16.11.2015 Views

studies

2015SupplementFULLTEXT

2015SupplementFULLTEXT

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

208A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1<br />

Two Combined Genetic Markers Identify Hepatitis C<br />

Recipients With a Lower Graft Survival From The Earliest<br />

Post-Transplant Period<br />

Silvia Martini 1 , Renato Romagnoli 2 , Francesco Tandoi 2 , Dominic<br />

Dell’Olio 3 , Paola Magistroni 3,4 , Francesca E. Bertinetto 3,4 , Ennia<br />

Dametto 3,4 , Stefano Mirabella 2 , Giorgia Rizza 2 , Donatella Cocchis<br />

2 , Antonio Ottobrelli 1 , Mario Rizzetto 1 , Mauro Salizzoni 2 ,<br />

Antonio Amoroso 3,4 ; 1 Gastrohepatology - Molinette Hospital,<br />

Turin, Italy; 2 Surgery, Liver Transplant Center - Molinette Hospital,<br />

Turin, Italy; 3 Regional Transplantation Center, Piedmont, Turin,<br />

Italy; 4 Immunogenetics Laboratory - Molinette Hospital, Turin, Italy<br />

Background and Aims. Human Leukocyte Antigen (HLA) variants<br />

and Interleukin 28B (IL28B) gene are associated with<br />

recurrence and progression of hepatitis C virus (HCV) liver<br />

disease in the transplant setting. We investigated the impact<br />

of these immunogenetic factors on graft survival after liver<br />

transplantation (LT) for HCV-related cirrhosis. Methods. This<br />

retrospective study consecutively included all the first adult LTs<br />

performed at a single center from 1999 to 2009. HLA-A/B/<br />

DRB1 frequencies were studied in 1,228 adult Caucasian LT<br />

recipients and their deceased donors. IL28B rs12979860<br />

polymorphism was determined only in the 449 HCV-positive<br />

viremic recipients and their HCV-negative donors, as it does<br />

not affect transplant outcome in non-HCV patients. Graft survival<br />

was the sole endpoint. Results. Median follow-up was 10<br />

years. HLA-DRB1*11 phenotypic frequency was significantly<br />

lower in HCV-positive recipients compared with HCV-negative<br />

recipients (28.2% vs 43.9%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 209A<br />

3<br />

Determinants and impact of early viral kinetics in<br />

patients with HCV-recurrence after liver transplantation<br />

treated by sofosbuvir + daclatasvir: results from the<br />

ANRS CO23 CUPILT study<br />

Camille Herve 1 , Audrey Coilly 2 , Claire Fougerou-Leurent 3 , Victor<br />

de Ledinghen 4 , Georges-Philippe Pageaux 5 , Pauline Houssel-Debry<br />

3 , Sylvie Radenne 6 , Christophe Duvoux 7 , Vincent Di<br />

Martino 8 , Nassim Kamar 9 , Louis d’Alteroche 10 , Valerie Canva 11 ,<br />

Pascal Lebray 12 , Jérôme Dumortier 13 , Christine Silvain 14 , Camille<br />

Besch 15 , Danielle Botta-Fridlund 16 , Albert Tran 17 , Helene Montialoux<br />

18 , Emilie Rossignol 3 , Alexandra Rohel 19 , Jean-Charles<br />

Duclos-Vallee 2 , Vincent Leroy 1 ; 1 Gastroenterology, CHU Grenoble,<br />

La Tronche, France; 2 Paul Brousse, Villejuif, France; 3 CHU<br />

Rennes, Rennes, France; 4 CHU Bordeaux, Pessac, France; 5 CHU<br />

Bordeaux, Bordeaux, France; 6 Croix Rousse, Lyon, France; 7 Henri<br />

Mondor, Créteil, France; 8 CHU Besançon, Besançon, France;<br />

9 CHU Toulouse, Toulouse, France; 10 CHU Tours, Tours, France;<br />

11 CHRU Lille, Lille, France; 12 La Pitié Salpétrière, Paris, France;<br />

13 Hôpital Edouard Herriot, Lyon, France; 14 CHU Poitiers, Poitiers,<br />

France; 15 CHU Strasbourg, Strasbourg, France; 16 Conception<br />

Hospital, Marseilles, France; 17 CHU Nice, Nice, France; 18 CHU<br />

Rouen, Rouen, France; 19 ANRS, Paris, France<br />

Early viral kinetic does not seem to impact the chance of SVR12<br />

with the use of new combinations of direct HCV antiviral agents.<br />

However, few data have been reported in HCV-recurrence after<br />

liver transplantation (LT). It could be hypothesized that immunosuppression<br />

might slow down HCV RNA decay. The aims<br />

of our study were to describe in a large cohort of treated LT<br />

patients the HCV RNA kinetics, to identify the predictive factors<br />

of slow response and to evaluate its impact on SVR. Methods:<br />

One hundred and ninety four LT patients prospectively included<br />

in the multi-centric CUPILT cohort and treated by sofosbuvir +<br />

daclatasvir ± ribavirin for 12 or 24 weeks were studied. HCV<br />

RNA was assessed at W0, W2, W4, W8, W12, EOT and<br />

FU12 by real time quantitative PCR with a lower limit of quantification<br />

of 15 IU/ml. Results: The characteristics of patients<br />

were as follows: age: 58.9 ± 8.7 years, baseline HCV RNA:<br />

6.4 log IU/ml, genotype 1: 79%, genotype 3: 11%, treatment<br />

experienced: 47%, HIV co-infection: 7%. The delay between LT<br />

and onset of treatment was 75.8 ± 69.1 months. The majority<br />

(60%) had severe HCV recurrence (F3, F4 or FCH). Immunosuppression<br />

was based on tacrolimus in 62%, cyclosporine in<br />

28% and everolimus in 8% of cases. MMF was administered<br />

in 57% of patients. Treatment duration was 24 weeks in 85%<br />

of cases and 12 weeks in 15% of cases. Ribavirin was used<br />

in 71%) of patients. The percentages of patients with HCV<br />

RNA ≥ 15 IU/ml and HCV RNA < 15 IU/ml but detectable<br />

were distributed as follow: W2: 83% and 13%, W4: 47% and<br />

29%, W8: 13% and 26%, W12 4% and 14%. By multivariate<br />

analysis, independent factors associated with HCV RNA ≥<br />

15 IU/ml at W4 were: use of MMF OR 1.90 (0.99 – 3.66),<br />

p3 treatments.<br />

Intention-to-treat 1- and 5-year overall survival was 84% and<br />

56%, respectively. Among 109 LT recipients, the 1- and 5-year<br />

post-LT survival was 95% and 80%, respectively, after a median<br />

post-LT follow-up of 4.3 years. The Kaplan-Meier probabilities<br />

of treatment failure at 1 and 5 years from first down-staging<br />

procedure were 25% and 44%, respectively. Treatment failure<br />

included dropout due to tumor progression (n=56), liver-related<br />

death without LT (n=12), and HCC recurrence after LT (n=12).<br />

In multivariable Cox proportional hazards regression analy-


210A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

sis, factors predicting treatment failure were pre-treatment AFP<br />

>1000 ng/mL (HR 3.3, p1000 ng/mL compared<br />

to 40% for AFP


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 211A<br />

and Status 1A FHF patients may be over-prioritized according<br />

to current allocation policies.<br />

Disclosures:<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

Patrick S. Kamath - Advisory Committees or Review Panels: Sequana Medical<br />

The following authors have nothing to disclose: Alina M. Allen, Julie Heimbach,<br />

Joseph J. Larson, Terry M. Therneau<br />

7<br />

End Stage Liver Disease Patients with MELD Scores<br />

Over 40 Have Significantly Higher Waitlist Mortality<br />

and Lower Probability of Receiving Liver Transplantation<br />

Compared to Status 1A Fulminant Hepatic Failure<br />

Patients<br />

Joseph C. Ahn 2 , Taft Bhuket 1 , Sasan Mosadeghi 1 , Catherine T. Frenette<br />

3 , Benny Liu 1 , Robert J. Wong 1 ; 1 Gastroenterology and Hepatology,<br />

Alameda Health System - Highland Hospital, Oakland, CA;<br />

2 School of Medicine, University of California, San Francisco, San<br />

Francisco, CA; 3 Organ Transplantation, Scripps Green Hospital,<br />

San Diego, CA<br />

Background: Current liver transplantation (LT) allocation systems<br />

prioritize Status 1A patients with fulminant hepatic failure<br />

(FHF) over end stage liver disease (ESLD) patients of all MELD<br />

scores. However, ESLD patients with the highest MELD scores<br />

may have higher waitlist mortality than Status 1A FHF patients,<br />

and thereby may require LT more urgently. Aim: Evaluate differences<br />

in LT waitlist mortality and probability of LT between Status1A<br />

FHF patients and ESLD patients with MELD>30. Methods:<br />

Using data from United Network for Organ Sharing registry,<br />

we retrospectively evaluated U.S. adults (age > 18) who were<br />

listed for LT from January 1, 2003 to December 31, 2013.<br />

ESLD patients were stratified into 3 groups of MELD scores (31-<br />

35, 36-40, > 40). Overall waitlist mortality and probability<br />

of LT were stratified by ESLD vs. Status 1A FHF and evaluated<br />

with Kaplan Meier curves and multivariate logistic regression<br />

models. The final multivariate model included adjustments for<br />

age, sex, race/ethnicity, obesity, hepatic encephalopathy,<br />

and ascites. Results: From 2003-2013, 15,049 ESLD patients<br />

with MELD>30 and 3,049 Status 1A FHF patients were listed<br />

for LT. ESLD patients with MELD 31-35 and MELD 36-40 had<br />

significant higher 28-day waitlist survival than Status 1A FHF<br />

patients, whereas MELD >40 patients had similar 28-day survival.<br />

Compared to Status 1A FHF patients, overall probability<br />

of LT was similar among ESLD patients with MELD 36-40 and<br />

MELD>40. On multivariate regression, compared to Status 1A<br />

FHF, ESLD patients with MELD scores >40 had significantly<br />

higher 14-day waitlist mortality (OR 1.92; 95% CI 1.56-2.36;<br />

p40 have significantly higher waitlist mortality compared<br />

to Status 1A FHF patients. Despite the higher waitlist mortality,<br />

ESLD patients with MELD >40 had similar probability of receiving<br />

LT, suggesting that these patients may be under-prioritized<br />

Disclosures:<br />

Catherine T. Frenette - Speaking and Teaching: Bayer, Salix, Gilead; Stock<br />

Shareholder: Gilead<br />

The following authors have nothing to disclose: Joseph C. Ahn, Taft Bhuket,<br />

Sasan Mosadeghi, Benny Liu, Robert J. Wong<br />

8<br />

Enhancing hepatocyte function using an advanced<br />

microfluidic system for clinical and pharmaceutical<br />

applications<br />

Maria Navarro-Zornoza 1,2 , Xavi Illa 2 , Carmen Peralta 3 , Rosa<br />

Villa 2 , Jordi Gracia-Sancho 1 ; 1 Barcelona Hepatic Hemodynamic<br />

Lab, IDIBAPS - Hospital Clinic de Barcelona - CIBEREHD, Barcelona,<br />

Spain; 2 CNM-CSIC, Barcelona, Spain; 3 IDIBAPS - CIBEREHD,<br />

Barcelona, Spain<br />

Background and Aims: Development of an advanced liveron-a-chip<br />

to support liver function in patients with acute liver<br />

failure and for in vitro drug discovery is still challenging. We<br />

hypothesized that representation of the unique architectural<br />

and dynamic features of the hepatic sinusoid using a superior<br />

microfluidic bioreactor system for co-culture of primary liver<br />

sinusoidal endothelial cells (LSECs) and hepatocytes would<br />

mimic the in vivo hepatic metabolism significantly better than<br />

using conventional methods for hepatocytes mono-culture or<br />

co-cultures. Methods: Hepatocytes and LSECs freshly isolated<br />

from Wistar rats were co-cultured in a bioreactor recently<br />

developed by our team. LSECs were grown in the upper area<br />

of the bioreactor on a 1mm pore size membrane with homogeneous<br />

and continuous shear stress stimulation (3dyn/cm 2 );<br />

whereas hepatocytes were plated in the lower area. After<br />

48-72h of co-culture, hepatocyte function and phenotype<br />

were characterized as albumin synthesis (enzymatic method),<br />

urea production (BUN enzymatic method), CYP3A4 (luminescence<br />

assay) and HNF4a expression (qPCR), and compared<br />

to hepatocytes co-cultured with LSECs in the microfluidic system<br />

but under static conditions or in conventional transwells,<br />

and with hepatocytes cultured alone. Results: Hepatocytes<br />

co-cultured in the bioreactor exhibited markedly better phenotype<br />

than those cultured using conventional methods. Indeed,<br />

hepatocytes from the bioreactor maintained polygonal shape<br />

differentiation, had higher albumin and urea production (Albumin:<br />

67-fold higher vs. transwells; 108-fold vs. mono-cultures;<br />

Urea: 79-fold superior vs. transwells; 108-fold vs. mono-cultures),<br />

exhibited markedly higher HNF4a expression (17-fold<br />

vs. transwells; 12-fold vs. mono-cultures), and superior CYP3A4<br />

activity (28-fold vs. transwells; 17-fold vs. mono-cultures). All<br />

differences were p


212A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

actions condition). Conclusion: The herein presented bioreactor<br />

allows co-culture of LSECs and hepatocytes, maintaining and<br />

enhancing hepatocyte function long-term significantly better<br />

than using monocultures or conventional co-culture methods.<br />

Our approach provides an overwhelming insight in the importance<br />

of the hepatic sinusoid in the development of technology<br />

for liver support in acute liver failure and for drug discovery.<br />

Disclosures:<br />

The following authors have nothing to disclose: Maria Navarro-Zornoza, Xavi<br />

Illa, Carmen Peralta, Rosa Villa, Jordi Gracia-Sancho<br />

9<br />

Prediction of waitlist mortality in patients with portopulmonary<br />

hypertension (POPH): An analysis of the UNOS<br />

database<br />

Hilary M. DuBrock 1,2 , David S. Goldberg 4 , Norman L. Sussman 5 ,<br />

Sonja Bartolome 6 , Zakiyah Kadry 7 , Reena Salgia 8 , Andre M.<br />

De Wolf 9 , David C. Mulligan 10 , Steven M. Kawut 4 , Michael J.<br />

Krowka 3 , Richard Channick 2 ; 1 Medicine, Beth Israel Deaconess<br />

Medical Center, Boston, MA; 2 Medicine, Massachusetts General<br />

Hospital, Boston, MA; 3 Mayo Clinic, Rochester, MN; 4 University<br />

of Pennsylvania, Philadelphia, PA; 5 Baylor College of Medicine,<br />

Houston, TX; 6 UT Southwestern, Dallas, TX; 7 Penn State Hershey<br />

Medical Center, Hershey, PA; 8 Henry Ford Hospital, Detroit, MI;<br />

9 Northwester University, Chicago, IL; 10 Yale New Haven Hospital,<br />

New Haven, CT<br />

Purpose: The current United Network for Organ Sharing<br />

(UNOS) system grants Model for End-Stage Liver Disease<br />

(MELD) exception points to patients with POPH and a mean<br />

pulmonary artery pressure (mPAP) 25 and pulmonary vascular<br />

resistance (PVR)>240] was built using a competing risk subdistribution<br />

hazard model and purposeful selection. Significance<br />

was defined as p16 years of age) were approved for<br />

a POPH MELD exception. In patients with true POPH (n=190),<br />

13(6.8%) died on the waitlist, 4(2.1%) died during transplant,<br />

27(14.2%) were removed for being too sick, 100(52.6%)<br />

patients underwent deceased donor liver transplant, 2(1.1%)<br />

underwent living donor transplant, 9(4.7%) were removed for<br />

other reasons and 35(18.4%) remained active on the waitlist.<br />

Initial MELDNa was a significant univariate predictor of waitlist<br />

mortality (HR 1.121, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 213A<br />

1%. Conclusion. In a national cohort of HCC patients undergoing<br />

LT, the recently enacted UNOS criteria improved the accuracy<br />

of radiographic staging but did not reduce false-positive<br />

HCC diagnoses. Overall utilization of HCC MELD exceptions<br />

did not meaningfully change due to a shift from automatic T2<br />

to RRB-approved exceptions.<br />

Disclosures:<br />

Jesse M. Civan - Consulting: Merck<br />

The following authors have nothing to disclose: Barry Schlansky, Willscott E.<br />

Naugler<br />

11<br />

Treatment with Optifast Reduces Hepatic Steatosis and<br />

Safely Increases Candidacy Rates for Live Donor Liver<br />

Transplantation<br />

Adam Doyle 1 , Jayne Dillon 1 , Oyedele Adeyi 2 , Sandra E. Fischer 2 ,<br />

Meaghan MacArthur 1 , Max Marquez 1 , Robert Smith 1 , David<br />

Grant 1 , Mark S. Cattral 1 , Ian McGilvray 1 , Paul Greig 1 , Anand<br />

Ghanekar 1 , Markus Selzner 1 , Eberhard L. Renner 1 , Les Lilly 1 , Gary<br />

Levy 1 , Nazia Selzner 1 ; 1 Multi Organ Transplant Program, University<br />

Health Network, Toronto, ON, Canada; 2 Department of<br />

Pathology, University Health Network, Toronto, ON, Canada<br />

Background: As a consequence of the increased prevalence<br />

of obesity and associated liver steatosis, the volunteer pool for<br />

live donor liver transplantation is threatened. The Optifast verylow<br />

calorie diet meal replacement system has demonstrated<br />

efficacy in the reduction in obesity and hepatic steatosis in<br />

candidates for bariatric surgery. We studied the ability of<br />

a 6-8 week course of Optifast in potential donors to reduce<br />

body weight and fatty liver converting them to suitable live<br />

liver donors. Methods: Among 447 potential donors evaluated<br />

in our live donor liver transplant program between January<br />

2012 to December 2014, 47 potential donors with a BMI<br />

greater than 30, or evidence of significant steatosis on liver<br />

biopsy, were administered Optifast for 6-8 weeks during their<br />

evaluation for live liver donation. Outcomes including changes<br />

in BMI and degree of steatosis on liver biopsy before and<br />

after Optifast treatment were studied. Post-surgical outcomes<br />

for donors and recipients were also recorded, and compared<br />

to 91 live donors who did not receive Optifast and underwent<br />

donor hepatectomy within the study period. Results: Median<br />

BMI of Optifast potential donors (n=47) was 35.0 (IQR 33.1-<br />

37.3), compared to 25.9 (IQR 23.1-29.1) in all non-Optifast<br />

potential donors (n=400) [p


214A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

13<br />

A simple educational tool for reducing 30-day hospital<br />

readmissions in patients with decompensated cirrhosis<br />

Dennis Kumral, Michael W. Crothers, Stephen H. Caldwell, Zachary<br />

Henry; University of Virginia, Charlottesville, VA<br />

Background: Cirrhosis is a chronic disease with acute on<br />

chronic decompensating events that lead to hospitalization,<br />

similar to congestive heart failure (CHF) and chronic obstructive<br />

pulmonary disease (COPD). While patients with CHF and<br />

COPD have widespread targeted discharge pathways that<br />

help prevent readmissions, patients with cirrhosis have no<br />

standardized discharge programs. We developed a quality<br />

improvement initiative aimed at utilizing a standardized discharge<br />

pathway for patients with cirrhosis. Our goal was to<br />

reduce 30-day readmissions to our inpatient hepatology service<br />

by at least ten percent. Methods: In April 2015 we began a<br />

quality improvement discharge education program for patients<br />

with cirrhosis complicated by either fluid overload or hepatic<br />

encephalopathy. A dedicated hepatology nurse coordinator<br />

performed a teaching session with patients and their families<br />

prior to discharge and provided them with a patient-friendly<br />

cirrhosis management booklet created by the study team. All<br />

patients were provided with a digital scale and pill organizer<br />

if they did not already own one. Within 72-hours of discharge,<br />

a phone call was made to patients to review medications and<br />

reinforce teaching. Patients were monitored for readmission up<br />

to 30 days post discharge as the primary quality outcome. A<br />

control group of cirrhosis patients without a dedicated teaching<br />

program was chosen from hospital discharges during the same<br />

time period one year earlier. Thirty-day readmission rates were<br />

assessed for comparison. Results: A total of 20 patients with<br />

cirrhosis complicated by fluid overload and/or hepatic encephalopathy<br />

went through the discharge pathway and reached<br />

30 days post discharge at the time of this submission. We<br />

identified 25 control patients from the same time period one<br />

year earlier. There was no significant difference in age, gender,<br />

discharge MELD, discharge child pugh score, or length<br />

of stay between the two groups. The 30 day readmission rate<br />

in the study group was 25% compared to 62% in the control<br />

group, p=0.02. This correlates to a relative risk reduction of<br />

60% with a number needed to teach (NNT) of 2.7 to prevent<br />

one 30-day readmission. On multivariate logistic regression<br />

the only significant variable for predicting 30 day readmission<br />

was participating in the educational discharge pathway which<br />

was found to be protective, OR 0.22, p=0.02. Conclusions:<br />

Implementation of a dedicated educational discharge pathway<br />

with detailed teaching of cirrhosis management and a post-discharge<br />

follow-up phone call can reduce 30-day readmissions<br />

for patients with decompensated cirrhosis.<br />

Disclosures:<br />

Stephen H. Caldwell - Advisory Committees or Review Panels: Vital Therapy;<br />

Grant/Research Support: Genfit, Gilead Sciences, Immuron, Hyperion, Immuron,<br />

NGM<br />

The following authors have nothing to disclose: Dennis Kumral, Michael W.<br />

Crothers, Zachary Henry<br />

14<br />

Development of a Model to Predict Post-Surgical<br />

Unplanned Readmissions in Patients with Decompensated<br />

Cirrhosis<br />

Monica Schmidt 1 , Paul H. Hayashi 2 , Alfred S. Barritt 2 ; 1 UNC Liver<br />

Center & Gillings School of Global Public Health, University of<br />

North Carolina Chapel Hill, Chapel Hill, NC; 2 UNC Liver Center,<br />

Chapel Hill, NC<br />

Background:In the U.S., liver cirrhosis is expected to affect more<br />

than 1 million individuals by 2020 and many will require surgery.<br />

Prediction of 30-day readmission will be key in designing<br />

appropriate discharge planning and preventing readmissions.<br />

We developed a predictive model for 30-day readmission that<br />

is specific to patients with decompensated liver cirrhosis undergoing<br />

surgery. Methods:We used the National Surgical Quality<br />

Improvement Program data (NSQIP) from 2011-2013. Patients<br />

with cirrhosis were identified. Predictions were obtained for<br />

the model and the MELD score for comparison. The AUROC,<br />

cut-point, sensitivity/specificity and decision curve analysis<br />

are reported. Results: There were 5,879 patients with 739<br />

readmissions within 30-days (13%). Predictors of 30-day<br />

readmission were insulin dependent diabetes (OR 3.4 CI 1.6-<br />

7.3), discharged home (OR 3.4 CI 1.5-8.7), ASA class 4-life<br />

threatening (OR 4.7 CI 1.8-12.1), and days from surgery to<br />

discharge (OR 0.94 CI 0.91-0.97). The AUROC for the MELD<br />

score was 0.50 while our model reached 0.80 (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 215A<br />

15<br />

Effect of HCV treatment outcome on hospitalization rate:<br />

Chronic Hepatitis Cohort Study (CHeCS)<br />

Eyasu H. Teshale 1 , Jian Xing 1 , Anne C. Moorman 1 , Scott D. Holmberg<br />

1 , Stuart C. Gordon 2 , Loralee B. Rupp 2 , Mei Lu 2 , Philip R.<br />

Spradling 1 , Joseph A. Boscarino 3 , Connie M. Trinacty 4 , Mark A.<br />

Schmidt 5 , Fujie Xu 1 ; 1 CDC, Atlanta, GA; 2 Henry Ford Health System,<br />

Detroit, MI; 3 Geisinger Health System, Danville, PA; 4 Kaiser<br />

Permanente Hawaii, Honolulu, HI; 5 Kaiser Permanente Northwest,<br />

Portland, OR<br />

Background: Chronic hepatitis C virus (HCV) infection is associated<br />

with high all-cause hospitalization rates. Hospitalization<br />

rates increase with increasing severity of liver disease. Successful<br />

treatment of chronic HCV infection results in reduced<br />

morbidity and mortality. The objective of this study is to determine<br />

the association between HCV treatment outcome and<br />

hospitalization rate. Methods: We used data from HCV-infected<br />

patients seen in four large U.S. healthcare systems from<br />

2006-2012. We included persons who received HCV treatment<br />

during this time and who had follow up before and after<br />

completion of treatment. Treatment outcome was grouped as<br />

sustained virologic response (SVR) or treatment failure (TF). We<br />

determined hospitalization rates [per 100 person years (PY)]<br />

before and after treatment (SVR or TF) and excluded hospitalizations<br />

during treatment. Results: Between 2006 and 2012,<br />

1409 persons received HCV therapy; 680 (48.3%) achieved<br />

SVR and 729 (51.7%) had TF. The overall before treatment<br />

hospitalization rate was 13.2/100 PY; before treatment hospitalization<br />

rates for those who achieved SVR and those with<br />

TF were 10.6/100 PY and 16.0/100 PY, respectively. The<br />

after treatment hospitalization rate decreased to 5.6/100 PY<br />

for those who achieved SVR (p


216A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

17<br />

A Prospective Study of a Protocol to Reduce Early Readmission<br />

after Liver Transplantation<br />

Mark W. Russo 1 , Amit Mori 1 , David Levi 2 , Ruth Pierce 2 , Siddesh<br />

Besur 1 , Vincent P. Casingal 2 , Paul A. Schmeltzer 1 , Philippe J.<br />

Zamor 1 , Andrew Delemos 1 , Lon Eskind 2 ; 1 Hepatology, Carolinas<br />

Healthcare System, Charlotte, NC; 2 Transplant, Carolinas Medical<br />

Center, Charlotte, NC<br />

A Prospective Study of a Protocol to Reduce Early Readmission<br />

after Liver Transplant Introduction: Hospital readmission<br />

is an important marker of quality and delivery of care. Studies<br />

have evaluated risk factors for readmission after liver transplant<br />

(OLT) but few <strong>studies</strong> evaluated interventions to reduce<br />

readmission. Aim: To reduce 30 day readmission rates after<br />

OLT by implementing a prospective protocol with a multistep<br />

strategy. Methods: In October 2013 a comprehensive strategy<br />

was initiated to address early readmission after OLT. Core<br />

components of the protocol included revising criteria for readmission,<br />

alternatives to readmission, emphasized processes to<br />

improve patient teaching and discharge planning and expansion<br />

of outpatient services. 2014 served as the first full year<br />

the protocol was implemented. Readmission rates for 2014<br />

were compared to 2012 and 2013 with Fisher’s exact test and<br />

ANOVA. Logistic regression was used to control for potential<br />

confounding variables. Results: From January 1st 2012 thru<br />

December 31st 2014 167 adult OLTs were performed at our<br />

center with a mean biologic MELD of 21. The most common<br />

indications were HCV (35%), NAFLD (18%), HCC (17%), and<br />

ETOH (13%). The mean age of the study group was 54 y/o,<br />

63% were male. The mean ICU LOS was 3.4 days and hospital<br />

LOS 12.6 days. Over the study period 30 day readmission<br />

rate was 34% and decreased after implementing the readmission<br />

protocol from 40% in 2012 and 39% in 2013 to 19%<br />

in 2014, p=0.04. The most common reasons for readmission<br />

were biliary complications, infection, and rejection. Discharge<br />

year (before or after implementing protocol) remained significant<br />

after controlling for age, MELD score, indication, dialysis<br />

pre and post liver transplant, distance from transplant center,<br />

length of stay, PRBC transfusion, insurance and weekend discharge<br />

OR=0.40 [95% CI 017,0.94], p=0.04. The top factors<br />

identified as reducing readmission were expanding outpatient<br />

services and alternatives to inpatient readmission. Conclusions:<br />

Early readmission after liver transplantation can be reduced<br />

by expanding outpatient services and implementing alternative<br />

approaches to inpatient readmission.<br />

Disclosures:<br />

Mark W. Russo - Grant/Research Support: Merck, Salix; Speaking and Teaching:<br />

janssen, Gilead, ABBVIE, Salix<br />

Philippe J. Zamor - Grant/Research Support: AbbVie, Bristol Myers Squibb,<br />

Gilead, Merck & Co. ; Speaking and Teaching: AbbVie, Bristol Myers Squibb,<br />

Janssen<br />

The following authors have nothing to disclose: Amit Mori, David Levi, Ruth<br />

Pierce, Siddesh Besur, Vincent P. Casingal, Paul A. Schmeltzer, Andrew Delemos,<br />

Lon Eskind<br />

18<br />

Reduced Work Productivity (WP), Absenteeism and<br />

Presenteeism of Patients Infected with Hepatitis C Virus<br />

(HCV) are Independently Predicted by Physical Component<br />

of Patient-Reported Outcomes (PROs)<br />

Zobair M. Younossi 1,2 , Maria Stepanova 3 , Linda Henry 3 , Issah<br />

Younossi 3 , Ali A. Weinstein 3 , Fatema Nader 1 , Sharon L. Hunt 3 ;<br />

1 Center For Liver Disease, Department of Medicine, Inova Fairfax<br />

Hospital, Falls Church, VA; 2 Betty and Guy Beatty Center for Integrated<br />

Research, Inova Health System, Falls Church, VA; 3 Center<br />

for Outcomes Research in Liver Disease, Washington, DC<br />

Background: Hepatitis C virus (HCV) infection is associated with<br />

significant impairment of health-related quality of life (HRQOL)<br />

and other PROs. Additionally, HCV infected patients are known<br />

to have reduced work productivity (WP), both in terms of presenteeism<br />

(impairment in work productivity while working) and<br />

absenteeism (productivity loss due to absence from work). Aim:<br />

The aim of this study was to identify PROs which are predictive<br />

of WP in untreated HCV-infected patients. Methods: We<br />

analyzed data from HCV patients prior to intitaion of anti-<br />

HCV regimens in a clinical trial setting. All patients had completed<br />

4 PRO questionnaires (CLDQ-HCV, SF-36, FACIT-F and<br />

WPAI:SHP). In subjects who reported being employed, WP<br />

and its absenteeism and presenteeism components were calculated<br />

using WPAI:SHP instrument. Independent PRO predictors<br />

of WP, absenteeism and presenteeism were assessed using<br />

multiple linear regression. Results: Of 4,121 HCV patients who<br />

completed pre-treatment WPAI:SHP, 2,480 (60.2%) reported<br />

to be employed, and of those, 2,121 had completed all PRO<br />

questionnaires before treatment initiation. The study cohort was<br />

51.1±9.6 years old, 16.7% cirrhotic, 76.1% HCV genotype 1,<br />

and 62.3% treatment-naïve. Average baseline impairment in<br />

WP was 11.0%, including 8.3% (75% of WP impairment) of<br />

presenteeism and 2.7% (25% of WP impairment) of absenteeism.<br />

Using multivariate regression analysis, WP was predicted<br />

by the activity/energy domain of CLDQ-HCV (beta=4.55±0.56<br />

per 1 point, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 217A<br />

19<br />

Administration of an Ileal Apical Sodium-dependent Bile<br />

Acid Transporter Inhibitor Protects Against Non-alcoholic<br />

Fatty Liver Disease in High Fat Diet-fed Mice<br />

Anuradha Rao 1 , Astrid Kosters 1 , Jamie Mells 1 , Bradley T. Keller 4 ,<br />

Hong Yin 2 , Sophia Banton 2 , Shuzhao Li 2 , Dean P. Jones 2 , Hao<br />

Wu 3 , Paul Dawson 1 , Saul J. Karpen 1 ; 1 Pediatrics, Emory University,<br />

Atlanta, GA; 2 Medicine, Emory University, Atlanta, GA; 3 Public<br />

Health, Emory University, Atlanta, GA; 4 Vasculox, Inc, St. Louis,<br />

MO<br />

Non-alcoholic fatty liver disease (NAFLD) is the most common<br />

chronic liver disease in the Western world. The mechanisms<br />

underlying NAFLD development and progression are not<br />

fully elucidated, and safe and effective NAFLD therapies are<br />

needed. Bile acids (BA) and their receptors have important<br />

roles in regulating whole body metabolism, including multiple<br />

hepatic targets coordinating lipid handling. We hypothesized<br />

that interruption of the enterohepatic BA circulation via<br />

a non-absorbable Apical Sodium-dependent BA Transporter<br />

(ASBT) inhibitor (ASBTi; SC435) would modify signaling in<br />

the gut-liver axis and diminish the development of NAFLD in<br />

High Fat Diet (HFD) fed mice. Methods: Male C57Bl/6 mice<br />

(n=12-16/group) were fed for 16 weeks with A) chow, B) HFD<br />

composed of 45% fat and 0.2% cholesterol, with 4% sucrose<br />

water (HFD), and C) HFD plus ASBTi (SC435; 60 ppm; HFD/<br />

ASBTi). Body weight, food intake, glucose and insulin tolerance,<br />

liver histology, and liver lipids were measured. Molecular<br />

mechanisms were examined using RNASeq, metabolomics,<br />

and real-time PCR analyses. Results: In HFD versus HFD/ASBTi<br />

mice, administration of the ASBTi increased fecal BA excretion,<br />

and mRNA expression for colonic Ibabp (7.0±2.0 fold;<br />

p


218A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Inflammasomes are new molecular platforms that respond to<br />

microbial products through the synthesis of pro-inflammatory<br />

cytokines. The Nlrp3 inflammasome activation is known to<br />

determine Il-1β and Il-18 synthesis and release, thereby modulating<br />

cell biology and eliciting pro-inflammatory signals in<br />

different cell types. We sought to verify whether the inflammasome<br />

is activated in sclerosing cholangitis and which are its<br />

effects in reactive cholangiocytes. Methods: Expression levels<br />

of the NLRP3 inflammasome were tested in cholangiocytes of<br />

normal and cholestatic livers. The effects of Nlrp3 activation<br />

induced by incubation with LPS/ATP were studied in vitro in<br />

normal and siRNA Nlrp3 knockdown cholangiocytes. In vivo,<br />

wild type (WT) and Nlrp3 A350VneoR<br />

(Nlrp3 -/- ) mice were fed<br />

with 3,5-diethoxycarbonyl-1,4-dihydrocollidine (DDC, a model<br />

of sclerosing cholangitis) for 4 weeks. Results: Expression of<br />

Nlrp3 and its components is increased in in cholangiocytes<br />

of mice subjected to DDC and in patients affected by PSC<br />

compared to normal conditions. LPS/ATP-induced activation<br />

of Nlrp3 in cholangiocytes stimulates the expression of Il-18<br />

and Il-6, but not of Il-1β. Nlrp3 activation also significantly<br />

decreased the expression of Zonulin-1 and E-cadherin but<br />

had no effect on cholangiocyte proliferation. The effects of<br />

LPS/ATP-induced activation of Nlrp3 in cholangiocytes were<br />

neutralized by knockdown of Nlrp3 by siRNA. In vivo, the<br />

increases in the liver CK-19-positive parenchyma induced by 4<br />

week DDC in WT animals were reduced in Nlrp3 -/- mice and<br />

expression of Zonulin-1 tended to be re-established. Conclusions:<br />

Nlrp3 is expressed in reactive cholangiocytes, both in<br />

murine models and in PSC. The activation of Nlrp3 leads to<br />

the synthesis of pro-inflammatory cytokines and influences the<br />

epithelial integrity of cholangiocytes. These findings suggest<br />

that microbial products may participate to the development of<br />

certain cholangiopathies by activating the inflammasome in<br />

cholangiocytes and altering the barrier function of the biliary<br />

epithelium.<br />

Disclosures:<br />

The following authors have nothing to disclose: Luca Maroni, Laura Agostinelli,<br />

Stefania Saccomanno, Eleonora Mingarelli, Chiara Rychlicki, Samuele De Minicis,<br />

Jesus M. Banales, Antonio Benedetti, Gianluca Svegliati Baroni, Marco<br />

Marzioni<br />

22<br />

Taurocholate Activates YAP via Sphingosine-1 Phosphate<br />

Receptor 2 in Cholangiocytes<br />

Kesong Peng 1 , Yunzhou Li 2 , Runping Liu 1,2 , Jing Yang 1 , Yongqing<br />

Wang 3,1 , Guang Liang 4 , Phillip B. Hylemon 1,2 , Huiping Zhou 1,2 ;<br />

1 Microbiology and Immunology, Virginia Commonwealth University,<br />

Richmond, VA; 2 McGuire Veterans Affairs Medical Center,<br />

Richmond, VA; 3 Nanjing Medical University, Nanjing, China;<br />

4 School of Pharmacy, Wenzhou Medical University, Wenzhou,<br />

China<br />

Introduction: The transcription co-activators YAP/TAZ are the<br />

core components of the Hippo pathway. Recent <strong>studies</strong> demonstrated<br />

that activation of several G protein coupled receptors<br />

(GPCRs), including sphingosine-1-phosphate receptor 2<br />

(S1PR2), causes dephosphorylation and nuclear translocation<br />

of YAP. Cholangiocytes are continuously exposed to high concentrations<br />

of conjugated bile acids. Obstruction of bile flow<br />

is a potent stimulus for cholangiocyte proliferation. We have<br />

recently reported that taurocholate (TCA) promotes cholangiocyte<br />

proliferation via activation of S1PR2, which is the predominant<br />

S1P receptor in cholangiocytes. Recently, it was reported<br />

that bile acids promote liver carcinogenesis through activation<br />

of YAP. However, whether activation of S1PR2 by conjugated<br />

bile acids induces YAP activation in cholangiocytes has not<br />

been examined and is the focus of this study Methods: A mouse<br />

cholangiocyte cell line was used for in vitro <strong>studies</strong>. Bile duct<br />

ligation (BDL) mouse models were used to examine the role<br />

of S1PR2 in YAP activation in vivo. A chemical antagonist of<br />

S1PR2, JTE-013, was used to inhibit S1PR2 activation. S1P<br />

was used as positive control. YAP activation was determined<br />

by Western blot analysis. The mRNA levels of the YAP target<br />

genes were determined by real-time RT-PCR. Results: Both<br />

TCA- and S1P-induced cell migration and proliferation were<br />

inhibited by JTE-013 in cholangiocytes. TCA- and S1P-induced<br />

activation of YAP was indicated by decreasing of phosphorylated<br />

YAP in cytosol and increasing of nuclear translocation of<br />

YAP, which was markedly inhibited by JTE-013. In addition,<br />

TCA and S1P also significantly up-regulated the expression<br />

of the down-stream target genes of YAP including: IQGAP1,<br />

cyclin D1, p21 and CTGF (connective tissue growth factor),<br />

in cholangiocytes. YAP activation was significantly reduced in<br />

BDL S1PR2 -/- mice as compared to wild-type mice. Discussion/<br />

Conclusion: Elevated conjugated bile acid levels are correlated<br />

with cholangiocyte proliferation and cholangiocarcinoma by<br />

unknown mechanism(s). YAP-activation has been linked to<br />

tumour growth. This study identified TCA/S1PR2/YAP/TAZ as<br />

a cell signalling pathway involved in TCA-mediated cholangiocyte<br />

proliferation and possibly transformation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kesong Peng, Yunzhou Li, Runping<br />

Liu, Jing Yang, Yongqing Wang, Guang Liang, Phillip B. Hylemon, Huiping<br />

Zhou<br />

23<br />

SRT1720 protects against cholestatic liver injury in mice<br />

by altering hepatic bile acid composition and enhancing<br />

urinary excretion of bile acids<br />

Supriya Kulkarni 1 , Carol J. Soroka 1 , Lee R. Hagey 2 , James L.<br />

Boyer 1 ; 1 Liver Center, Internal Medicine, Digestive Diseases, Yale<br />

University School of medicine, New Haven, CT; 2 Division of Gastroenterology,<br />

Department of Medicine, University of California at<br />

San Diego, La Jolla, CA<br />

BACKGROUND: Sirtuin1 (SirT1, mammalian homolog of S. Cerevisiae<br />

enzyme Sir2) is a critical transcriptional and transactivational<br />

regulator of murine Fxr. Liver specific knockout of Sirt1<br />

deregulates Fxr activity. Our previous <strong>studies</strong> demonstrated that<br />

bile duct ligation (BDL), or cholic acid (CA) feeding decrease<br />

hepatic Sirt1 expression and that a Sirt1 activator SRT1720<br />

improves cholestatic liver injury. AIM: To determine the mechanisms<br />

by which Sirt1 activation may alleviate cholestatic liver<br />

injury. METHODS: C57Bl/6 mice (n=11-13, male, 8-9 weeks<br />

old) were fed standard rodent chow or chow supplemented<br />

with 1% CA for 5 days. Both groups (n=5-6) were administered<br />

either vehicle or Sirt1 activator, SRT1720 (50mg/kg/<br />

day) for duration of the diet. Liver injury (ALT, ALP), BA levels in<br />

plasma, liver, ileum, urine and the bile acid pool size as well<br />

as gene and protein expression of BA transporter and synthesis<br />

genes were determined. Immunoprecipitation assays were<br />

used to determine Sirt1 acetylation and Fxr activity. RESULTS:<br />

Sirt1 mRNA, protein and acetylation activity were significantly<br />

reduced (75%) in CA fed mouse livers. SRT1720 administration<br />

in CA fed mice reduced plasma ALT (40%) and plasma<br />

BA (50%) levels while hepatic and total BA pool remained<br />

unchanged. SRT1720 significantly increased tetra- hydroxylated<br />

BA by 1.96x and decreased di-hydroxylated BA fraction<br />

to 21% over CA fed mice and increased Cy2b10 (1.75x)<br />

expression. SRT1720 increased ileal Fgf15 (30x) and hepatic<br />

Fgfr4 (2X), decreased hepatic Cyp7a1 (99%), 27a1 (75%)<br />

expression. SRT1720 also increased renal Mrp2 and Mrp4<br />

expression (2-4.5x) along with increased urinary BA concen-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 219A<br />

tration (2x) relative to CA fed mice, thus increasing BA excretion<br />

in the urine. Hepatic Jnk was activated by CA feeding,<br />

a finding that that was reversed by SRT1720 administration<br />

which restored Sirt1 and Fxr mRNA, protein and activity levels<br />

to normal. CONCLUSION: Sirt1 protein levels and acetylation<br />

activity are decreased in CA feeding model of cholestasis Sirt1<br />

activator SRT1720 decreased cholestatic liver injury by several<br />

mechanisms: SRT1720 increased the hepatic concentration of<br />

hydrophilic hydroxylated BA likely via increased expression<br />

of Cyp2b10. In addition, SRT1720 decreased BA synthesis<br />

via the Fgf15-Fgfr-Cyp7a1 pathway and finally SRT1720<br />

decreased serum bile acid levels by increasing renal Mrp2<br />

and Mrp4 expression and urinary BA excretion. These observations<br />

could be attributed to SRT1720 mediated reversal of Jnk<br />

activation and hence restoration of Sirt1 and Fxr expression in<br />

the CA fed mice.<br />

Disclosures:<br />

James L. Boyer - Advisory Committees or Review Panels: Pfizer; Consulting:<br />

abbvie<br />

The following authors have nothing to disclose: Supriya Kulkarni, Carol J.<br />

Soroka, Lee R. Hagey<br />

24<br />

FXR-β-catenin complex and bile acid homeostasis: therapeutic<br />

implications in cholestatic liver disease<br />

Kari Nejak-Bowen, Satdarshan (Paul) S. Monga; University of Pittsburgh,<br />

Pittsburgh, PA<br />

Cholestatic liver diseases are characterized by bile acid (BA)<br />

accumulation in the liver, which can lead to inflammation<br />

and fibrosis, and eventually result in end-stage liver disease.<br />

We utilize the bile duct ligation (BDL) model and identify a<br />

dramatic reduction in liver injury and fibrosis in liver-specific<br />

β-catenin knockout (KO) mice. This is due to a previously unrecognized<br />

role of β-catenin in regulation of BA synthesis and<br />

transport through association with farnesoid X receptor (FXR).<br />

We confirm this association through immunoprecipitation (IP)<br />

of FXR and β-catenin from wild-type (WT), β-catenin KO, and<br />

FXR KO livers and showed association of the two proteins in<br />

WT but not in either KO. Further, the association with FXR<br />

occurs at the C-terminal of β-catenin, since IP of β-catenin and<br />

FXR from HepG2 cells reveals association with both the fulllength<br />

and N-terminal truncated β-catenin forms. Additionally,<br />

the β-catenin-FXR pool is unresponsive to BA stimulation as<br />

demonstrated by IP <strong>studies</strong> on GW4064-treated Hep3B cells.<br />

To address the functional relevance of this complex, we show<br />

that absence of β-catenin allows FXR activation to occur more<br />

readily, both in vitro and in vivo. In fact, a more pronounced<br />

FXR activation in KO is the major reason of decreased total<br />

hepatic BA and injury after BDL. Conversely, when Hep3B cells<br />

were transfected with S33Y-mutated stable β-catenin, we find<br />

increased FXR-β-catenin association by IP. This coincides with<br />

loss of FXR-RXRα association, which normally causes activation<br />

of SHP and suppresses BA synthesis. Expression of S33Y-βcatenin<br />

also abrogates the GW4064-induced increase in SHP<br />

reporter activity seen in Hep3B cells treated with control plasmid,<br />

confirming the inhibitory role of this association. This was<br />

validated in HepG2 cells, which show less SHP induction after<br />

β-catenin silencing than Hep3B cells, as the truncated/stable<br />

form of β-catenin is still present in these cells after β-catenin suppression.<br />

The complex interplay between β-catenin, FXR and<br />

RXRα was further validated when overexpression of RXRα led<br />

to a switch in association of FXR from β-catenin to RXRα. Thus,<br />

we have identified a novel FXR-β-catenin complex and propose<br />

its manipulation may provide a novel therapeutic opportunity<br />

for alleviating BA-associated hepatic injury during cholestasis.<br />

Disclosures:<br />

Satdarshan (Paul) S. Monga - Advisory Committees or Review Panels: Abbvie;<br />

Grant/Research Support: Dicerna Pharmaceuticals<br />

The following authors have nothing to disclose: Kari Nejak-Bowen<br />

25<br />

HIV and HCV co-infection in hepatocytes and stellate<br />

cells reveals cooperative transcriptional activation<br />

between viruses and cell types<br />

Shadi Salloum 1 , Cynthia Brisac 1 , Rohit Jindal 2 , Shyam Sundhar<br />

Bale 2 , Nadia Alatrakchi 1 , Annie Kruger 1 , Anna Lidofsky 1 , Martin<br />

L. Yarmush 2 , Raymond T. Chung 1 ; 1 GI, MGH, Boston, MA; 2 Center<br />

for Engineering in Medicine, Shriner, Boston, MA<br />

Background: HIV/HCV co-infection accelerates progression to<br />

liver fibrosis, a major public health problem. Currently, the<br />

molecular and cellular mechanisms underlying progression of<br />

this dynamic disease remain incompletely understood. We previously<br />

demonstrated, in monoculture experiments, that HCV<br />

and HIV independently led to the production of profibrogenic<br />

TGFβ through an increase in reactive oxygen species (ROS)<br />

and NFkB activation in hepatocytes as well as in hepatic stellate<br />

cells (HSCs), all of which are associated with the development<br />

of hepatic fibrosis. Aim: We sought to create a co-culture<br />

model combining hepatocytes and HSCs to effectively simulate<br />

HCV and HIV co-infection in vitro. This model will be of<br />

great utility to evaluate the mechanistic contributions of HIV<br />

and HCV infection to hepatic fibrogenesis in a multicellular<br />

context. Methods: We first constructed stable monoclonal<br />

green fluorescent protein (GFP) reporter cells expressing functional<br />

antioxidant response element (ARE), NFkB, and SMAD3<br />

driven reporter transgenes, which enables these transcriptional<br />

regulators to be studied dynamically in living cells. Then, we<br />

adapted the reporter cells to create a model of co-infection<br />

that allows investigation of transcriptional dynamics associated<br />

with HIV/HCV co-infection. Finally, we used the reporter cells<br />

to create co-cultures of hepatocytes and stellate cells to further<br />

investigate the effect of HIV/HCV co-infection. Results: The<br />

efficiency of reporter cell lines was validated by using positive<br />

controls which activated 85% to 60% of total cells versus only<br />

2 to 10% mock activated cells. In In monoculture, HIV and<br />

HCV induced ARE, and SMAD3 in both cell lines (suggesting<br />

an increase in cellular oxidative stress and TGF-β production).<br />

In contrast, NFkB was activated only in Huh7.5.1 but not in<br />

LX2 cells. Co-infection with HCV and HIV led to a significant<br />

increase (2-3 fold) in the activation of these reporters in comparison<br />

to monoinfection. The co-culture model confirmed HCV<br />

and/or HIV effects, on SMAD, ARE, and NFkB activation.<br />

Interestingly, the viruses’ effects on these reporters increased<br />

additively compared to monoculture in Huh7.5.1 cells. Conclusions:<br />

HIV accentuates the HCV-related profibrogenic program<br />

in HSCs and hepatocytes through initial ROS induction, NF-kB,<br />

and TGF-B1 upregulation. Co-culture <strong>studies</strong> indicated additive<br />

effects compared with monoculture. Our novel co-culture<br />

reporter cell model represents an efficient system for studying<br />

transcriptional responses and has the potential to provide<br />

important insights into the progression of fibrosis liver disease,<br />

as well as study of potential therapeutic interventions.<br />

Disclosures:<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Shadi Salloum, Cynthia Brisac,<br />

Rohit Jindal, Shyam Sundhar Bale, Nadia Alatrakchi, Annie Kruger, Anna<br />

Lidofsky, Martin L. Yarmush


220A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

26<br />

Elevation of Galectin-9 is associated with the cytotoxicity<br />

of NK cells in chronic hepatitis<br />

Akira Nishio, Tomohide Tatsumi, Takatoshi Nawa, Takahiro<br />

Suda, Atsuo Takigawa, Tadashi Kegasawa, Yoshiki Onishi, Seiichi<br />

Tawara, Teppei Yoshioka, Satoshi Aono, Minoru Shigekawa,<br />

Hayato Hikita, Ryotaro Sakamori, Naoki Hiramatsu, Tetsuo Takehara;<br />

Department of Gastroenterology and Hepatology, Osaka<br />

University Graduate School of Medicine, Suita, Japan<br />

Background & Aims: NK cells play an essential role in the<br />

pathogenesis of chronic hepatitis C (CHC). NK cells are activated<br />

in CHC and activated NK cells exhibit elevated cytotoxicity<br />

leading to liver injury. Recently, the activated NK cells<br />

contribute to exhaustion of virus-specific T cells, which is associated<br />

with persistent viral infection. However, the mechanism<br />

of NK cell activation has not been clearly understood. The<br />

immunoregulatory protein Galectin-9 (Gal-9) is elevated in the<br />

serum and liver in CHC patients, suggesting that Gal-9 might<br />

be involved in the pathogenesis of CHC. In this study, we investigated<br />

the effect of Gal-9 on the activation of NK cells in CHC.<br />

Methods: Serum Gal-9 levels were analyzed in 39 healthy<br />

donors, 50 CHC patients and 24 IFN-treated patients achieving<br />

sustained viral response (SVR) by ELISA. The cytotoxicity of<br />

NK cells and Gal-9 levels in the supernatant were evaluated by<br />

flow cytometry and ELISA. Immunofluorescence staining was<br />

performed in CHC liver tissue. Results: Serum Gal-9 levels in<br />

CHC patients were significantly higher than those in healthy<br />

donors, and those in SVR patients were significantly lower<br />

than those in CHC patients (median, 0-90 percentile, CHC:<br />

146 pg/ml, 0-508, healthy donors: 0, 0-55.4, SVR: 53.6,<br />

0-185.6). In CHC patients, serum levels of Gal-9 in elevated<br />

ALT patients were significantly higher than those in normal ALT<br />

patients, suggesting that elevated Gal-9 might be associated<br />

with liver injury. Serum Gal-9 levels were not associated with<br />

HCV viral loads. In vitro study revealed that addition of recombinant<br />

Gal-9 resulted in significant increase of the cytotoxicity<br />

of naïve NK cells in a dose dependent manner, independent of<br />

Tim-3. Gal-9-activated NK cells showed the cytotoxicity against<br />

CD4+ or CD8+ T cells, which might contribute to persistent<br />

infection by depleting T cell. We found that a large amount of<br />

Gal-9 production was induced in the co-culture of PBMCs with<br />

JFH-1/Huh7.5.1 cells, but not with uninfected Huh7.5.1 cells.<br />

A transwell system revealed that cell-cell contact was required<br />

for Gal-9 production. In vitro depletion study revealed CD14+<br />

monocytes were required for Gal-9 in the co-culture, and both<br />

Gal-9 and CD14 positive cells were observed in immunofluorescence<br />

staining, indicating that Gal-9 was produced by<br />

Kupffer cells in the liver. The cytotoxicity of NK cells was significantly<br />

elevated when co-cultured with JFH-1 cells compared<br />

with control. Conclusion: Monocytes-derived Gal-9 induces the<br />

cytotoxicity of NK cells, which might be involved in liver injury<br />

and persistent HCV infection. These results suggest the role of<br />

Gal-9 in the pathogenesis of chronic hepatitis C.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Akira Nishio, Tomohide Tatsumi,<br />

Takatoshi Nawa, Takahiro Suda, Atsuo Takigawa, Tadashi Kegasawa, Yoshiki<br />

Onishi, Seiichi Tawara, Teppei Yoshioka, Satoshi Aono, Minoru Shigekawa,<br />

Ryotaro Sakamori, Naoki Hiramatsu<br />

27<br />

De novo formation and maturation of hepatitis C virus<br />

replication membranes visualized by pulse-chase imaging<br />

Hongliang Wang 1 , Andrew W. Tai 1,2 ; 1 Internal Medicine, University<br />

of Michigan, Ann Arbor, MI; 2 VA Ann Arbor Healthcare<br />

System, Ann Arbor, MI<br />

Hepatitis C virus (HCV) replicates on an altered host membrane<br />

structure called the “membranous web” (MW). Little is known<br />

regarding temporal events in MW formation and association<br />

with putative sites of virus assembly at or near lipid droplets<br />

(LDs). For example, it is not known whether membranous webs<br />

are stable structures that are continually renewed by newly<br />

synthesized viral proteins, or whether newly synthesized viral<br />

proteins are directed to form new MWs de novo. The HCV<br />

nonstructural protein NS5A is believed to be a component of<br />

MWs. We have developed pulse-chase fluorescent labeling<br />

and imaging approaches that allow us to discriminate ‘old’<br />

from ‘new’ NS5A-positive membranous structures (NS5A foci).<br />

With this system, we find that ‘new’ NS5A-positive foci form at<br />

sites distinct from ‘old’ NS5A foci, supporting a model of de<br />

novo MW formation instead of renewal of previously formed<br />

MWs. We also find that the phosphatidylinositol 4-kinase<br />

PI4KA and oxysterol-binding protein, which are known to be<br />

required for normal MW morphogenesis, are required to maintain<br />

the normal morphology of both ‘old’ and ‘new’ foci, and<br />

that these two proteins appear to play a role in de novo NS5A<br />

focus formation. Finally, we observe increasing colocalization<br />

of NS5A-positive foci with HCV core protein and LDs over time.<br />

Overall, these data support a model in which HCV MWs are<br />

formed de novo rather than renewed. We hypothesize that<br />

newly-formed NS5A foci undergo a process of maturation and<br />

movement towards sites of putative virion assembly at or near<br />

LDs. This may have implications for our understanding of how<br />

replication membranes are formed for other positive-sense RNA<br />

viruses.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hongliang Wang, Andrew W.<br />

Tai<br />

28<br />

Rapid normalization of antiviral and pro-inflammatory<br />

transcriptional signatures in peripheral blood of patients<br />

and livers of humanized mice treated with DAAs<br />

Matthew A. Burchill 1 , Lucy Golden-Mason 1 , Megan Wind-Rotolo 2 ,<br />

Michael Gale 3 , Hugo R. Rosen 1 ; 1 GI-Hepatology, University of<br />

Colorado-Denver, Aurora, CO; 2 Exploratory Clinical and Translational<br />

Research, Bristol-Myers Squibb, Lawrenceville, NJ; 3 Immunology,<br />

University of Washington, Seattle, WA<br />

Chronic HCV infection results in sustained immune activation<br />

in both the periphery and hepatic tissue. Recently, novel<br />

direct acting antiviral (DAA) regimens have been shown to<br />

induce rapid and sustained clearance of HCV. DAAs have<br />

provided the unique opportunity to determine whether successful<br />

treatment-induced eradication of viral RNA normalizes<br />

the dysregulated antiviral innate immune response in patients<br />

chronically infected with HCV. Methods: We used both a<br />

prospective cohort of patients and a novel humanized mouse<br />

model. Results: First, in 35 patients receiving DAAs, we found<br />

that a regimen of daclatasvir, asunaprevir, and BMS-791325<br />

induced not only rapid viral clearance, but also a re-setting<br />

of antiviral responses in the peripheral blood. Specifically,<br />

we observed that following treatment with DAAs, there was<br />

a significant reduction in IFITM1 and the chemokines CXCL10


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 221A<br />

and CXCL11 in the peripheral blood by Affymetrix microarray<br />

and confirmatory RT-PCR. Next, using a novel humanized<br />

mouse model (FRG), we assessed the transcriptional changes<br />

that occur in the hepatic tissue following DAA treatment. After<br />

establishing a stable chronic HCV infection in the mice, they<br />

were treated with daclatasvir, asunaprevir, and sofosbuvir that<br />

induced the clearance of circulating HCV within 48 hours.<br />

Despite the rapid clearance of circulating viral particles, we<br />

observed HCV NS3 protein within infected hepatocytes 14<br />

days after initiation of DAA therapy. The presence of residual<br />

NS3 protein did not affect the blunting of the expression<br />

of proinflammatory responses as we observed a significant<br />

reduction in the chemokines CXCL10 and CXCL11 in the liver<br />

of treated mice. These data suggested that the transcriptional<br />

changes that occur in the peripheral blood of patients treated<br />

with DAAs may partially reflect intrahepatic transcriptional<br />

changes. However, in livers of DAA-treated humanized mice<br />

we observed a restoration of mitochondrial anti-viral signaling<br />

protein (MAVS) and an increase in the protein IFITM1, the<br />

latter a tight junction protein typically induced only by IFN.<br />

Conclusions: These findings highlight how viral clearance elicited<br />

by DAA treatment leads to the rapid restoration of innate<br />

signaling moieties that are suppressed by HCV infection. Collectively,<br />

our data demonstrate that the rapid viral clearance<br />

following treatment with DAAs results in the rebalancing of<br />

innate antiviral response in both the peripheral blood and the<br />

liver as well as enhanced antiviral signaling within previously<br />

infected hepatocytes.<br />

Disclosures:<br />

Megan Wind-Rotolo - Employment: Bristol-Myers Squibb; Stock Shareholder:<br />

Bristol-Myers Squibb<br />

The following authors have nothing to disclose: Matthew A. Burchill, Lucy Golden-Mason,<br />

Michael Gale, Hugo R. Rosen<br />

29<br />

Activation of Hepatic Stellate Cells Drives the Rapid<br />

Development of Liver Fibrosis During Acute HCV Infection<br />

in HIV-infected Men<br />

Robert D. Gillies 1,2 , Daniel S. Fierer 1 , Marcus Yip 1,2 , Scott L. Friedman<br />

3 , Andrea D. Branch 3 , M. Isabel Fiel 4 ; 1 Infectious Diseases,<br />

Mount Sinai School of Medicine, New York, NY; 2 Monash University,<br />

Melbourne, VIC, Australia; 3 Liver Diseases, Icahn School<br />

of Medicine at Mount Sinai, New York, NY; 4 Pathology, Icahn<br />

School of Medicine at Mount Sinai, New York, NY<br />

Background The international epidemic of sexually-transmitted<br />

HCV infection among HIV-infected men is characterized by<br />

rapid development of liver fibrosis of unknown mechanism(s).<br />

The purpose of this study is to determine whether increased<br />

fibrogenesis during acute HCV infection is associated with<br />

activation of hepatic stellate cells (HSC), the principal collagen-producing<br />

cell in liver injury. Methods Liver biopsies were<br />

evaluated from HIV-infected men with acute and recent HCV<br />

infection in New York City. Date of HCV diagnosis was set<br />

as the date of the first-noted ALT elevation. The presence of<br />

alpha-smooth muscle actin (ASMA), indicative of activated<br />

HSC, was determined by counting stained foci in 10 random<br />

lobular areas per high-powered field (hpf) (40x magnification)<br />

on slides immunostained with antibodies to ASMA. Fibrosis<br />

stage (Scheuer, 0-4) was determined by examining Masson<br />

trichrome-stained slides. Collagen-proportionate area (CPA)<br />

was determined by computer digital image analysis of Sirius<br />

red-stained slides. The effect of time to biopsy on these three<br />

measurements was examined as a dichotomous variable, using<br />

1 year after HCV diagnosis. The Chi-Square test<br />

for trend was used to compare Scheuer stage of fibrosis; and<br />

the Mann-Whitney U Test was used to compare the median<br />

CPA and median activated HSC/hpf between groups. Results<br />

Liver biopsies from 20 men were evaluated. Median age at<br />

HCV diagnosis was 42 years, with a median of 9 years since<br />

HIV diagnosis; and median CD4 count was 501 cells/uL (all<br />

had >200). The median number of activated HSC/hpf was<br />

significantly higher in biopsies obtained 1 year after HCV diagnosis (p = 0.002, Table), indicating<br />

greater HSC activation early in HCV infection and lower<br />

HSC activation later in infection. Despite this lower activation<br />

of HSC in biopsies >1 year after HCV diagnosis, fibrosis<br />

stage (p = 0.01), and CPA (p = 0.06) were both higher in<br />

those with biopsy >1 year after HCV infection. Conclusion The<br />

rapid development of fibrosis that occurs in HIV-infected men<br />

during the first year after acute HCV infection is associated<br />

with ASMA-staining fibrogenic myofibroblasts. The number of<br />

activated HSC was lower after the first year of HCV infection<br />

while the stage of fibrosis was higher, indicating that once<br />

fibrogenesis has been set in motion, this process continues,<br />

although possibly at a slower pace. This rapid development of<br />

fibrosis can therefore not be considered to be benign.<br />

Disclosures:<br />

Daniel S. Fierer - Stock Shareholder: Gilead<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

The following authors have nothing to disclose: Robert D. Gillies, Marcus Yip,<br />

M. Isabel Fiel<br />

30<br />

Alcohol increases the production of hepatitis C virus<br />

(HCV) lipo-viro-particles in primary human hepatocytes<br />

Veronique Pene 1 , Céline Hernandez 1 , Etienne Blanc 2 , Lynda<br />

Aoudjehane 3,4 , Béatrice Le Grand 2 , Arnaud Carpentier 1 , Jean-<br />

François Méritet 1,5 , Filomena Conti 4,6 , Yvon Calmus 4,6 , Hélène<br />

Rouach 2 , Philippe Podevin 1,7 , Arielle R. Rosenberg 1,5 ; 1 EA 4474<br />

“Hepatitis C Virology”, Université Paris Descartes, Paris, France;<br />

2 UMRS 1124 “Toxicology Pharmacology and Cellular Signaling”,<br />

Université Paris Descartes, Inserm, Paris, France; 3 Human HepCell,<br />

Hôpital Saint-Antoine, Paris, France; 4 UMRS 938 “CDR Saint-Antoine”,<br />

Inserm, Paris, France; 5 Service de Virologie, AP-HP, Hôpital<br />

Cochin, Paris, France; 6 Unité de Transplantation Hépatique, AP-HP,<br />

Hôpital Pitié-Salpêtrière, Paris, France; 7 Centre de Référence en<br />

addictologie, AP-HP, Hôpital Pitié-Salpêtrière, Paris, France<br />

BACKGROUND & AIM: Alcoholism is a major aggravating<br />

factor of HCV infection, yet the molecular mechanisms of this<br />

comorbidity remain elusive. Clinical <strong>studies</strong> found higher viral<br />

loads in patients with chronic excessive alcohol consumption.<br />

However, the so-called “viral load” reflects a wide heterogeneity<br />

of HCV particles, the most infectious ones being those<br />

of lowest buoyant density due to their association with verylow-density<br />

lipoproteins (VLDL, the apolipoprotein B-containing<br />

triglyceride-rich lipoproteins secreted by hepatocytes) in hybrid<br />

structures termed lipo-viro-particles (LVP). With the aim of investigating<br />

the impact of alcohol on HCV infectious cycle, we took


222A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

advantage of our HCV culture system in primary human adult<br />

hepatocytes (PHH), which, contrary to the widely used hepatocarcinoma-derived<br />

Huh-7 sublines, retain the liver metabolism<br />

of ethanol and secrete authentic VLDL and LVP. METHODS:<br />

PHH were infected with the HCV strain JFH1 or a chimeric<br />

virus derived thereof and treated with increasing doses of ethanol<br />

(0-100 mM) for 2 weeks. Cultures were monitored for<br />

HCV genome replication (negative strand RT-qPCR), viral load<br />

(clinically used test), production of infectious virus (titration by<br />

focus-formation assay), and VLDL secretion (apolipoprotein B<br />

ELISA). The density of viral particles was assessed by isopycnic<br />

iodixanol ultracentrifugation. RESULTS: Ethanol exposure of<br />

HCV-infected PHH caused a time- and dose-dependent increase<br />

in the viral load, comparable to that reported in clinical <strong>studies</strong>.<br />

Most strikingly, it caused an even greater increase in the<br />

infectious titer but did not significantly affect the viral genome<br />

replication, thus pointing to an effect on virus morphogenesis.<br />

This effect was not seen in Huh-7.5.1 cells treated in parallel,<br />

suggesting that it involves the liver metabolism of ethanol. The<br />

higher specific infectivity of HCV particles produced by PHH<br />

in the presence of ethanol correlated with lower mean buoyant<br />

density, consistent with triglyceride enrichment. Finally, in<br />

either HCV-infected or naïve PHH, addition of ethanol caused<br />

a time-dependent increase in VLDL secretion. CONCLUSION:<br />

This study in the relevant PHH model reveals that ethanol via its<br />

metabolites increases the production of HCV particles of lowest<br />

buoyant density and highest infectivity, i.e., LVP, most likely<br />

due to the impact of ethanol on triglyceride metabolism that<br />

results in increased VLDL secretion. Drugs targeting this host<br />

metabolic pathway may be useful in difficult-to-treat alcoholic<br />

patients, often poorly compliant with therapy and therefore at<br />

risk for resistance if treated by direct antivirals only. VP & CH:<br />

equal contribution.<br />

Disclosures:<br />

The following authors have nothing to disclose: Veronique Pene, Céline Hernandez,<br />

Etienne Blanc, Lynda Aoudjehane, Béatrice Le Grand, Arnaud Carpentier,<br />

Jean-François Méritet, Filomena Conti, Yvon Calmus, Hélène Rouach, Philippe<br />

Podevin, Arielle R. Rosenberg<br />

followed by 15 weeks of combined therapy with 250mg REP<br />

2139-Ca and 180ug Pegasys® and then 33 weeks with<br />

Pegasys® monotherapy. Viremia (HDV RNA and HBV DNA),<br />

HBsAg and anti-HBs are followed every two weeks (Robogene<br />

RT-PCR, Abbott RealTime HBV, Abbott Architect) performed at<br />

the Institute of Virology, University of Duisburg-Essen (Essen,<br />

Germany). HDV RNA is validated on separate test platforms<br />

at the National Genetics Institute (Los Angeles, USA) and the<br />

Institute of Virology, Technische Universität München (Munich,<br />

Germany). Results: REP 2139-Ca treatment is well tolerated<br />

with mild and quickly resolving IV infusion reactions. Serum<br />

HBsAg is currently reduced 1-6 log in 11/12 patients (5 with<br />

serum HBsAg < 1 IU / ml) and HDV RNA is currently reduced<br />

1.5-7 log in 12 /12 patients (undetectable in 6 patients). Anti-<br />

HBs is detected in 10/12 patients, with 6 patients < 10mIU /<br />

ml and after combined exposure to Pegasys®, 5 patients have<br />

substantially increased anti-HBs titers from 50 to > 800 mIU /<br />

ml. In all patients with pre-treatment HBV DNA < 10 IU / ml,<br />

de-repression of serum HBV DNA is observed. Conclusions:<br />

Updated interim data from the REP 301 protocol assessing the<br />

safety and antiviral efficacy of REP 2139 (first in monotherapy<br />

and then with add on Pegasys® at week 16) in 12 Caucasian<br />

patients with chronic HBV / HDV co-infection demonstrated<br />

substantial reductions in serum HBsAg and HDV RNA as well<br />

as appearance of anti-HBs. HDV RNA reductions appear stronger<br />

than HBsAg reductions, suggesting an additional antiviral<br />

mechanism other than the inhibition of subviral particle assembly<br />

may affect HDV directly. REP 2139-Ca may become an<br />

important new therapeutic option for patients with chronic HBV<br />

/ HDV infection.<br />

Disclosures:<br />

Michel Bazinet - Board Membership: REPLICor Inc.; Employment: REPLICor Inc.;<br />

Management Position: REPLICor Inc.; Patent Held/Filed: REPLICor Inc.; Stock<br />

Shareholder: REPLICor Inc.<br />

Andrew Vaillant - Employment: REPLICor; Stock Shareholder: REPLICor<br />

The following authors have nothing to disclose: Victor Pantea, Valentin<br />

Cebotarescu, Lilia Cojuhari, Paulina Jimbei, Jeffrey Albrecht, Peter Schmid, Hadi<br />

Karimzadeh, Michael Roggendorf<br />

31<br />

Update on the safety and efficacy of REP 2139 monotherapy<br />

and subsequent combination therapy with<br />

pegylated interferon alpha-2a in chronic HBV / HDV<br />

co-infection in Caucasian patients<br />

Michel Bazinet 1 , Victor Pantea 2 , Valentin Cebotarescu 2 , Lilia<br />

Cojuhari 3 , Paulina Jimbei 3 , Jeffrey Albrecht 4 , Peter Schmid 4 , Hadi<br />

Karimzadeh 5 , Michael Roggendorf 5 , Andrew Vaillant 1 ; 1 Replicor<br />

Inc., Montreal, QC, Canada; 2 N. Testemitanu State University of<br />

Medicine and Pharmacy, Chisinau, Moldova (the Republic of);<br />

3 Toma Ciorba Infectious Clinical Hospital, Chisinau, Moldova<br />

(the Republic of); 4 National Genetics Institute, Los Angeles, CA;<br />

5 Institure of Virololgy, Technische Universität München, Munich,<br />

Germany<br />

Introduction: HBV / HDV co-infection causes rapid progression<br />

of liver disease and with no approved therapy, presents a<br />

significant unmet medical need. Nucleic acid polymers (NAPs)<br />

block HBV subviral particle assembly and release from infected<br />

hepatocytes and can eliminate serum HBsAg. As the NAP REP<br />

2139 was previously been shown to clear serum HBsAg and<br />

improve the ability of immunotherapy to elicit SVR in Asian<br />

patients with HBV, its activity in combination with Pegasys® in<br />

HBV / HDV co-infected Caucasian patients is currently being<br />

examined. Methods: In a phase II proof of concept trial (REP<br />

301; NCT02233075), patients with chronic HBV / HDV co-infection<br />

received once weekly dosing of 500mg REP 2139-Ca<br />

(calcium chelate complex) by 2h IV infusion for 15 weeks,<br />

32<br />

Reductions in cccDNA under NUC and ARC-520 therapy<br />

in chimpanzees with chronic hepatitis B virus infection<br />

implicate integrated DNA in maintaining circulating<br />

HBsAg<br />

Christine I. Wooddell 1 , Deborah Chavez 2 , Jason E. Goetzmann<br />

3 , Bernadette Guerra 2 , Ryan M. Peterson 1 , Helen Lee 2 ,<br />

Julia O. Hegge 1 , Robert Gish 4 , Stephen Locarnini 5 , Christopher<br />

R. Anzalone 1 , Robert E. Lanford 2 , David L. Lewis 1 ; 1 Arrowhead<br />

Madison, Arrowhead Research Corporation, Madison, WI; 2 Texas<br />

Biomedical Research Institute, San Antonio, TX; 3 New Iberia<br />

Research Center, University of Louisiana at Lafayette, New Iberia,<br />

LA; 4 Department of Medicine, Stanford University Medical Center,<br />

San Diego, CA; 5 Victorian Infectious Diseases Reference Laboratory,<br />

Melbourne, VIC, Australia<br />

Background: RNAi therapeutic ARC-520 designed to target<br />

all cccDNA-derived transcripts reduces viral antigenemia for<br />

>1 month after single doses in HBV patients. Here we report<br />

the effect of multiple ARC-520 doses on hepatic HBV DNA<br />

and RNA in HBV chimps. Methods: 9 chimps (5 M, 4 F; 9-37<br />

yrs) received 6-11 monthly injections of ARC-520 concurrent<br />

with NUC therapy. 5 were HBeAg-positive (HBeAg+), baseline<br />

DNA 8-9 log 10<br />

IU/mL serum, and 4 were HBeAg-negative<br />

(HBeAg-), ≤3 log 10<br />

IU/mL. Chimps received NUCs for<br />

8-24 weeks prior to ARC-520 dosing. Liver biopsies from 8<br />

chimps were taken at baseline, completion of NUC lead-in<br />

and on study. HBV DNA, +/- plasmid-safe DNase digestion to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 223A<br />

enrich cccDNA, was measured by qPCR. Pre-core/core RNA<br />

(C probe) and total HBV RNA (Total probe) were measured<br />

by RT-qPCR. The Guide for the Care and Use of Laboratory<br />

Animals was strictly adhered to. Results: During NUC lead-in,<br />

total liver HBV DNA decreased 1.1-2.5 log 10<br />

in HBeAg+ but<br />

not appreciably in HBeAg- chimps. cccDNA in HBeAg+ chimps<br />

decreased 0.7 ± 0.6 log 10<br />

. Following addition of ARC-520 in<br />

HBeAg+, total liver DNA decreased from baseline by 1.5–2.9<br />

log 10<br />

and cccDNA by 1.4 ± 0.7 log 10<br />

, the degree of reduction<br />

generally correlating with duration of treatment. Neither<br />

total HBV DNA nor cccDNA levels changed remarkably in<br />

HBeAg- during the study, which at baseline had 2-4 orders of<br />

magnitude less cccDNA than HBeAg+ chimps. HBeAg- chimps<br />

had 50-fold more DNase-sensitive HBV DNA, possibly indicating<br />

the majority is integrated DNA rather than cccDNA.<br />

HBV RNA was not reduced by NUCs, but with addition of<br />

ARC-520 RNA reductions tracked qHBsAg reductions. In<br />

HBeAg+, Total probe detected 1-2x as many transcripts as<br />

the C probe, suggesting similar levels of core/pre-core and<br />

S transcripts. In HBeAg-, the Total probe detected 37x more<br />

transcripts than the C probe, supporting a greater proportion of<br />

HBsAg transcripts being produced from integrated HBV DNA<br />

in HBeAg- chimps. Integration between DR1 and DR2 would<br />

result in HBV RNA lacking ARC-520 target sites, consistent<br />

with greater HBsAg reduction in HBeAg+ (1.7 ± 0.5 log 10<br />

)<br />

than HBeAg- chimps (0.7 ± 0.1 log 10<br />

). Administration of siRNA<br />

targeted to integrant-produced transcripts resulted in HBsAg<br />

reductions up to 2.3 log 10<br />

beyond those produced by ARC-520<br />

in HBeAg-. Conclusions: 1) ARC-520 reduced total liver DNA<br />

and cccDNA beyond levels achieved in HBeAg+ with NUCs<br />

during lead-in; 2) ARC-520 but not NUCs reduced HBV RNA<br />

and antigens; 3) integrated HBV DNA may be important in<br />

maintaining HBsAg in chronic HBV, especially in HBeAg-. This<br />

finding, if confirmed, has important implications for development<br />

of new HBV therapies.<br />

Disclosures:<br />

Christine I. Wooddell - Employment: Arrowhead Research Corporation<br />

Ryan M. Peterson - Employment: Arrowhead Research Corporation<br />

Julia O. Hegge - Employment: Arrowhead Research Corp<br />

Robert Gish - Advisory Committees or Review Panels: Gilead, AbbVie, Arrowhead;<br />

Consulting: Eiger, Isis, Genentech; Speaking and Teaching: Gilead, Abb-<br />

Vie; Stock Shareholder: Arrowhead<br />

Stephen Locarnini - Consulting: Gilead, Arrowhead; Employment: Melbourne<br />

Health<br />

Robert E. Lanford - Grant/Research Support: Arrowhead Research<br />

David L. Lewis - Employment: Arrowhead Research Corporation<br />

The following authors have nothing to disclose: Deborah Chavez, Jason E. Goetzmann,<br />

Bernadette Guerra, Helen Lee, Christopher R. Anzalone<br />

33<br />

Inhibition of Hepatitis B Virus Replication by the HBV<br />

Core Inhibitor NVR 3-778<br />

Angela Lam, Suping Ren, Robert Vogel, Christine Espiritu, Mollie<br />

Kelly, Vincent Lau, George D. Hartman, Lalo Flores, Klaus Klumpp;<br />

Novira Therapeutics Inc., Doyleston, PA<br />

Background: Hepatitis B virus (HBV) chronically infects more<br />

than 350 million people worldwide, and is a major cause of<br />

severe liver disease including cirrhosis and hepatocellular carcinoma.<br />

NVR 3-778 represents a new class of HBV core inhibitors<br />

and is currently in clinical development for the treatment of<br />

chronic hepatitis B. We present the preclinical characterization<br />

of NVR 3-778 and its antiviral properties in cells expressing<br />

HBV. Methods: HBV core assembly was examined using fluorescence<br />

quenching and electron microscopy. HBV replication<br />

inhibition was measured as reduction of secreted HBV DNA in<br />

HBV producing HepG2.2.15 cells. The production, encapsidation<br />

and secretion of HBV RNA was determined using Northern<br />

blot and Quantigene assays. Antiviral activity of NVR 3-778<br />

against nucleoside resistant HBV variants was determined by<br />

cell based phenotyping. HBV broad spectrum antiviral activity<br />

was examined by testing NVR 3-778 across all HBV genotypes<br />

(A-H). Results: NVR 3-778 inhibited HBV replication in<br />

HepG2.2.15 cells and could fully block the production of HBV<br />

DNA.. The compound could effectively induce mis-assembly of<br />

recombinant HBV core protein and electron microscopy indicated<br />

the formation of capsid-like particles. In HBV expressing<br />

cells, NVR 3-778 blocked the encapsidation of HBV pgRNA,<br />

HBV DNA formation, and the release of HBV DNA and RNA<br />

containing particles. NVR 3-778 was similarly active against<br />

wild-type and HBV nucleos(t)ide resistant variants with defined<br />

amino acid changes within the reverse transcriptase protein:<br />

rtL180M/M204V, rtN236T, rtA181V, rtA181V/N236T, and<br />

rtL180M/M204V/N236T. Furthermore, in transfection assays<br />

NVR 3-778 inhibited HBV replication across representative<br />

HBV strains from genotypes A to H (EC 50<br />

values from 0.20 to<br />

0.58 mM). Conclusions: Preclinical antiviral profiling <strong>studies</strong><br />

showed that the antiviral HBV replication inhibitor NVR 3-778<br />

could induce mis-assembly of core protein into capsid-like particles<br />

in vitro. Treatment of HBV producing cells with NVR 3-778<br />

was associated with inhibition of pgRNA encapsidation. Consequently,<br />

the downstream processes of HBV DNA production<br />

and secretion of infectious HBV DNA-containing virions (Dane<br />

particles) were inhibited. In addition, NVR 3-778 also blocked<br />

the secretion of HBV RNA-containing particles. NVR 3-778 was<br />

active against HBV variants resistant to nucleos(t)ide analogs<br />

and was broadly active across all HBV genotypes. These results<br />

support the clinical development of NVR 3-778 as a potential<br />

new therapy for chronic hepatitis B patients alone or in combination<br />

with nucleoside analogs or other antiviral agents.<br />

Disclosures:<br />

Angela Lam - Employment: Novira Therapeutics, Merck & Co<br />

Christine Espiritu - Employment: Novira Therapeutics<br />

Mollie Kelly - Independent Contractor: Novira Therapeutics<br />

George D. Hartman - Management Position: Novira Therapeutics<br />

Klaus Klumpp - Board Membership: Riboscience LLC; Employment: Novira Therapeutics<br />

Inc<br />

The following authors have nothing to disclose: Suping Ren, Robert Vogel, Vincent<br />

Lau, Lalo Flores<br />

34<br />

Dual-gRNAs and gRNA-microRNA (miRNA)-gRNA ternary<br />

cassette combined CRISPR/Cas9 system and RNAi<br />

approach promotes the clearance of HBV cccDNA<br />

Jie Wang, Fengmin Lu; Peking University Health Science Center,<br />

Beijing, China<br />

Purpose: To investigate the potential use of CRISPR/Cas9<br />

approach for eradicating hepatitis B virus (HBV) infection, we<br />

designed and screened for the most effective gRNAs against<br />

HBV of genotypes A-D, and developed a novel gRNA-microRNA<br />

(miRNA)-gRNA ternary cassette. Methods: A total of<br />

15 gRNAs against HBV of genotypes A-D were designed. 11<br />

combinations of two above gRNAs (dual-gRNAs) covering the<br />

different regions of HBV genome were chosen. Based on the<br />

efficiency of dual-gRNAs in suppressing HBV replication, a<br />

novel gRNA-miRNA-gRNA ternary cassette driven by a single<br />

U6 promoter was designed by integrating an anti-HBV primiR31<br />

mimic between two HBV-specific gRNAs, in which two<br />

gRNAs could be released through Drosha/DGCR8 processing.<br />

The HBV replication was examined by the measurement<br />

of HBV surface antigen (HBsAg) and e antigen (HBeAg) in<br />

the culture supernatant. The destruction of HBV-expressing vec-


224A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tor was examined by polymerase chain reaction (PCR) and<br />

sequencing method, and the destruction of cccDNA was examined<br />

by KCl precipitation, plasmid-safe ATP-dependent DNase<br />

digestion, rolling circle amplification and quantitative PCR combined<br />

method. Results: All of gRNAs could significantly reduce<br />

the HBsAg or HBeAg production in the culture supernatant,<br />

which was dependent on the region in which gRNA against.<br />

All of dual-gRNAs could efficiently suppress the HBsAg and<br />

HBeAg production for HBV of genotypes A-D, and the efficacy<br />

of dual-gRNAs in suppressing HBsAg and HBeAg production<br />

was significantly increased when compared to the single gRNA<br />

used alone. Furthermore, with PCR direct sequencing we confirmed<br />

that these dual-gRNAs could specifically destroy HBV<br />

expressing template by removing the fragment between the<br />

cleavage sites of the two used gRNAs. The gRNA-miR-HBVgRNA<br />

ternary cassette exerted a synergistic effect in efficiently<br />

inhibiting HBV replication and destroying HBV-expressing template<br />

in vitro and in vivo. We determined that the optimal stem<br />

extended length of pri-miRNA was 38 base pairs in our ternary<br />

cassette. Most importantly, together with RNAi approach<br />

the gRNA-miR-HBV-gRNA ternary cassette showed the potent<br />

activity on the destruction of HBV cccDNA. Conclusions: These<br />

results suggested that dual-gRNAs guided CRISPR/Cas9 system<br />

could efficiently destroy HBV expressing templates (genotypes<br />

A-D). Together with RNAi approach, the gRNA-miR-HBV-gRNA<br />

ternary cassette showed the potent activity on the destruction<br />

of HBV cccDNA. Since HBV cccDNA was an obstacle for the<br />

elimination of chronic HBV infection, thus the CRISPR/Cas9 system<br />

may be a potential tool for the clearance of HBV cccDNA.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jie Wang, Fengmin Lu<br />

both GSEA and IPA. Strikingly, there was a comparable transcriptional<br />

response to TLR7-CM treatment in HBV-infected and<br />

uninfected PHH (p0.80), including a comparable<br />

induction of IFN-stimulated gene (ISG) signature. Notably,<br />

APOBEC3B (A3B) mRNA levels were only weakly induced<br />

by TLR7-CM (mean = 2.6-fold), and A3A expression was not<br />

significantly stimulated. Consistent with the lack of viral interference<br />

in the hepatocyte response to TLR7-CM, HBV infection<br />

did not result in prominent global transcriptional response in<br />

Mock-CM treated PHH. Knock-down of IFNAR1 mRNA (~90%<br />

reduction) abrogated the antiviral activity of TLR7-CM, suggesting<br />

that one or more of the ISGs strongly induced by treatment<br />

(e.g. MX1, IFI16, IFIT1 and IFITM1) may be HBV restriction<br />

factors. Conclusion: These data demonstrate that type I IFN is a<br />

key antiviral cytokine induced by GS-9620 in PBMCs, and that<br />

HBV does not interfere with IFN-signaling in human hepatocytes.<br />

In addition, transcriptional profiling identified candidate<br />

HBV restriction factors, but indicated that APOBEC3A likely<br />

does not mediate the antiviral activity of TLR7-CM in HBV-infected<br />

hepatocytes.<br />

Disclosures:<br />

Li Li - Employment: Gilead Sciences<br />

Congrong Niu - Employment: Gilead Science<br />

Stephane Daffis - Employment: Gilead Sciences<br />

Hilario Ramos - Employment: Gilead Sciences<br />

Eduardo Salas - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Christian Voitenleitner - Employment: Gilead Sciences<br />

William E. Delaney - Employment: Gilead Sciences; Patent Held/Filed: Gilead<br />

Sciences; Stock Shareholder: Gilead Sciences<br />

Simon P. Fletcher - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

35<br />

HBV Does Not Modulate the Transcriptional Response<br />

to TLR7-Induced Cytokines in Highly Infected Primary<br />

Human Hepatocytes<br />

Li Li, Congrong Niu, Stephane Daffis, Hilario Ramos, Eduardo<br />

Salas, Christian Voitenleitner, William E. Delaney, Simon P.<br />

Fletcher; Gilead Sciences, Foster City, CA<br />

Background & Aims: GS-9620, an oral agonist of toll-like<br />

receptor 7 (TLR7), is in phase 2 trials for the treatment of<br />

chronic hepatitis B (CHB). We previously demonstrated that<br />

prolonged exposure to antiviral cytokines directly induced by<br />

TLR7 activation, such as interferon-α (IFN-α), potently inhibited<br />

HBV in primary human hepatocytes (PHH) in vitro. Here we<br />

characterized the transcriptional response of PHH to TLR7-induced<br />

cytokines in order to identify potential HBV restriction<br />

factors and to determine whether HBV infection modulates<br />

signaling of these cytokines in hepatocytes. Methods: PHH<br />

from two different donors were infected with HBV or mock<br />

infected, and 13 days later were treated with media from<br />

healthy human PBMCs stimulated with either GS-9620 (TLR7<br />

conditioned media; TLR7-CM) or DMSO (control; Mock-CM).<br />

High level infection (≥75% PHH infected) was confirmed by<br />

HBV core staining. Whole transcriptome analysis of PHH<br />

± HBV at 8, 24 and 48 hours post-treatment with TLR7-CM<br />

and Mock-CM was performed by RNA-Seq. Transcriptional<br />

responses were characterized by gene set enrichment analysis<br />

(GSEA) and Ingenuity Pathway Analysis (IPA). Knock-down of<br />

IFN-α/β receptor 1 (IFNAR1) mRNA was performed by siRNA,<br />

and included a negative (scramble) control. Results: TLR7-CM<br />

induced a comparable transcriptional signature in PHH from<br />

two independent donors. In both donors, TLR7-CM induced the<br />

greatest transcriptional response at 8 hours post-treatment, with<br />

IFN signaling pathways being the top induced pathways by<br />

36<br />

GalNAc-siRNA conjugate ALN-HBV targets a highly<br />

conserved, pan-genotypic X-orf viral site and mediates<br />

profound and durable HBsAg silencing in vitro and in<br />

vivo<br />

Laura Sepp-Lorenzino, Andrew G. Sprague, Tara Mayo, Tuyen<br />

M. Nguyen, Svetlana Shulga Morskaya, Huilei Xu, Stuart Milstein,<br />

Greg Hinkle, Pia Kasperkovitz, Richard G. Duncan, Natalie Keirstead,<br />

Brenda Carito, Lauren Moran, Prasoon Chaturvedi, Krishna<br />

C. Aluri, Husain Attarwala, Renta M. Hutabarat, Ju Liu, Chris Tran,<br />

Qianfan Wang, Benjamin S. Brigham, Akin Akinc, Klaus B. Charisse,<br />

Vasant Jadhav, Satya Kuchimanchi, Martin A. Maier, Muthiah<br />

Manoharan, Rachel Meyers, Tadeusz Wyrzykiewicz, Haroon<br />

Hashmi, Julie Donovan, Tim Mooney, Daniel Freedman, Tanya<br />

P. Sengupta, Karin Galil, Eoin Coakley, Patrick Haslett; Alnylam<br />

Pharmaceuticals, Cambridge, MA<br />

Background: Despite the prevalence of chronic HBV infection<br />

(CHB), there are no highly effective therapies allowing sustained<br />

off-treatment viral control and/or cure of HBV infection.<br />

An RNAi therapeutic targeting the HBV genome has the potential<br />

to achieve a “functional cure” by effectively decreasing<br />

expression of all viral gene products, including tolerogenic viral<br />

antigens, e.g., HBsAg. Methods and Results: ALN-HBV is a<br />

hepatocyte-targeted GalNAc-siRNA conjugate with Enhanced<br />

Stabilization Chemistry (ESC) for subcutaneous (SC) administration.<br />

It targets a site in the HBV X protein open reading<br />

frame (ORF), thereby targeting all four HBV RNA transcripts<br />

for degradation by RNA interference. Bioinformatic analyses<br />

indicate a >95% perfect sequence homology with a database<br />

of 3,950 full length HBV viral genomes incorporating genotypes<br />

A through J, and a low potential for off-target activity<br />

against host genes. Pan-genotypic activity was confirmed in


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 225A<br />

vitro, using reporter constructs representative of genotypes A-J.<br />

In HepG2.2.15 cells, ALN-HBV mediates viral transcript silencing<br />

with ED 50<br />

s of 12-19 pM, as determined for amplicons in<br />

the P and S ORFs. In vivo, in an AAV-HBV mouse model, a<br />

single 3 mg/kg SC dose of ANL-HBV resulted in mean 1.6<br />

log 10<br />

reduction in plasma HBsAg (3.6 log 10<br />

max reduction) at<br />

10-15 days post dosing. After 3 weekly doses of 3 mg/kg SC,<br />

mean HBsAg silencing >2.9 log 10<br />

was observed, with several<br />

animals being below the level of quantitation (100 days after the last dose.<br />

In vitro metabolic stability, and mouse and rat pharmacokinetics<br />

and liver exposure were consistent with ESC chemistry<br />

which enables the long-lived pharmacologic response. Exploratory<br />

toxicology and histopathology were performed in rats<br />

sacrificed 24 hr after 3 weekly doses of 30 and 100 mg/kg<br />

SC, and revealed excellent tolerability. Additional pharmacology<br />

<strong>studies</strong> are in progress, including combination <strong>studies</strong><br />

with ALN-PDL, an RNAi drug targeting PD-L1 in hepatocytes.<br />

ALN-PDL is expected to result in augmentation of HBV-specific<br />

cellular immunity in the liver, thereby improving immunological<br />

control of HBV infection locally, while avoiding the immunotoxicity<br />

of systemic immune checkpoint blockade. Conclusion:<br />

ALN-HBV is a single, hepatocyte-targeted siRNA conjugate<br />

targeting a highly conserved region in the HBV genome. It<br />

mediates specific, potent and durable silencing of HBV viral<br />

transcripts and HBsAg expression. ALN-HBV is being developed<br />

as finite treatment in combination with standard-of-care<br />

nucleos(t)ide analogs as a means for inducing a functional cure<br />

in CHB patients. A CTA filing is planned for late 2015.<br />

Disclosures:<br />

Laura Sepp-Lorenzino - Employment: Alnylam<br />

Andrew G. Sprague - Employment: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Tuyen M. Nguyen - Employment: Alnylam Pharmaceutical; Stock Shareholder:<br />

Alnylam Pharmaceutical<br />

Svetlana Shulga Morskaya - Employment: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Stuart Milstein - Employment: Alnylam<br />

Greg Hinkle - Employment: Alnylam Pharmaceuticals<br />

Richard G. Duncan - Employment: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Krishna C. Aluri - Employment: Alnylam Pharmaceuticals<br />

Husain Attarwala - Employment: Alnylam Pharmaceuticals<br />

Renta M. Hutabarat - Employment: Alnylam Pharmaceuticals<br />

Ju Liu - Employment: Alnylam<br />

Benjamin S. Brigham - Employment: Alnylam Pharmaceuticals<br />

Akin Akinc - Employment: Alnylam Pharmaceuticals<br />

Klaus B. Charisse - Employment: Alnylam Pharmaceuticals<br />

Vasant Jadhav - Employment: Alnylam Pharmaceuticals<br />

Satya Kuchimanchi - Employment: Alnylam Pharmaceuticals<br />

Martin A. Maier - Employment: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Muthiah Manoharan - Employment: Alnylam<br />

Rachel Meyers - Employment: Alnylam Pharmaceuticals<br />

Haroon Hashmi - Employment: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Daniel Freedman - Employment: Alnylam Pharmaceuticals<br />

Tanya P. Sengupta - Employment: Alnylam Pharmaceuticals<br />

Karin Galil - Consulting: Alnylam Pharmaceuticals<br />

Eoin Coakley - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Patrick Haslett - Employment: Alnylam Pharmaceuticals<br />

The following authors have nothing to disclose: Tara Mayo, Huilei Xu, Pia<br />

Kasperkovitz, Natalie Keirstead, Brenda Carito, Lauren Moran, Prasoon<br />

Chaturvedi, Chris Tran, Qianfan Wang, Tadeusz Wyrzykiewicz, Julie Donovan,<br />

Tim Mooney<br />

37<br />

Safety and efficacy of daclatasvir plus sofosbuvir with<br />

or without ribavirin for the treatment of chronic HCV<br />

genotype 3 infection: Interim results of a multicenter<br />

European compassionate use program<br />

Tania M. Welzel 1 , Joerg Petersen 2 , Peter Ferenci 3 , Michael<br />

Gschwantler 4 , Kerstin Herzer 5 , Markus Cornberg 6 , Eckart Schott 7 ,<br />

Thomas Berg 8 , Ulrich Spengler 9 , Ola Weiland 10 , Marc van der<br />

Valk 11 , Andreas Geier 12 , Jürgen K. Rockstroh 9 , Markus Peck-Radosavljevic<br />

3 , Yue Zhao 13 , Maria Jesus Jimenez Exposito 14 , Stefan<br />

Zeuzem 1 ; 1 Universitätsklinikum der Johann Wolfgang Goethe<br />

Universität, Frankfurt, Germany; 2 IFI Institut für Interdisziplinäre<br />

Medizin, Hamburg, Germany; 3 Medizinische Universität Wien,<br />

Vienna, Austria; 4 Wilhelminenspital, Vienna, Austria; 5 Universitätsklinikum<br />

Essen (AöR), Essen, Germany; 6 Medizinische<br />

Hochschule Hannover, Hannover, Germany; 7 Charité Universitätmedizin<br />

Berlin, Berlin, Germany; 8 Universitätsklinikum Leipzig,<br />

Leipzig, Germany; 9 Universitätsklinikum Bonn, Bonn, Germany;<br />

10 Karolinska Institutet, Solna, Sweden; 11 Academic Medical Center,<br />

Amsterdam, Netherlands; 12 Universitätsklinikum Würzburg,<br />

Würzburg, Germany; 13 Bristol-Myers Squibb, Hopewell, NJ;<br />

14 Bristol-Myers Squibb, Princeton, NJ<br />

BACKGROUND: Despite progress in the development of well<br />

tolerated, all-oral, interferon-free therapies, the treatment of<br />

HCV genotype (GT) 3 infection remains challenging. Current<br />

treatment options are limited, and debate continues regarding<br />

the optimal regimen and duration of treatment, especially for<br />

treatment-experienced patients with liver cirrhosis. Here we<br />

report interim results on the combination daclatasvir (DCV) plus<br />

sofosbuvir (SOF) with or without ribavirin (RBV) for the treatment<br />

of GT3-infected patients from a European compassionate use<br />

program (CUP). METHODS: This CUP enrolled adult patients<br />

with chronic HCV infection who were at a high risk of hepatic<br />

decompensation or death within 12 months if left untreated,<br />

and who had no available treatment options. Patients received<br />

DCV 60mg + SOF 400mg QD, with RBV added at the physician’s<br />

discretion. Recommended duration of treatment was<br />

24 weeks. This interim analysis includes 101 GT 3-infected<br />

patients enrolled in the CUP with available data on March 16,<br />

2015, of whom 24 patients have available SVR12 data to<br />

date. RESULTS: Most patients were male (68%), white (90%);<br />

median age: 54 years (31–75). Fifty-nine (58%) were treatment<br />

experienced. Most patients had cirrhosis (81%); 46 of them<br />

(56%) were Child-Pugh class B/C; and 13 (16%) had MELD<br />

scores ≥15. Seven patients were liver transplant recipients; 15<br />

were HIV/HCV coinfected. Median baseline HCV RNA was<br />

5.6 log 10<br />

IU/mL. Overall, SVR12 rates among GT3-infected<br />

patients with HCV RNA available data were 92% (22/24).<br />

One patient meets criteria for relapse. SVR12 rates among<br />

sub-groups are presented in the Table. Updated SVR12 data<br />

will be presented. Twenty-two patients experienced at least one<br />

serious adverse event (AE); 5 patients discontinued therapy due<br />

to AEs (pneumonia/multi-organ failure, physical deterioration,<br />

hepatic encephalopathy, syncope, undetermined); 2 patients<br />

died (multi-organ failure, liver related). CONCLUSION: In this<br />

preliminary analysis, DCV in combination with SOF±RBV led<br />

to high SVR12 rates in HCV GT3-infected patients, regardless<br />

of cirrhosis status, and was well tolerated. The use of RBV did<br />

not affect efficacy outcomes.


226A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

†All liver transplant patients had cirrhosis, ‡Includes 3 liver transplant<br />

patients<br />

Disclosures:<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

Peter Ferenci - Advisory Committees or Review Panels: Idenix, Gilead, MSD,<br />

Janssen, Salix, AbbVie, BMS; Patent Held/Filed: Madaus Rottapharm; Speaking<br />

and Teaching: Gilead, Roche<br />

Michael Gschwantler - Advisory Committees or Review Panels: Janssen, BMS,<br />

Gilead, AbbVie; Speaking and Teaching: Janssen, BMS, Gilead, AbbVie<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

Eckart Schott - Advisory Committees or Review Panels: Gilead, Roche, Bayer,<br />

BMS, Abbvie; Speaking and Teaching: Gilead, Novartis, Roche, MSD, Bayer,<br />

Falk, BMS, Janssen, Abbvie<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Ola Weiland - Advisory Committees or Review Panels: Merk, BMS, Medivir, Gilead,<br />

AbbVie; Grant/Research Support; Speaking and Teaching: Merk, Roche,<br />

BMS, Novartis, Janssen, Medivir, Gilead, AbbVie<br />

Marc van der Valk - Advisory Committees or Review Panels: gilead, msd, bms,<br />

abbvie, janssen cilag, viiV, roche<br />

Andreas Geier - Advisory Committees or Review Panels: AbbVie, Janssen; Speaking<br />

and Teaching: BMS, Gilead, Sequana, Falk, Novartis<br />

Jürgen K. Rockstroh - Advisory Committees or Review Panels: Abbvie, BI, BMS,<br />

Merck, Roche, Tibotec, Abbvie, Bionor, Tobira, ViiV, Gilead, Janssen; Consulting:<br />

Novartis; Grant/Research Support: Merck; Speaking and Teaching: Abbott,<br />

BI, BMS, Merck, Roche, Tibotec, Gilead, Janssen, ViiV<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Maria Jesus Jimenez Exposito - Employment: Bristol-Myers Squibb<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

The following authors have nothing to disclose: Kerstin Herzer, Ulrich Spengler,<br />

Yue Zhao<br />

38<br />

Sofosbuvir/GS-5816+GS-9857 for 6 or 8 Weeks in<br />

Genotype 1 or 3 HCV-infected Patients<br />

Edward J. Gane 1 , Robert H. Hyland 2 , Yin Yang 2 , Luisa M. Stamm 2 ,<br />

Diana M. Brainard 2 , John G. McHutchison 2 , Catherine A. Stedman<br />

3 ; 1 Auckland Clinical Studies, Auckland, New Zealand; 2 Gilead<br />

Sciences, Inc, Foster City, CA; 3 Christchurch Clinical Studies<br />

Trust, Christchurch, New Zealand<br />

Background: Sofosbuvir (SOF), GS-5816, and GS-9857 target<br />

3 distinct viral proteins, NS5B, NS5A, and NS3, respectively.<br />

The combination of these pangenotypic agents has the potential<br />

to improve efficacy and may reduce treatment duration<br />

across different patient populations. We previously reported<br />

safety and efficacy of SOF/GS-5816 + GS-9857 administered<br />

for 4 or 6 weeks in patients with HCV genotype 1 (GT1) infection,<br />

including treatment-naïve (TN) GT1 patients without cirrhosis;<br />

TN GT1 patients with cirrhosis; and GT1 patients who had<br />

failed a regimen containing at least 2 direct antiviral agents.<br />

We now evaluate the safety and efficacy of this regimen<br />

administered for 6 or 8 weeks in more difficult-to-treat populations,<br />

including treatment experienced (TE) GT 1 patients with<br />

cirrhosis who had failed PEG/RBV; TE GT 1 patients who had<br />

failed protease inhibitor (PI) + PEG/RBV; TN GT 3 patients with<br />

cirrhosis; and TE GT 3 patients with cirrhosis who had failed<br />

PEG/RBV. Methods: All patients received once daily SOF/<br />

GS-5816 400mg/100mg in a fixed dose combination, plus<br />

GS-9857 100mg. Treatment duration of each group differed<br />

according to the patient’s baseline characteristics: 8 weeks<br />

for TE GT 1 patients who had failed prior PEG/RBV, 8 weeks<br />

for TE GT 1 patients who had failed PI + PEG/RBV, 6 weeks<br />

for TN GT 3 patients and 8 weeks for TE GT 3 patients. The<br />

primary efficacy endpoint was SVR12, defined as HCV RNA<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 227A<br />

39<br />

SVR12 results from the Phase II, open-label IMPACT<br />

study of simeprevir (SMV) in combination with daclatasvir<br />

(DCV) and sofosbuvir (SOF) in treatment-naïve and<br />

-experienced patients with chronic HCV genotype 1/4<br />

infection and decompensated liver disease<br />

Eric Lawitz 1 , Fred Poordad 1 , Julio A. Gutierrez 1 , Thomas Kakuda 2 ,<br />

Gaston Picchio 3 , Greet Beets 4 , Ann Vandevoorde 4 , Peter Van<br />

Remoortere 2 , Bert Jacquemyn 4 , Gemma Quinn 4 , Donghan Luo 2 ,<br />

Sivi Ouwerkerk-Mahadevan 5 , Leen Vijgen 4 , Veerle Van Eygen 4 ,<br />

Maria Beumont 5 ; 1 Texas Liver Institute, University of Texas Health<br />

Science Center, San Antonio, TX; 2 Janssen Research & Development,<br />

LLC, Titusville, NJ; 3 Infectious Diseases, Janssen Research<br />

& Development, San Francisco, CA; 4 Janssen Infectious Diseases<br />

BVBA, Beerse, Belgium; 5 Janssen Research & Development,<br />

Beerse, Belgium<br />

Purpose: The ongoing IMPACT study (NCT02262728)<br />

assessed SMV+DCV+SOF in HCV genotype (GT)1/4-infected<br />

patients (pts) with decompensated liver disease. We present<br />

sustained virologic response 12 weeks (w) after end of treatment<br />

(SVR12) from an interim analysis (IA). Methods: Treatment-naïve<br />

or PegIFN±RBV treatment-experienced cirrhotic<br />

adults with Child-Pugh (CP) score


228A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

patients ± cirrhosis ± HIV co-infection. Injection drug use is the<br />

major risk factor for HCV transmission in many settings; however,<br />

data on interferon-free treatment outcomes among PWID<br />

are very limited. C-EDGE COSTAR is a Phase 3 trial evaluating<br />

efficacy and safety of GZR/EBR among treatment naïve<br />

HCV GT1/4/6-infected patients ± cirrhosis who are receiving<br />

opioid agonist therapy (OAT) (methadone or buprenorphine).<br />

METHODS: This double-blind, placebo-controlled study randomized<br />

patients 2:1 to an immediate treatment group (ITG) or<br />

a deferred treatment group (DTG). The ITG received GZR/<br />

EBR for 12 weeks; the DTG received placebo for 12 weeks,<br />

followed by 12 weeks of open label GZR/EBR. Use of non-prescribed<br />

drugs was monitored by urine drug screening, but<br />

was not exclusionary to study participation. HCV RNA levels<br />

were measured using COBAS TaqMan v2.0 (lower limit of<br />

quantitation 5x<br />

ULN after normalization while on therapy. SVR12 results will<br />

be presented. CONCLUSION: Preliminary data indicate that<br />

GZR/EBR is safe and highly effective in patients with chronic<br />

HCV GT1, 4, or 6 receiving OAT, and supports enhanced<br />

efforts to address barriers to HCV treatment access for PWID.<br />

Disclosures:<br />

Gregory Dore - Board Membership: Gilead, Merck, Abbvie, Bristol-Myers<br />

Squibb; Grant/Research Support: Gilead, Merck, Abbvie, Bristol-Myers Squibb;<br />

Speaking and Teaching: Gilead, Merck, Abbvie, Bristol-Myers Squibb<br />

Frederick Altice - Grant/Research Support: Gileaad, NIH, NIDA, SAMHSA,<br />

HRSA; Speaking and Teaching: Merck, Bristol Myers Squibb, Gilead, Rush Practice<br />

Point Communications<br />

Alain H. Litwin - Advisory Committees or Review Panels: Merck, Gilead; Grant/<br />

Research Support: Merck, Gilead<br />

Olav Dalgard - Advisory Committees or Review Panels: MSD, Janssen Cilag,<br />

Medivir, Gilead, Abbvie; Grant/Research Support: MSD, Medivir, Gilead<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Anne Luetkemeyer - Grant/Research Support: Gilead, Abbvie, Pfizer, Bristol<br />

Myers Squibb, Merck<br />

Ronald Nahass - Advisory Committees or Review Panels: Gilead, MErck, Janssen,<br />

BMS; Grant/Research Support: Gilead, Merck, Janssen, BMS; Speaking and<br />

Teaching: Gilead, Merck, Janssen<br />

Cheng-Yuan Peng - Advisory Committees or Review Panels: AbbVie, BMS, Gilead,<br />

MSD, Roche<br />

Brian Conway - Advisory Committees or Review Panels: Vertex Pharmaceuticals,<br />

Merck, Boehringer Ingelheim, Jannsen Pharmaceuticals; Grant/Research Support:<br />

Vertex Pharmaceuticals, Merck, Boehringer Ingelheim, Jannsen Pharmaceuticals,<br />

AbbVie, Gilead Sciences, Gilead Sciences<br />

Jason Grebely - Advisory Committees or Review Panels: Merck, Gilead; Grant/<br />

Research Support: Merck, Gilead, Abbvie, BMS<br />

Anita Y. Howe - Employment: Merck Research Laboratory<br />

Bach-Yen T. Nguyen - Employment: Merck<br />

Janice Wahl - Employment: Merck & Co,<br />

Eliav Barr - Employment: Merck<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

The following authors have nothing to disclose: Oren Shibolet, Heather L. Platt<br />

41<br />

98%–100% SVR4 in HCV Genotype 1 Non-Cirrhotic<br />

Treatment-Naïve or Pegylated Interferon/Ribavirin Null<br />

Responders With the Combination of the Next Generation<br />

NS3/4A Protease Inhibitor ABT-493 and NS5A<br />

Inhibitor ABT-530 (SURVEYOR-1)<br />

Fred Poordad 2 , Franco Felizarta 3 , Armen Asatryan 1 , Tarek I. Hassanein<br />

4 , Humberto I. Aguilar 5 , Jacob P. Lalezari 6 , J. Scott Overcash<br />

7 , Teresa Ng 1 , Ran Liu 1 , Chih-Wei Lin 1 , Federico J. Mensa 1 ,<br />

Jens Kort 1 ; 1 AbbVie, Inc., North Chicago, IL; 2 Texas Liver Institute,<br />

University of Texas Health Science Center, San Antonio, TX; 3 Private<br />

practice, Bakersfield, CA; 4 Southern California GI and Liver<br />

Centers and Southern California Education and Research Center,<br />

Coronado, CA; 5 Louisiana Research Center, Shreveport, LA;<br />

6 Quest Clinical Research, San Francisco, CA; 7 eStudySite, San<br />

Diego, CA<br />

Background: The next-generation HCV direct-acting antivirals<br />

(DAAs) ABT-493, an NS3/4A protease inhibitor identified by<br />

AbbVie and Enanta, and ABT-530, an NS5A inhibitor, are<br />

characterized by potent pangenotypic in vitro antiviral activity<br />

against major HCV genotypes (GTs), including activity against<br />

key known resistance-associated variants and a high barrier<br />

to resistance selection. Monotherapy with ABT-493 or ABT-<br />

530 resulted in a mean 4 log 10<br />

IU/mL decline from baseline<br />

in HCV plasma viral load in GT1-infected subjects with and<br />

without compensated cirrhosis. Purpose: In this phase 2 study,<br />

treatment with ABT-493 and ABT-530 for 12 weeks is evaluated<br />

in HCV GT1-infected subjects without cirrhosis. Efficacy<br />

and safety results are reported here. Methods: Non-cirrhotic<br />

GT1-infected treatment-naïve (TN) or pegylated interferon/ribavirin<br />

(pegIFN/RBV) null responder subjects received once-daily<br />

ABT-493 200 mg + ABT-530 120 or 40 mg for 12 weeks,<br />

and subsequently were followed for 24 weeks. The primary<br />

efficacy endpoint was sustained virologic response 12 wks<br />

after the last dose of study drug (SVR12). Safety was evaluated<br />

by adverse event (AE) monitoring, laboratory testing,<br />

and other standard assessments. Results: 79 subjects (male,<br />

52%; median [range] age, 54.0 [26.0–70.0] years; GT1a,<br />

81%; GT1b, 19%; TN, 63%; pegIFN/RBV null responders,<br />

37%; fibrosis >F2, 25%; median [range] HCV RNA log 10<br />

IU/<br />

mL, 6.8 [4.4–7.5]) were enrolled, 40 received ABT-493 200<br />

mg+ABT-530 120 mg and 39 received ABT-493 200 mg and<br />

ABT-530 40 mg. SVR 4 weeks after the last dose of study drug<br />

(SVR4) is available in all 79 subjects. SVR4 was achieved in<br />

29 of 29 (100%) pegIFN/RBV null responders and 49 of 50<br />

(98%) TN subjects. One TN subject relapsed at post-treatment<br />

week 4; further characterization of this relapse and complete<br />

post-treatment week 12 data (SVR12) will be available. There<br />

were no treatment-related serious AEs or clinically relevant<br />

laboratory findings. The most common AEs (reported in >5%<br />

of subjects) were fatigue, headache, nausea, diarrhea, and<br />

anxiety. Conclusions: Once-daily 12-week treatment with the<br />

next-generation HCV DAAs ABT-493 and ABT-530 of GT1<br />

infection in non-cirrhotic TN and pegIFN/RBV null responders<br />

resulted in high SVR4 (98%–100%) rates. One treatment<br />

relapse was observed. Based on these encouraging results,<br />

the SURVEYOR-1 study with once-daily ABT-493 and ABT-530<br />

has been expanded to include GT1 subjects with compensated<br />

cirrhosis and evaluate a shorter, 8 week treatment duration in<br />

non-cirrhotic subjects.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 229A<br />

Disclosures:<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Franco Felizarta - Grant/Research Support: AbbVie, Gilead, Janssen, Merck,<br />

BMS, Boehringer-Ingelheim, Vertex, Roche; Speaking and Teaching: AbbVie,<br />

Gilead, Janssen, Merck<br />

Armen Asatryan - Employment: AbbVie<br />

Tarek I. Hassanein - Advisory Committees or Review Panels: AbbVie, Bristol-Myers<br />

Squibb; Grant/Research Support: AbbVie Pharmaceuticals, Obalon, Bristol-Myers<br />

Squibb, Eiasi Pharmaceuticals, Gilead Sciences, Janssen R&D, Idenix<br />

Pharmaceuticals, Ikaria Therapeutics, Merck Sharp & Dohme, NGM BioPharmaceuticals,<br />

Ocera Therapeutics, Salix Pharmaceuticals, Sundise, TaiGen Biotechnology,<br />

Takeda Pharmaceuticals, Vital Therapies, Tobria; Speaking and<br />

Teaching: Baxter, Bristol-Myers Squibb, Gilead, Salix, AbbVie<br />

Humberto I. Aguilar - Speaking and Teaching: abbvie, gilead<br />

Teresa Ng - Employment: AbbVie; Patent Held/Filed: AbbVie; Stock Shareholder:<br />

AbbVie<br />

Ran Liu - Employment: Abbvie<br />

Chih-Wei Lin - Employment: Abbvie<br />

Federico J. Mensa - Employment: Abbvie Inc.; Stock Shareholder: Abbvie Inc.<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

The following authors have nothing to disclose: Jacob P. Lalezari, J. Scott Overcash<br />

42<br />

An Integrated Analysis of 402 Compensated Cirrhotic<br />

Patients With HCV Genotype (GT) 1, 4 or 6 Infection<br />

Treated With Grazoprevir/Elbasvir<br />

Ira M. Jacobson 1 , Eric Lawitz 2 , Paul Y. Kwo 3 , Christophe Hezode 4 ,<br />

Cheng-Yuan Peng 5 , Anita Y. Howe 6 , Peggy Hwang 6 , Janice<br />

Wahl 6 , Michael Robertson 6 , Eliav Barr 6 , Barbara A. Haber 6 ;<br />

1 Medicine, Mt. Sinai Beth Israel, New York, NY; 2 Texas Liver Institute,<br />

University of Texas Health Science Center, San Antonio, TX;<br />

3 Indiana University School of Medicine, Indianapolis, IN; 4 Henri<br />

Mondor Hospital, Creteil, France; 5 China Medical University Hospital,<br />

Taichung, Taiwan; 6 Merck & Co., Inc., Kenilworth, NJ<br />

Purpose: Treatment of cirrhotic HCV-infected patients (pts) often<br />

requires ribavirin (RBV) and long durations of therapy. In the<br />

C-EDGE trials, 12-16 weeks’ therapy with grazoprevir 100<br />

mg/elbasvir 50 mg (GZR/EBR) ± RBV was highly efficacious<br />

in a diverse population of pts. An analysis of the efficacy/<br />

safety of GZR/EBR ± RBV among compensated cirrhotics in<br />

the phase 2/3 GZR/EBR program was conducted. Methods:<br />

Treatment-naive (TN) and treatment-experienced (TE) pts with<br />

compensated cirrhosis received GZR/EBR ±RBV for 12 (TN) or<br />

12-18 weeks (TE) in 6 <strong>studies</strong>. The primary end point was HCV<br />

RNA


230A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

could predict the presence of various cancer entities in patients<br />

with hepatocellular, colorectal and non-small cell lung carcinoma<br />

(HCC, CRC, NSCLC, resp.), as a tool for cancer screening<br />

and therapy response monitoring. Methods: MPs were<br />

isolated from serum of patients with HCC, CRC, NSCLC and<br />

from healthy donors. MP profiling was done after differential<br />

ultracentrifugation using FACS analysis for the presence and<br />

quantity of AnnexinV + /EpCAM + and AnnexinV + /EpCAM + /<br />

CD147 + taMPs. Results: AnnexinV + /EpCAM + and AnnexinV<br />

+ /EpCAM + /CD147 + taMPs highly significantly separated<br />

patients with carcinomas from healthy controls. The calculated<br />

cut-off values for AnnexinV + /EpCAM + taMPs were 36.6<br />

/1AnnexinV + MPs in NSCLC, 35.8/1kAnnexinV + MPs in CRC<br />

and 31.4/1kAnnexinV + taMPs in HCC. Additionally, AnnexinV<br />

+ /EpCAM + /CD147 + taMPs indicated the presence of cancer<br />

as well as AnnexinV + /EpCAM + taMPs alone. AnnexinV + /<br />

EpCAM + taMPs and AnnexinV + /EpCAM + /CD147 + taMPs<br />

were increased 2.4-fold (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 231A<br />

respectively. The median (IQR) time intervals were 46.5 (62)<br />

and 47.5 (87) days between MRE and biopsy and between<br />

ARFI and biopsy, respectively. For diagnosing any fibrosis,<br />

MRE had an area under ROC curve (AUROC) of 0.799 (95%<br />

CI, 0.723-0.875), which was significantly (p=0.012) higher<br />

than ARFI, which had an AUROC of 0.664 (95% CI, 0.568-<br />

0.760). The AUROCs of MRE for diagnosing ≥ stages 2, 3,<br />

and 4 fibrosis were 0.885 (95% CI, 0.816-0.953), 0.934<br />

(95% CI, 0.863-1.000), and 0.882 (95% CI, 0.729-1.000),<br />

and those for ARFI were 0.848 (95% CI, 0.776-0.921), 0.896<br />

(95% CI, 0.824-0.968), and 0.862 (95% CI, 0.721-1.000).<br />

MRE had a higher AUROC than ARFI for discriminating dichotomized<br />

fibrosis stages at every dichotomization cutpoint, but<br />

the difference in AUROC was smaller as the dichotomization<br />

cutpoint increased. Conclusion: In a head-to-head comparison<br />

and using contemporaneous liver biopsy as the gold standard,<br />

MRE provides significantly higher diagnostic accuracy than<br />

ARFI for diagnosing any fibrosis in NAFLD patients.<br />

Disclosures:<br />

Claude B. Sirlin - Grant/Research Support: GE, Pfizer, Bayer, Guerbet, Siemens<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

The following authors have nothing to disclose: Jeffrey Y. Cui, Elhamy Heba,<br />

Carolyn Hernandez, William Haufe, Jonathan Hooker<br />

46<br />

The predictive value of 13 C-Methacetin breath test (MBT)<br />

for liver transplantation (LT) and death: a seven year<br />

follow-up study of 206 chronic hepatitis C patients<br />

Oliver Goetze 2,1 , Monika Breuer 1 , Andreas Geier 2,1 , Michael<br />

Fried 1 , Achim Weber 4 , Yaron Ilan 3 , Beat Mullhaupt 1 ; 1 Department<br />

of Gastroenterology & Hepatology, University Hospital Zurich,<br />

Zurich, Switzerland; 2 Division of Hepatology, University Hospital<br />

Würzburg, Würzburg, Germany; 3 Department of Medicine,<br />

Hebrew University-Hadassah Medical Center, Jerusalem, Israel;<br />

4 Institute of Surgical Pathology, University Hospital Zurich, Zurich,<br />

Switzerland<br />

In patients with chronic liver disease, currently used biomarkers<br />

demonstrate limited mid- to long-term predictive values to<br />

forecast deterioration of liver function leading to liver transplantation<br />

(LT) or liver related death. Aim: To determine the<br />

predictive value of the 13 C-Methacetin Breath Test (MBT)<br />

(BreathID ® ; Exalenz) in forseeing deterioration of liver function<br />

leading to LT and/or death in patients with chronic HCV infection.<br />

Methods: A total of 265 chronic HCV infected patients<br />

were enrolled, out of which 206 patients (71 females) without<br />

co-infections and/or liver comorbidities were considered for<br />

this analysis. The average BMI was 24.8 kg/m 2 (SD 4.0 kg/<br />

m 2 ), 183 patients underwent liver biopsy and ISHAK scoring<br />

and 196 patients underwent MBT. 13 C-Methacetin peak percent<br />

dose recovery rate (PDR-Peak) was used as determinant<br />

parameter for this analysis. Patients were followed for up to<br />

7 years according to standard of care. During the course of<br />

follow-up some patients underwent antiviral treatment as part of<br />

routine patient management. Primary analysis was performed<br />

on all patients up to one of the following: liver-transplantation,<br />

death or response to antiviral treatment, resulting in 161<br />

and 158 patients with biopsy and/or MBT results respectively.<br />

Predictive value was assessed using ROC curve and survival<br />

analyses. Results: In seven years follow up, PDR-Peak, using<br />

cut-off of 20%/h, predicted LT and/or death with HR=12.07<br />

(95%CI 3.48;41.84),p


232A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

black tea. The mean ALT noted in 657 individuals within 3<br />

months of TE was 60.6 U/L. The average BMI was 25.7 kg/m 2<br />

and weight was 71.7 kg. Average daily alcohol consumption<br />

was 5 g/day. The average liver stiffness for the 1155 individuals<br />

was 8.4 kPa. CAP was available from 2013 in 719 individuals<br />

with average steatosis measures of 214 dB/m. After<br />

MLR taking into account age, alcohol, smoking and weight,<br />

individuals with HCV benefited from coffee intake with LSM<br />

on average 2 kPa less with intake of 2 or more cups of coffee<br />

daily (p = 0.026). This was not seen in those with HBV or<br />

NAFLD. In NAFLD a decrease in CAP of 23 dB/m (9%) was<br />

noted in individuals who drank ≥2 cups of coffee (p = 0.032).<br />

There was no effect of coffee consumption on CAP in other<br />

groups. Conclusions: Individuals with self-reported daily coffee<br />

intake ≥2 cups of coffee/ day had a significant decrease in<br />

liver stiffness, which was not seen in those with HBV or NAFLD.<br />

Similar daily coffee intake resulted in significant decrease in<br />

liver steatosis in NAFLD. These findings are consistent with the<br />

literature to date and add to the growing body of evidence<br />

suggesting coffee may be a beneficial supplement in some<br />

liver diseases.<br />

Disclosures:<br />

The following authors have nothing to disclose: Alexander Hodge, Sarah P. Lim,<br />

Evan Goh, Ophelia H. Wong, Philip Marsh, Virginia Knight, Yong Song<br />

48<br />

Serum Lipidomic Profiling for Screening Potential Biomarkers<br />

of Liver Cirrhosis among Patients with Chronic<br />

Hepatitis C<br />

Ana Maria Passos-Castilho 1 , Maria Lucia Ferraz 2 , Valdemir M.<br />

Carvalho 3 , Karina H. Cardozo 3 , Luciana Kikuchi 4 , Aline Chagas 4 ,<br />

João Renato R. Pinho 4 , Michele S. Gomes-Gouvêa 4 , Fernanda<br />

Malta 4 , Flair J. Carrilho 4 , Celso Granato 1,3 ; 1 Division of Infectious<br />

Diseases, Federal University of Sao Paulo, Sao Paulo, Brazil;<br />

2 Department of Gastroenterology, Federal University of Sao Paulo,<br />

Sao Paulo, Brazil; 3 Fleury SA Group, Sao Paulo, Brazil; 4 Department<br />

of Gastroenterology, University of Sao Paulo School of Medicine,<br />

Sao Paulo, Brazil<br />

Background: Liver cirrhosis (LC) is the 14th most common cause<br />

of death in adults worldwide, resulting in 1.03 million deaths<br />

per year. Chronic hepatitis C (CHC) is one of the main causes<br />

of LC. Non-invasive markers of fibrosis have been increasingly<br />

studied as they may represent a more informative, safe and<br />

accessible diagnostic and/or screening tool for at-risk patients.<br />

Aim: To explore the serum lipidome profiles of LC to identify<br />

potential biomarkers. Methods: A total of 182 subjects were<br />

enrolled from 2011 to 2014. An ultraperformance liquid chromatography–mass<br />

spectrometry (UPLC-MS) lipidomic method<br />

was used to characterize serum profiles from LC (n=59), CHC<br />

(n=65) and healthy subjects (n=58). Patients were diagnosed<br />

by clinical, laboratory and imaging evidence of LC while<br />

healthy controls had normal liver function and no infectious diseases.<br />

Reversed-phased analysis was performed on a Waters<br />

ACQUITY IClass UPLC system coupled to a Waters Synapt-MS<br />

hidrid quadrupole-time of flight mass spectrometer operating<br />

in the positive ion electrospray mode with a mass scan range<br />

200–1200 m/z for data acquisition in continuous mode. All<br />

data were processed with Progenesis software (Waters). The<br />

resulting multivariate data set was analyzed with MetaboAnalyst<br />

(TMIC) using supervised partial least-squares discriminate<br />

analysis (PLS-DA). Potential biomarkers were selected according<br />

to the variable importance in the projection (VIP) and statistical<br />

significance was evaluated using Mann-Whitney test. The<br />

receiver operating characteristic (ROC) curve was performed<br />

to assess the accuracy of selected biomarkers in LC diagnosis.<br />

To identify the potential biomarkers, the HMDB and Lipid<br />

Maps databases were queried with the exact mass. Results:<br />

The UPLC–MS-based serum lipidomic profile detected 51 of<br />

59 LC cases, performing with a diagnostic accuracy of 87%.<br />

The 3 potential biomarkers differentiated LC from CHC individually<br />

with 76 to 88% sensitivity and 72 to 74% specificity.<br />

They all presented with fold change higher than 2 and p <<br />

0.0001 in univariate analyses. The combination of the 3 potential<br />

biomarkers resulted in an area under the curve of 0.95,<br />

86% sensitivity and 88% specificity. The model was validated<br />

with 1000 permutation tests (p < 0.001) to prevent overfitting.<br />

3-Hydroxy-cis-5-tetradecenocylcarnitine, a hydroxy fatty acid<br />

involved in many metabolic and proinflammatory pathways,<br />

exhibited a significant increase in LC and a decrease in CHC<br />

and was proposed as an important indicator of LC. Conclusions:<br />

UPLC-MS lipid profiling proved to be an efficient and<br />

convenient tool for diagnosis and screening of LC in an at-risk<br />

population.<br />

Disclosures:<br />

João Renato R. Pinho - Employment: Hospital Israelita Albert Einstein<br />

The following authors have nothing to disclose: Ana Maria Passos-Castilho,<br />

Maria Lucia Ferraz, Valdemir M. Carvalho, Karina H. Cardozo, Luciana Kikuchi,<br />

Aline Chagas, Michele S. Gomes-Gouvêa, Fernanda Malta, Flair J. Carrilho,<br />

Celso Granato<br />

49<br />

Comparison of Liver Transplant-related Survival Benefit<br />

in Patients With vs Without Hepatocellular Carcinoma in<br />

the United States.<br />

George N. Ioannou 1,2 , Kristin Berry 1 ; 1 Veterans Affairs Puget<br />

Sound Healthcare System, Seattle, WA; 2 University of Washington,<br />

Seattle, WA<br />

Background and Aims Patients with T2 hepatocellular carcinoma<br />

(HCC) can obtain an exception allowing them to undergo<br />

liver transplantation at much lower actual Model for End Stage<br />

Liver Disease (MELD) scores than patients without HCC. We<br />

aimed to compare patients transplanted with and without HCC<br />

with regards to transplantation-related survival benefit. Methods<br />

We modeled the post-transplant survival of adult, first-time<br />

liver transplant recipients with (n=9135) or without (n=25,890)<br />

HCC from 2002-2013 using Cox proportional hazards regression.<br />

We modeled waitlist survival of patients listed for transplantation<br />

with (n=15,605) or without (n=85,229) HCC using<br />

competing risks analysis and a combined outcome of death<br />

or liver failure, defined as MELD score ≥30. We used these<br />

survival models to calculate monthly transition probabilities and<br />

5-year life expectancies. Survival benefit was calculated as the<br />

difference between post-transplant and waitlist life expectancy.<br />

Results Five-year survival benefit increased with actual MELD<br />

score for both patients with and without HCC from just a few<br />

months in patients with low MELD scores (6-8) to 4 years in<br />

patients with the highest MELD scores (36-40). The survival<br />

benefit of patients with HCC was similar to that of patients<br />

without HCC who had the same actual MELD score, irrespective<br />

of tumor burden or serum alpha fetoprotein level. Because<br />

patients with HCC were transplanted at a lower mean MELD<br />

score (13.3±6.2) than patients without HCC (21.8±8.0), a<br />

dramatically lower mean 5-year survival benefit was achieved<br />

by transplanting patients with HCC (0.12 years/patient) than<br />

patients without HCC (1.47 years/patient). Conclusions The<br />

HCC MELD exception policy unintentionally resulted in a dramatic<br />

reduction in transplant-related survival benefit.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 233A<br />

Five-year transplant-related survival benefit as a function of MELD<br />

score shown separately in patients with or without HCC<br />

considered, wait time 18 months (HR 1.6, 95% CI<br />

1.01-2.5, p=0.043) and AFP >400 ng/mL at HCC diagnosis<br />

(HR 3.0, 95% CI 1.7-5.5, p100 ng/mL at LT<br />

(HR 1.6, 95% CI 1.04-2.6, p


234A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

LT. Therefore, MHE, i) should not to be ignored and ii) deserves<br />

to be treated in order to reduce the risk of neurological complications<br />

occurring post-LT.<br />

Disclosures:<br />

The following authors have nothing to disclose: Marc-André Clément, Cristina R.<br />

Bosoi, Mélanie Tremblay, Chantal Bemeur, Christopher F. Rose<br />

52<br />

The Functional Basis for Cognitive Improvement after<br />

Liver Transplant: A Prospective Brain MRI Study<br />

Vishwadeep Ahluwalia, James Wade, HoChong Gilles, Melanie<br />

White, Ariel Unser, Edith A. Gavis, Mohammad S. Siddiqui, Velimir<br />

A. Luketic, Arun J. Sanyal, Puneet Puri, Douglas M. Heuman,<br />

Scott C. Matherly, R. Todd Stravitz, Richard K. Sterling, Michael<br />

Fuchs, Jasmohan S. Bajaj; VCU and McGuire VAMC, Richmond,<br />

VA<br />

Cirrhotics have varying degrees of cognitive recovery post-liver<br />

transplant (LT) depending on their prior hepatic encephalopathy(HE)<br />

status but the mechanism is unclear. Aim: Define the<br />

functional basis of cognitive change in LT recipients. Method:<br />

Cirrhotics on the transplant list underwent standard cognitive<br />

testing & questionnaires regarding depression (Beck Depression,<br />

BDI), QOL (Sickness impact profile, SIP has psych &<br />

physical components) pre/6 months post-LT. Cognitive tests<br />

were number connection A/B (NC), digit symbol (DS) & block<br />

design (BD) which test for psychomotor speed, set-shifting and<br />

visuo-motor coordination. We also performed brain functional<br />

MRI while subjects underwent the inhibitory control test (ICT)<br />

within the scanner. ICT tests for response inhibition, working<br />

memory & psychomotor speed. The outcomes are lures (that<br />

need to be inhibited) & targets (need to be responded to).<br />

Changes in brain activation on correct lure inhibition & target<br />

response were compared before & after LT. Results: We<br />

enrolled 67 cirrhotics of which 11 died prior, 12 still await &<br />

5 pts were


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 235A<br />

may help refine post-LT HCC surveillance strategies and identify<br />

patients who would derive the most benefits from future post-LT<br />

adjuvant therapies.<br />

RETREAT Score<br />

0 points for: AFP


236A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

from BM macrophages and Kupffer cells in a dose-dependent<br />

manner; 5-10nM MCC950 abrogated such IL-1β release. Conclusions:<br />

These data demonstrate that MCC950, a potent and<br />

selective NLRP3 inflammasome inhibitor, prevents or reverses<br />

inflammation, injury and fibrosis in two different murine models<br />

of NASH. These effects could be explained, at least in part, by<br />

the effective abrogation of cholesterol crystal-induced NLRP3<br />

activation and IL-1β release in BM macrophages and Kupffer<br />

cells. Targeting NLRP3 is a logical new direction in NASH<br />

pharmacotherapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Auvro R. Mridha, Alexander<br />

Wree, Avril A. Robertson, Narci C. Teoh, Matthew A. Cooper, Ariel E. Feldstein,<br />

Geoffrey C. Farrell<br />

56<br />

Role of Fatty Acid Desaturase 1 (FADS1) in Nonalcoholic<br />

Fatty Liver Disease (NAFLD)<br />

Shaminie Athinarayanan 1 , Yang-Yi Fan 2 , Robert Chapkin 2 , Wanqing<br />

Liu 1 ; 1 Medicinal Chemistry and Molecular Pharmacology,<br />

Purdue University, West Lafayette, IN; 2 Center for Translational<br />

Environmental Health Research, Texas A&M, College Station, TX<br />

Background: NAFLD and nonalcoholic steatohepatitis (NASH)<br />

are common chronic liver diseases. Previous <strong>studies</strong> have<br />

demonstrated that decreased hepatic and blood levels of<br />

n-3 polyunsaturated fatty acids (PUFA) are associated with<br />

NAFLD/NASH. The FADS1 gene encodes delta-5 desaturase<br />

(D5D), one of the rate-limiting enzymes involved in n-3 PUFA<br />

metabolism. We hypothesized that FADS1 is involved in the<br />

pathogenesis of NAFLD/NASH via modulating hepatic lipid<br />

homeostasis and other pathways. Methods: In order to corroborate<br />

our hypothesis, we used both the HepG2 cell line<br />

and a Fads1 knockout (null) mouse model. The FADS1 gene<br />

was knocked down using siRNA in HepG2 cells. Palmitic acid<br />

(PA)-induced lipid accumulation was compared between the KD<br />

cells and the control siRNA group. In addition, RNA-seq based<br />

hepatic transcriptome data were compared between the null<br />

(n=4) and wild-type (n=4) mice. Differential gene expression<br />

data were analyzed using EdgeR from Bioconductor. Detailed<br />

pathway analyses were performed by combining both a targeted<br />

approach to examine well-established lipid homeostasis<br />

pathways as well as a global, unsupervised enrichment analysis<br />

using Ingenuity Pathway Analysis (IPA). Results: FADS1<br />

knock down in HepG2 cells significantly increased PA-induced<br />

intracellular lipid accumulation by 60% (p < 0.001). Liver samples<br />

from Fads1 null mice exhibited a significantly (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 237A<br />

58<br />

Hepatocytes Release Ceramide-rich Proinflammatory<br />

Extracellular Vesicles in an IRE1alpha-dependent manner<br />

Eiji Kakazu 1,2 , Amy S. Mauer 1 , Harmeet Malhi 1 ; 1 Gastroenterology<br />

and Hepatology, Mayo Clinic, Rochester, MN; 2 Department of<br />

Community Medical Supports, Tohoku University, Aobaku, Japan<br />

Palmitate (PA), a lipotoxic free fatty acid, is implicated in<br />

hepatocyte apoptosis, macrophage-mediated liver inflammation,<br />

and activation of the IRE1alpha branch of the endoplasmic<br />

reticulum (ER) stress response - all key features of<br />

progressive nonalcoholic steatohepatitis (NASH). Recently,<br />

extracellular vesicles (EVs) have been implicated in NASH.<br />

However, the exact pathways of hepatocyte EV generation<br />

under lipotoxic conditions and their downstream effects on<br />

macrophages are undefined. Therefore, we hypothesized that<br />

hepatocytes release proinflammatory PA-induced EVs via an<br />

ER stress response. To test this hypothesis we employed immortalized<br />

mouse hepatocytes (IMH) from wild-type (WT) or IRE1a<br />

knockout (KO) mice, Huh 7 and primary mouse hepatocytes<br />

(PMH). Cells were treated with 400μM PA or thapsigargin<br />

(Tg) to induce ER stress. EVs were isolated from cell culture<br />

supernatants by ultracentrifugation and characterized by immunofluorescence<br />

and western blotting for EV markers, electron<br />

microscopy (EM) and by nanoparticle tracking analysis (NTA)<br />

for morphology and size. Ceramides were measured by mass<br />

spectrometry. CRISPR/Cas9 technology was used to delete<br />

XBP1. Results: PA and Tg significantly increased EV release<br />

in all three hepatocyte cells tested (IMH, Huh7 and PMH). The<br />

released EVs were 100nm in size, decorated with EV markers<br />

CD63, TSG101 and LAMP1, and demonstrated characteristic<br />

cup-shaped morphology on EM. As PA and Tg activate an ER<br />

stress response, we next tested cells lacking in each of the three<br />

canonical ER stress sensors, IRE1alpha, ATF6alpha and PERK.<br />

Both PA and Tg induced a significant EV release in cells lacking<br />

ATF6 alpha or PERK; however, PA and Tg-induced EV release<br />

was significantly suppressed in IRE1alpha KO cells. We tested<br />

the involvement of XBP1 signaling in this phenomenon and<br />

found a reduction in EV response in cells lacking XBP1. Next<br />

we measured ceramides in PA-induced EVs as they are implicated<br />

in EV biogenesis. PA-induced EVs were enriched in C16<br />

ceramide; this enrichment occurred in an IRE1alpha-dependent<br />

manner. Inhibition of de novo ceramide synthesis by myriocin<br />

ameliorated PA-induced EV release, and exogenous C16 ceramide<br />

resulted in EV release in IRE1alpha KO and WT cells.<br />

Lastly, PA-stimulated EVs were chemotactic to macrophages.<br />

Conclusions: PA induces C16 ceramide-enriched EV release<br />

in an IRE1alpha-dependent manner. PA-induced EVs stimulate<br />

macrophage chemotaxis and this may be a mechanism for the<br />

recruitment of macrophages to the liver under lipotoxic conditions.<br />

We hypothesize that interference with this macrophage<br />

recruitment response may be a therapeutic avenue in NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Eiji Kakazu, Amy S. Mauer,<br />

Harmeet Malhi<br />

59<br />

Deficiency of intestinal mucin-2 protects mice from<br />

diet-induced fatty liver disease and obesity<br />

Phillipp Hartmann, Caroline T. Seebauer, Magdalena Mazagova,<br />

Angela Horvath, Lirui Wang, Cristina Llorente-Izquierdo, Katharina<br />

Brandl, Samuel B. Ho, David A. Brenner, Bernd Schnabl; University<br />

of California San Diego, La Jolla, CA<br />

Background: Non-alcoholic fatty liver disease (NAFLD) and<br />

obesity are characterized by altered gut microbiota, inflammation,<br />

and gut barrier dysfunction. The aim of this study was to<br />

investigate the role of mucin-2 (Muc2) as the major component<br />

of the intestinal mucus layer in the development of fatty liver<br />

disease and obesity. Methods and Results: We studied experimental<br />

fatty liver disease and obesity induced by high-fat diet<br />

(HFD) feeding for 16 weeks in wild-type (WT) littermate and<br />

mucin-2 knockout mice. Muc2 deficiency protected mice from<br />

HFD-induced obesity and fatty liver disease as evidenced by a<br />

significantly reduced body weight and lower hepatic triglyceride<br />

concentrations, respectively. Furthermore, Muc2 -/- mice displayed<br />

significantly decreased weights of subcutaneous and<br />

epididymal white adipose tissue and brown adipose tissue<br />

relative to WT mice on HFD. There was no difference in food<br />

intake and fecal triglyceride content between the HFD groups<br />

that would explain differences in phenotype. HFD-fed Muc2 -/-<br />

mice exhibited improved glucose tolerance and responsiveness<br />

to insulin, as evidenced by intraperitoneal glucose tolerance<br />

and insulin tolerance tests, as well as reduced inflammation<br />

in white adipose tissue than WT littermates. Furthermore, they<br />

showed an upregulated expression of key genes involved in<br />

lipolysis, such as adipose triglyceride lipase (ATGL) and hormone-sensitive<br />

lipase (HSL), and in fatty acid β-oxidation, e.g.<br />

palmitoyl acyl-CoA oxidase 1 (ACOX1), in white adipose tissue<br />

compared with WT mice after 16 weeks of HFD feeding.<br />

Interleukin-22 (IL-22) is not only crucial for maintaining intestinal<br />

homeostasis, but administration of IL-22 to obese mice has<br />

been shown to improve insulin sensitivity, decrease chronic<br />

inflammation and reduce body weight. Most interestingly,<br />

Muc2 -/- mice showed increased intestinal gene expression<br />

of IL-22 and IL-22 target genes Reg3b and Reg3g relative to<br />

WT mice fed HFD. Plasma levels of IL-22 were significantly<br />

higher in HFD-fed Muc2 -/- mice as compared with WT mice.<br />

After bypassing the mucus layer by intraperitoneal injections<br />

of bacteria-derived flagellin, differences in IL-22 plasma levels<br />

between HFD-fed WT and Muc2 -/- mice vanished, suggesting<br />

that the mucosal immune system produces increased amounts<br />

of IL-22 in the absence of a protective mucus barrier. Conclusion:<br />

An impaired gut barrier elicits a strong intestinal immune<br />

response in Muc2 deficient mice, which is characterized by the<br />

induction of IL-22. Increased systemic IL-22 mediates beneficial<br />

metabolic and anti-inflammatory effects, and protects against<br />

obesity and fatty liver disease.<br />

Disclosures:<br />

Angela Horvath - Grant/Research Support: Instutut Allergosan<br />

Samuel B. Ho - Grant/Research Support: Gilead, Genentech; Speaking and<br />

Teaching: Prime Education, Inc<br />

The following authors have nothing to disclose: Phillipp Hartmann, Caroline T.<br />

Seebauer, Magdalena Mazagova, Lirui Wang, Cristina Llorente-Izquierdo, Katharina<br />

Brandl, David A. Brenner, Bernd Schnabl


238A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

60<br />

FoxO3 increases miR-34a to cause palmitate-induced<br />

cholangiocytes lipoapoptosis<br />

Sathish Kumar Natarajan, Cody J. Wehrkamp, Ashley M. Mohr,<br />

Mary A. Smith, Sizhao Lu, Duygu Dee Harrison-Findik, Justin L.<br />

Mott; UNIVERSITY OF NEBRASKA MEDICAL CENTER, Omaha,<br />

NE<br />

Abstract: Non-alcoholic fatty liver disease (NAFLD) patients<br />

have elevated plasma saturated free fatty acid (FFA) levels.<br />

We had recently demonstrated the critical role of FoxO3 and<br />

its downstream target, PUMA in FFA-induced cholangiocyte<br />

lipoapoptosis and provided insights for biliary damage in a<br />

subset of NAFLD patients. Here, we explored whether miR-34a<br />

have a role in cholangiocyte lipoapoptosis. Methods: Apoptosis<br />

was assessed by TUNEL, characteristic nuclear morphology<br />

staining with DAPI, and caspase 3/7 activity assay. miR-34a<br />

levels were measured using quantitative real time-PCR. FoxO3,<br />

cleaved PARP, cleaved caspase 3 and miR-34a target expression<br />

was examined by Western blot. Results: Treatment of cholangiocytes<br />

with palmitate (PA) showed 7-20 fold increase in<br />

caspase3/7 activity in normal immortalized cholangiocytes,<br />

H69 and cell lines derived from cholangiocarcinoma, KMCH,<br />

Mz-ChA-1 and HuCCT cells. PA treatment also resulted in a dramatic<br />

increase in the levels of caspase cleaved products such<br />

as cleaved PARP and cleaved caspase 3 suggesting cholangiocyte<br />

lipoapoptosis. Interestingly, treatment with PA significantly<br />

increased levels of miR-34a in cholangiocytes and hepatocyte<br />

cell line, Huh7. PA-induced increase in miR-34a levels were<br />

abolished in cholangiocytes transduced with FoxO3 shRNA,<br />

implicating that miR-34a is a transcriptional target of FoxO3.<br />

Further, cholangiocytes transfected with anti-miR-34a were protected<br />

from PA-induced lipoapoptosis suggesting a major role<br />

for miR-34a. Further, direct and indirect targets of miR34a such<br />

as Sirt1, receptor tyrosine kinase (cMET), Kruppel like factor 4<br />

(KLF4), fibroblast growth factor receptor (FGFR1 and 4) were<br />

all decreased in PA-treated cholangiocytes. In addition, our<br />

data also shows that cholangiocyte apoptosis and miR-34a are<br />

dramatically increased in the liver from mice fed with high-fat<br />

high-sucrose for 3 months compared to control fed animals.In<br />

conclusion, PA induces the expression of proapoptotic miR-34a<br />

via FoxO3 and increased miR-34a levels results in targeting<br />

protein deacetylase Sirt1 and anti-apoptotic proteins like cMET<br />

and KLF4 resulting in cholangiocyte lipoapoptosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sathish Kumar Natarajan, Cody<br />

J. Wehrkamp, Ashley M. Mohr, Mary A. Smith, Sizhao Lu, Duygu Dee Harrison-Findik,<br />

Justin L. Mott<br />

61<br />

Prior Blockade of TNF-α Signaling Promotes Liver<br />

Regeneration and Offers Strategies for Improving Outcomes<br />

after Liver Resection Surgery<br />

Fadi L. Jaber; GASTROENTEROLOGY AND LIVER DISEASES,<br />

ALBERT EINSTEIN COLLEGE OF MEDICINE, Bronx, NY<br />

The presumed beneficial role in liver regeneration (LR) of<br />

pro-inflammatory cytokines, such as TNF-α or IL6, contrasts<br />

with deleterious effects of these cytokines in other settings,<br />

e.g., after ischemia/reperfusion, inflammation, etc. Recently,<br />

cell transplantation-induced liver injury was shown to activate<br />

multiple chemokines/cytokines/receptors, which were essentially<br />

normalized after prior blockade of TNF-α by the drug,<br />

Etanercept (ETN), with superior transplanted cell engraftment<br />

and proliferation during liver repopulation in retrorsine (Ret)/<br />

partial hepatectomy (PH) rat model. To resolve what role TNF-α<br />

may serve in LR, we hypothesized that administering ETN to<br />

block TNF-α before PH will allow determination of whether LR<br />

will be impaired or if it would still proceed. Groups of F344<br />

rats were established for 2/3 PH alone or ETN followed by<br />

2/3 PH along with untreated controls. Time-course analysis<br />

at 6, 32 or 72 h and 7 d after these procedures showed ETN<br />

was well tolerated and liver/body weight ratios were actually<br />

superior rather than inferior in ETN-pretreated rats with PH,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 239A<br />

experience a fibrotic phase dominated by inflammatory macrophages,<br />

followed by a phase of fibrosis remodelling characterised<br />

by restorative macrophages. Gpnmb- mice show a<br />

higher level of circulating hepatic enzymes in the phase of<br />

fibrosis remodelling, associated with higher proliferation of<br />

parenchymal cells and higher levels of inflammatory cytokines<br />

in the tissue. Moreover those mice show a lower number of<br />

macrophages and T cells in the phase of fibrosis resolution,<br />

determined by immunohistochemistry. Flow cytometry analysis<br />

unveiled a possible skew to a pro-inflammatory phenotype in<br />

the recovering liver. Key cytokines have been investigated as<br />

a possible cross-regulator of phagocytosis and tissue proliferation.<br />

The role of IL6 as a pro-proliferative agent in the liver<br />

is well known. No information is available on its role as a<br />

modulator of phagocytosis. Both in vivo and in vitro IL6 shows<br />

a pro-phagocytic effect possibly through the activation of the<br />

STAT3 pathway downstream the IL6 receptor. Collectively<br />

our data suggest that phagocytosis could be the gatekeeper<br />

between the necessary phase of initial inflammation and the<br />

following phase of tissue remodelling.<br />

Disclosures:<br />

The following authors have nothing to disclose: Lara Campana, Antonella Pellicoro,<br />

Rebecca Aucott, Stephen Greenhalgh, Katherine L. Hull, Stuart J. Forbes,<br />

John P. Iredale<br />

63<br />

Dynamic shift of microbiota and its relationship with<br />

hepatic gene expression during liver regeneration<br />

Hui-Xin Liu 1 , Clarissa S. Rocha 2 , Satya Dandekar 2 , Yu-Jui Yvonne<br />

Wan 1 ; 1 Medical Pathology and Laboratory Medicine, UC DAVIS,<br />

Sacramento, CA; 2 Medical Microbiology and Immunology, UC<br />

DAVIS, Davis, CA<br />

ABSTRACT: The pathways that regulate hepatocyte proliferation<br />

and liver regeneration have been extensively studied within<br />

the liver. However, the signaling contribution specifically from<br />

the gut microbiota involved in liver regeneration is poorly<br />

understood. We have temporally profiled the microbiota, bile<br />

acids (BA), and hepatic transcriptomes in regenerating livers<br />

obtained from mice 0, 1 hour, 1, 2, 3, 5, 7, and 9 days post<br />

2/3 partial hepatectomy (Phx) and established their interactive<br />

relationships. Our data showed that liver resection led<br />

to rapid changes in the gut microbiota that was reflected in<br />

increased abundance of Bacteroidetes including S24-7 and<br />

Rikenellaceae and decreased abundance of Firmicutes including<br />

Clostridiales, Lachnospiraceae, and Ruminococcaceae.<br />

These changes persisted throughout the course of liver regeneration.<br />

Short Time-series Expression Miner analysis of RNA-Sequencing<br />

data of 6,125 genes, which had greater than 2 folds<br />

changes at their mRNA levels compared day 0 controls, generated<br />

six unique patterns: peaked at day 1 (691), 2 (818),<br />

and 3 (275), dipped at day 2 (175), continuously increasing<br />

(340) or decreasing (558) during the course of regeneration.<br />

Correlation analyses revealed close associations between the<br />

expression of genes involved in BA homeostasis, metabolism,<br />

and immune function with the altered abundance of Ruminococcacea,<br />

Lachnospiraceae, and S24-7. Quantification of BA<br />

demonstrated a rapid increase of free and conjugated BAs 1<br />

hour after liver resection due to a sudden loss of liver mass. The<br />

hepatic conjugated BA overload was correlated with increased<br />

abundance of Bacteroidetes, which mediate BA conjugation<br />

by expressing 7α-dehydroxylase. The possibility of shifts in the<br />

microbiota profile as a consequence of changes in BA composition<br />

following liver resection remains to be determined. Conclusion:<br />

This is the first report to show that microbiota profile is<br />

altered during liver regeneration. Moreover, there appears to<br />

be a close relationship between microbiota composition and<br />

hepatic gene expression pattern. The data suggested a significant<br />

role of microbiota in contributing to the liver regeneration.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hui-Xin Liu, Clarissa S. Rocha,<br />

Satya Dandekar, Yu-Jui Yvonne Wan<br />

64<br />

Annexin A6 is necessary for liver regeneration and glucose<br />

homeostasis in mice<br />

Carles Rentero Alfonso 1,2 , Anna Alvarez-Guaita 1 , Stephen E.<br />

Moss 3 , Thomas Grewal 4 , Carlos Enrich 1,2 ; 1 Biologia Cel.lular,<br />

Immunologia i Neurociences, Universitat de Barcelona, Barcelona,<br />

Spain; 2 Centre de Recerca Biomèdica CELLEX, Institut d’Investigacions<br />

Biomèdiques August Pi i Sunyer (IDIBAPS), Barcelona, Spain;<br />

3 Department of Cell Biology, UCL Institute of Ophthalmology, London,<br />

United Kingdom; 4 Faculty of Pharmacy, University of Sydney,<br />

Sydney, NSW, Australia<br />

Annexin A6 (AnxA6) belongs to a conserved family of Ca 2+ -<br />

and phospholipid-binding proteins which interact with membranes<br />

upon increase of intracellular Ca 2+ concentration. We<br />

have previously shown the relevance of AnxA6 in the regulation<br />

of endocytic and exocytic pathways, intracellular cholesterol<br />

homeostasis, in the EGFR/PKCα/Ras/MAPK signaling<br />

pathway and in membrane remodeling. However, although<br />

AnxA6 is one of the major proteins in the liver (0.25% total<br />

protein), its function in the physiology of this organ remains<br />

unknown. After partial hepatectomy (PHx), an acute proliferative<br />

and metabolic stress, quiescent hepatocytes are triggered<br />

to progress through the cell cycle, showing a synchronous onset<br />

of DNA synthesis with a cellular response that involves cell<br />

activation and tissue remodeling. During this period of liver<br />

regeneration, the liver has to retain the major physiological<br />

tasks such as the synthesis and secretion of plasma proteins,<br />

lipid homeostasis and its metabolic function to ensure viability<br />

of the organism. The aim of the present study was to investigate<br />

the in vivo function of AnxA6 during liver regeneration using<br />

an AnxA6 knock-out (AnxA6 -/- ) mouse model. To this objective,<br />

2/3 PHx was performed in wild-type C57BL/6 control and<br />

AnxA6 -/- mice. Histological, electron microscopy, biochemical<br />

and metabolic <strong>studies</strong> after PHx were performed. Primary<br />

hepatocytes were also isolated and cultured to confirm the<br />

liver-specificity. After PHx, AnxA6 -/- mice exhibit a dramatic<br />

reduction of survival rate (87.5% at 72h post-PHx). However,<br />

the exogenous administration of glucose before and during<br />

PHx restored AnxA6 -/- survival rate after PHx, suggesting a link<br />

between AnxA6 and hepatic energetic metabolism. A comprehensive<br />

analysis of glucose metabolism experiments pointed to<br />

an impairment of liver gluconeogenesis in AnxA6 -/- mice, also<br />

observed in starving mice. A biochemical approach allowed us<br />

to elucidate a role for AnxA6 in the intracellular trafficking of<br />

SLC38A2, the major liver L-Alanine transporter, which is essential<br />

for gluconeogenic hepatic substrate incorporation during<br />

liver regeneration and mice starvation. We conclude that<br />

AnxA6 is a new regulator of hepatic gluconeogenesis essential<br />

for the trafficking of the alanine transporter SLC38A2 into<br />

the hepatocyte sinusoidal plasma membrane and subsequent<br />

alanine uptake, the major gluconeogenic substrate during both<br />

liver regeneration and fasting.<br />

Disclosures:<br />

The following authors have nothing to disclose: Carles Rentero Alfonso, Anna<br />

Alvarez-Guaita, Stephen E. Moss, Thomas Grewal, Carlos Enrich


240A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

65<br />

Akt-mediated FoxO1 inhibition is required for liver<br />

regeneration after partial hepatectomy in mice<br />

Montserrat Pauta 1,3 , Noemi Rotllan 4,5 , Ana Fernandez-Hernando 6 ,<br />

Cedric Langhi 7 , Jordi Ribera 1,2 , Loreto Boix 8,2 , Jordi Bruix 8,2 ,<br />

Wladimiro Jiménez 1,9 , Yajaira Suarez 4,5 , David A. Ford 7 , Angel<br />

Baldan 7 , Morris J. Birnbaum 10 , Manuel Morales-Ruiz 1,2 , Carlos<br />

Fernandez-Hernando 4,5 ; 1 Biochemistry and Molecular Genetics,<br />

Hospital Clinic, Barcelona, Spain; 2 IDIBAPS, CIBERehd, Barcelona,<br />

Spain; 3 IDIBAPS, Barcelona, Spain; 4 Vascular Biology and Therapeutics<br />

Program, Yale University School of Medicine, New Haven,<br />

CT; 5 Integrative Cell Signaling and Neurobiology of Metabolism<br />

Program, Section of Comparative Medicine and Department of<br />

Pathology, Yale University School of Medicine, New Haven, CT;<br />

6 Departments of Medicine and Cell Biology, Leon H. Charney<br />

Division of Cardiology, New York University School of Medicine,<br />

New York, NY; 7 Edward A. Doisy Department of Biochemistry &<br />

Molecular Biology, and Center for Cardiovascular Research, Saint<br />

Louis University, Saint Louis, MO; 8 Barcelona Clinic Liver Cancer<br />

(BCLC) Group, Liver Unit, Hospital Clinic of Barcelona, University<br />

of Barcelona, Barcelona, Spain; 9 Department of Physiological<br />

Sciences I, IDIBAPS, CIBERehd, University of Barcelona School of<br />

Medicine, Barcelona, Spain; 10 Institute for Diabetes, Obesity, and<br />

Metabolism, Perelman School of Medicine, University of Pennsylvania,<br />

Philadelphia, PA<br />

Introduction and aim: Understanding the hepatic regenerative<br />

process has clinical interest since the effectiveness of many<br />

treatments for chronic liver diseases is conditioned by an efficient<br />

liver regeneration. Experimental evidence points to the<br />

need of a temporal coordination between cytokines, growth<br />

factors and metabolic signaling pathways to enable successful<br />

liver regeneration. One intracellular mediator that acts as a<br />

signal integration node for these processes is the serine-threonine<br />

kinase Akt, being Akt1 and Akt2 the most abundant<br />

isoforms expressed in the liver. Therefore, the aim of this study<br />

was to demonstrate whether Akt, globally or through any of<br />

its isoforms, has a significant role in liver regeneration. Methods:<br />

Two-thirds partial hepatectomy (PH) was performed in<br />

eight-old-week male Akt1 -/- , Akt2 -/- , Akt1 loxP/loxP , Akt2 loxP/loxP ,<br />

FoxO1 loxP/loxP and wild-type (WT) mice. In addition, we generated<br />

the Akt1 loxP/loxP ;Akt2 loxP/loxP ;FoxO1 loxP/loxP mice (TLKO)<br />

by crossing the Akt1 loxP/loxP ;Akt2 loxP/loxP mice (DLKO) with the<br />

FoxO1 loxP/loxP mice. The liver specific DLKO and TLKO deficient<br />

mice were generated by treating the floxed mice with adeno-associated<br />

virus encoding for the Cre-recombinase driven<br />

by the Alpha-Fetoprotein promoter. Mice from each group<br />

were sacrificed at various times post-hepatectomy (0 days, 2d,<br />

4d and 7d; n = 5-10) for evaluating biochemical and liver<br />

function parameters. Results: The hepatic deficiency of single<br />

Akt1 or Akt2 does not influence liver regeneration after PH.<br />

However, DLKO mice show impaired liver regeneration and<br />

a significant increase mortality (compared with WT, Akt1 -/-<br />

and Akt2 -/- mice; P10-fold;<br />

p50%) by 15min,<br />

and only approached basal levels by 96h post-PH. TMX-DTG<br />

mice (with reduced Hh signaling) accumulated 50% fewer<br />

Gli2+ cells, and 2-3 fold less Gli1 and Gli2 mRNA (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 241A<br />

approaches that inhibit Hh signaling down-regulate YAP, profibrogenic<br />

YAP targets and aSMA, showing that Hh regulates<br />

YAP and suggesting that YAP may be a downstream effector of<br />

pro-EMT Hh signaling.<br />

Disclosures:<br />

The following authors have nothing to disclose: Marzena Swiderska-Syn, Wing-<br />

Kin Syn, Guanhua Xie, Mark Jewell, Richard T. Premont, Gregory A. Michelotti,<br />

Anna Mae Diehl<br />

67<br />

Region-Based Projections of Deceased Liver Donors from<br />

2015-2025: An Analysis of Proposed UNOS Liver Allocation<br />

Redistricting<br />

Neehar D. Parikh 1 , Wesley J. Marrero Colon 2 , Yongcai Xu 2 , Anna<br />

S. Lok 1 , David W. Hutton 3 , Mariel S. Lavieri 2 ; 1 Division of Gastroenterology,<br />

University of Michigan, Ann Arbor, MI; 2 Industrial and<br />

Operations Engineering, University of Michigan, Ann Arbor, MI;<br />

3 School of Public Health, University of Michigan, Ann Arbor, MI<br />

Background: To help remedy the geographic disparity in liver<br />

transplant (LT) allocation the United Network of Organ Sharing<br />

(UNOS) has recently proposed regional redistricting plans.<br />

While current geographic inequity will improve, it is unclear to<br />

what extent demographic trends in the US will impact redistricting.<br />

We aimed to determine the impact of population demographic<br />

trends on liver redistricting proposals. Method: We<br />

performed a secondary analysis of the Organ Procurement<br />

and Transplantation Network database of all adult donors from<br />

2000 to 2012 and calculated the total number of donors available<br />

and transplanted donor livers stratified by age, race, and<br />

body mass index (BMI) group per year. We used National<br />

Health and Nutrition Examination Survey and Centers for Disease<br />

Control and Prevention historical data to stratify the US<br />

population by age, sex, race, and BMI. We then used US population<br />

age and race projections provided by the US Census<br />

Bureau and the Weldon Cooper Center for Public Service and<br />

made regional projections of available donors from 2015 to<br />

2025, incorporating the proposed 8- and 4-region redistricting<br />

plans proposed by UNOS. Given the uncertainty in LT demand,<br />

we used donors/100,000 population age 18-84 (D/100K) as<br />

a measure of equity. We calculated the mean and variation<br />

between regions, by calculating the percentage difference<br />

between the highest and lowest D/100K per allocation model.<br />

Results: The overall projected D/100K will decrease from 2.72<br />

to 2.65 over the next decade. The projected percentage variation<br />

in 2015 and 2025 in each model is shown in the Table.<br />

In the current 11-region allocation system, the 2015 range in<br />

D/100K is 2.45-2.89, and is projected to be 2.42-2.83 in<br />

2025. In the proposed 8-region allocation system the 2015<br />

range is 2.49-2.89, and is projected to be 2.42-2.83 in 2025.<br />

In the 4-region allocation system, the 2015 range is 2.49-2.79<br />

and is projected to be 2.42-2.71 in 2025. Conclusions: Of<br />

the current UNOS allocation schemes, the 4-region allocation<br />

model reduces the geographic variation to the greatest extent.<br />

This is projected to be maintained as the US population ages<br />

and undergoes demographic shifts over the next decade. However,<br />

this should be balanced by consequences of redistricting<br />

such as increased cold ischemia time and costs associated with<br />

organ transport.<br />

68<br />

Liver transplantation waiting list mortality in PSC<br />

patients is low as compared to non-PSC patients and<br />

consistent across laboratory MELD and MELD exception<br />

candidates: a nationwide study in the Netherlands<br />

Annemarie C. de Vries 1 , Madelon Tieleman 1 , Bart van Hoek 2 ,<br />

Aad P. van den Berg 3 , Wojciech G. Polak 4 , Jan Ringers 5 , Robert J.<br />

Porte 6 , Cynthia Konijn 7 , Robert A. de Man 1 , Henk R. van Buuren 1 ,<br />

Bettina E. Hansen 1 , Herold J. Metselaar 1 ; 1 Gastroenterology and<br />

Hepatology, Erasmus MC University Medical Center, Rotterdam,<br />

Netherlands; 2 Gastroenterology and Hepatology, Leiden University<br />

Medical Center, Leiden, Netherlands; 3 Gastroenterology and<br />

Hepatology, University Medical Center Groningen, Groningen,<br />

Netherlands; 4 Surgery, Erasmus MC University Medical Center,<br />

Rotterdam, Netherlands; 5 Surgery, Leiden University Medical<br />

Center, Leiden, Netherlands; 6 Surgery, University Medical Center<br />

Groningen, Groningen, Netherlands; 7 Dutch Transplantation Foundation,<br />

Leiden, Netherlands<br />

Background: PSC patients with end-stage disease form a heterogeneous<br />

group due to their varying complications. This<br />

hinders laboratory MELD score prioritization on the liver transplantation<br />

(LTx) waiting list, resulting in both lab MELD (LM)<br />

and MELD exception (ME) PSC candidates. Aim: To assess LTx<br />

waiting list mortality of LM and ME PSC candidates as compared<br />

to non-PSC patients. Methods: Patients aged ≥18 years<br />

who were listed for LTx in the Netherlands after introduction of<br />

the MELD score prioritization to December 31th 2013 were<br />

included. Data were recorded until November 2014. Exclusion<br />

criteria were reLTx, HU status or combined organ transplantation.<br />

A competing risk analysis was performed. Results: During<br />

the study period 852 candidates (M 579/ F 273; median age<br />

54.0 yrs) were listed for LTx. Median lab MELD score at listing<br />

was not significantly different between PSC patients (n=146)<br />

and non-PSC patients (n=706) (13.5 vs. 13.0; p=0.51).<br />

ME points were granted in resp. 22 PSC and 227 non-PSC<br />

patients (HR 0.34; p


242A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Bart van Hoek - Advisory Committees or Review Panels: Janssen-Cilag, Bristol<br />

Meyers Squib, Gilead, Merck, Abbvie<br />

Robert A. de Man - Advisory Committees or Review Panels: Norgine; Grant/<br />

Research Support: Biotest<br />

Henk R. van Buuren - Grant/Research Support: Intercept, Zambon Nederland BV<br />

The following authors have nothing to disclose: Annemarie C. de Vries, Madelon<br />

Tieleman, Aad P. van den Berg, Wojciech G. Polak, Jan Ringers, Robert J. Porte,<br />

Cynthia Konijn, Bettina E. Hansen, Herold J. Metselaar<br />

69<br />

Liver Transplant Outcomes and Survival Benefit for<br />

Obese Patients in the United States: Are We Disadvantaging<br />

the Morbidly Obese?<br />

Barry Schlansky 1 , Willscott E. Naugler 1 , Susan L. Orloff 2 , C. Kristian<br />

Enestvedt 2 ; 1 Gastroenterology/Hepatology, Oregon Health &<br />

Science University, Portland, OR; 2 Abdominal Organ Transplantation,<br />

Oregon Health & Science University, Portland, OR<br />

Purpose. Over 85% of U.S. centers adhere to national practice<br />

guidelines that consider morbid obesity to be a contraindication<br />

to liver transplantation (LT) based on inferior post-LT<br />

survival compared to the non-obese. In the present era, LT outcomes<br />

and survival benefit for obese patients have not been<br />

well studied. Methods. We evaluated the association of body<br />

mass index (BMI) with wait list and post-LT outcomes in patients<br />

wait listed for LT from 2005 to 2014, using the United Network<br />

for Organ Sharing database. We categorized BMI by the<br />

World Health Organization classification, with 18.5-29.9 kg/<br />

m 2 defining normal weight, 30-34.9 obesity, 35-39.9 severe<br />

obesity, and ≥40 morbid obesity. We followed patients from<br />

LT to death or graft loss, from wait listing to death before LT<br />

or receipt of LT, and from wait listing to death before or after<br />

LT (intention-to-treat, ITT). We used Cox regression to evaluate<br />

post-LT and ITT mortality and LT survival benefit, and competing<br />

risk regression to evaluate wait list mortality versus receipt<br />

of LT. We also explored temporal trends of these outcomes<br />

in an expanded cohort from 2002. Results. 3.9% of 80,221<br />

waitlisted patients and 3.5% of 38,177 transplanted patients<br />

were morbidly obese. ITT survival was lower for morbidly<br />

obese compared to lower BMI patients (log-rank p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 243A<br />

sis of complications during the donor hospitalization. Baseline<br />

data (PRE) 2/08-7/10 (29 months) were compared to donors<br />

receiving the pain protocol (POST) 3/13-4/15 (26 months).<br />

An external monitor validated the findings. Each center was IRB<br />

approved. RESULTS: 97 donors received SOC (PRE) and 75<br />

donors received the pain protocol (POST). Complications due<br />

to opioids were nausea (82%), hypoxia


244A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

73<br />

Incidence and Impact of Decompensating Events in Primary<br />

Biliary Cirrhosis - Results of an International Follow<br />

Up Study of 3030 Patients.<br />

Maren H. Harms 1 , Willem J. Lammers 1 , Harry L. Janssen 2 , Christophe<br />

Corpechot 3 , Pietro Invernizzi 4 , Marlyn J. Mayo 5 , Pier Maria<br />

Battezzati 6 , Annarosa Floreani 7 , Albert Pares 8 , Frederik Nevens 9 ,<br />

Andrew Mason 10 , Kris V. Kowdley 11 , Cyriel Y. Ponsioen 12 , Tony<br />

Bruns 13 , George N. Dalekos 14 , Douglas Thorburn 15 , Gideon<br />

Hirschfield 16 , Nicholas F. LaRusso 17 , Ana Lleo 4 , Nora Cazzagon<br />

7 , Irene Franceschet 7 , Llorenc Caballeria 8 , Kalliopi Zachou 14 ,<br />

Raoul Poupon 3 , Angela C. Cheung 2 , Palak J. Trivedi 16 , Marco<br />

Carbone 4 , Keith D. Lindor 18 , Henk R. van Buuren 1 , Bettina E.<br />

Hansen 1 ; 1 Gastroenterology and Hepatology, Erasmus University<br />

Medical Center, Rotterdam, Netherlands; 2 Liver Clinic, Toronto<br />

Western & General Hospital, University Health Network, Toronto,<br />

ON, Canada; 3 Centre de Référence des Maladies Inflammatoires<br />

des VoiesBiliaires, Hôpital Saint-Antoine, Paris, France; 4 Liver Unit<br />

and Center for Autoimmune Liver Diseases, Humanitas Clinical<br />

and Research Center, Rozzano, Italy; 5 Digestive and Liver diseases,<br />

UT Southwestern Medical Center, Dallas, TX; 6 Department<br />

of Health Sciences, Università degli Studi di Milano, Milan, Italy;<br />

7 Department of Surgery, Oncology and Gastroenterology, University<br />

of Padua, Padua, Italy; 8 Liver Unit, Hospital Clínic, CIBERehd,<br />

IDIBAPS, University of Barcelona, Barcelona, Spain; 9 Dept of<br />

Hepatology, University Hospitals Leuven, KULeuven, Leuven, Belgium;<br />

10 Divison of Gastroenterology and Hepatology, University of<br />

Alberta, Edmonton, AB, Canada; 11 Liver Care Network, Swedish<br />

Medical Center, Seattle, WA; 12 Dept of Gastroenterology and<br />

Hepatology, Academic Medical Center, Amsterdam, Netherlands;<br />

13 Department of Gastroenterology and Hepatology, University of<br />

Jena, Jena, Germany; 14 Institute of Internal Medicine and Hepatology,<br />

University of Thessaly, Larissa, Greece; 15 The Sheila Sherlock<br />

Liver Centre, The Royal Free Hospital, London, United Kingdom;<br />

16 NIHR Biomedical Research Unit and Centre for Liver Research,<br />

University of Birmingham, Birmingham, United Kingdom; 17 Dept<br />

Gastroenterology and Hepatology, Mayo Clinic, Rochester, MN;<br />

18 Arizona State University, Phoenix, AZ<br />

Background Clinical events of decompensation are considered<br />

indicators of poor prognosis in primary biliary cirrhosis (PBC).<br />

Few <strong>studies</strong> have assessed the incidence of decompensating<br />

events and the related outcome of patients. Methods Long-term<br />

follow-up (FU) data of ursodeoxycholic acid (UDCA) treated<br />

patients were derived from 17 North-American and European<br />

centers. Decompensation was defined as a first event of ascites,<br />

variceal bleeding, or encephalopathy, whichever came<br />

first. Patients with events prior to baseline or in first year of FU<br />

were excluded. Risk factor analysis was performed using Cox<br />

proportional hazard models. Results Of 3030 UDCA-treated<br />

PBC patients, 92 patients (3.0%) were excluded. Median FU<br />

was 8.4 yrs (IQR 4.7-12.9). A decompensating event occurred<br />

in 275 patients: ascites:167 (61%), variceal bleeding:76<br />

(27%), encephalopathy:24 (9%), multiple:8 (3%). 1-, 3- and<br />

5- year transplantation-free survival with or without event was<br />

59% vs 99%, 35% vs 97% and 19% vs 94% respectively (time<br />

dependent HR40.6; 95%CI 29.6-55.7). Survival did not significantly<br />

differ between different types of events. Multivariable<br />

analysis showed that at time of first event, the following<br />

factors are predictive of survival: age at time of event (per<br />

10 yrs) (HR 1.39; 95%CI 1.09-1.63), calendar year of event<br />

(HR0.97; 95%CI 0.94-0.99), bilirubin>2x upper limit of normal<br />

(ULN) (HR2.98; 95%CI 1.99-4.45) and normal albumin<br />

(HR1.68; 95%CI 1.12-2.53). Survival of patients with normal<br />

albumin and bilirubin<br />

2xULN (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 245A<br />

LTx. However, prognostic models based on population-based<br />

cohorts are lacking. Various biomarkers of disease progression<br />

have been assessed for PSC, and the potential prognostic<br />

value of serum alkaline phosphatase (ALP) over time has been<br />

discussed. The aim of this study was to create a prognostic<br />

model for PSC consisting of disease characteristics and biochemical<br />

variables, and to validate this model in an external<br />

cohort. Methods 692 PSC patients were identified in a large,<br />

population-based PSC cohort from the Netherlands. Disease<br />

characteristics and biochemical variables during long term<br />

follow-up (FU) were retrieved from patient records. Clinical<br />

endpoints were: development of cholangiocarcinoma, LTx,<br />

or PSC-related death. Biochemical tests were transformed by<br />

log transformation, missing values were imputed by multiple<br />

imputation. All variables were assessed as potential predictors<br />

of survival by univariate analysis. To calculate the prognostic<br />

index (PI), a Cox proportional hazards model was developed<br />

and internally validated with bootstrap. The model was<br />

then validated in an external cohort of 259 PSC patients from<br />

UK. Results The median FU was 85 months (range 0-468<br />

months). All disease characteristics and biochemical variables<br />

were considered for the model. After variable selection by<br />

LASSO, multivariate Cox models were fitted, and parameters<br />

estimated from 20 imputation datasets were averaged. The<br />

following formula was created: PI=1.409*PSC type (0/1) 1<br />

+ 0.021*Age_PSC diagnosis - 2.420*log_AlbuminxULN 2 +<br />

2.073*abs(log_TrombocytesxULN-0.5) 2 + 0.469*log_AspartateAminotransferase(AST)xULN<br />

2 + 0.565*log_ALPxULN 2 +<br />

0.528*log_TotalBilirubinxULN 2 1: PSC type: Large duct=1;<br />

Small duct=0 2: xULN= times upper limit of normal A higher<br />

PI indicated a worse prognosis, corresponding to a shorter<br />

endpoint-free survival. The PI yielded a c-statistic of 0.715 in<br />

the development dataset, 0.705 in internal validation (adjusted<br />

for optimism with 1000 times bootstrap), and 0.683 in the<br />

external validation dataset. Conclusion A novel prognostic PSC<br />

model based on PSC type, age at PSC diagnosis, albumin,<br />

thrombocytes, AST, ALP and bilirubin which shows adequate<br />

performance in internal and external validation cohorts, allows<br />

to determine complication-free survival probability for PSC<br />

patients over time.<br />

Disclosures:<br />

Ulrich Beuers - Consulting: Intercept via University of Amsterdam, Novartis via<br />

University of Amsterdam; Grant/Research Support: Falk, Zambon; Speaking and<br />

Teaching: Falk Foundation, Gilead, Roche, Shire<br />

Cyriel Y. Ponsioen - Advisory Committees or Review Panels: Takeda; Consulting:<br />

AbbVIE; Grant/Research Support: AbbVIE, Schering Plough, Dr. Falk Pharma,<br />

Tramedico Netherlands, Takeda<br />

The following authors have nothing to disclose: Elisabeth M. de Vries, Junfeng<br />

Wang, Kate D. Williamson, Mariska M. Leeflang, Kirsten Boonstra, Roger W.<br />

Chapman, Ronald Geskus<br />

75<br />

Clinical Epidemiology of Primary Biliary Cirrhosis based<br />

on a Large US Laboratory Database: Incidence and<br />

Trends in Serum Alkaline Phosphatase<br />

W. Ray Kim 1 , Tracy J. Mayne 2 , Tonya Marmon 3 , David Shapiro 4 ,<br />

Keith D. Lindor 5 ; 1 Division of Gastroenterology and Hepatology,<br />

Stanford University School of Medicine, Stanford, CA; 2 Outcomes<br />

Research, Intercept Pharmaceuticals, New York, NY; 3 Biostatistics,<br />

Intercept Pharmaceuticals, San diego, CA; 4 Office of the Chief<br />

Medical Officer, Intercept Pharmaceuticals, San Diego, CA; 5 College<br />

of Health Solutions, Arizona State University, Tempe, AZ<br />

Backround/Aims: Large scale epidemiological data on primary<br />

biliary cirrhosis (PBC) are scarce. Using a nationwide database<br />

covering approximately 30% of US adults, we estimate<br />

PBC incidence and describe trends in serum alkaline phosphatase<br />

(ALP) activities following the diagnosis. Methods: Electronic<br />

data from a large commercial laboratory were queried<br />

to extract all AMA and ALP results between Jan 2010 and Dec<br />

2013. Following AASLD guidelines, we defined probable PBC<br />

by positive anti-mitochondrial antibodies (AMA) and elevated<br />

serum ALP. AMA positivity was defined by a titer > 1:20 or M2<br />

antibody > 25.0 U, and elevated ALP (> upper limit of normal<br />

(ULN), 120 U/L) by the peak ALP before and up to 30-days<br />

after the first positive AMA. In subjects with PBC, serum ALP<br />

was followed for 24 months after the first positive AMA (baseline)<br />

and the highest ALP was identified per 6-month interval.<br />

Results: Out of over 576,000 persons receiving AMA testing,<br />

16,492 unique individuals (2.9%) tested positive. In 5,380<br />

(33%), serum ALP was elevated at baseline, while another 727<br />

AMA-positive patients developed elevated ALP subsequent to<br />

a positive AMA result. Thus, a total of 6,107 met the criteria<br />

for probable PBC, giving rise to a crude incidence rate of<br />

5.8 per 100,000 over 4 years. In the figure, of the 5,380<br />

patients, 47% had ALP > x2ULN, including 25% with ALP ><br />

3xULN. Following the diagnosis, the proportion of elevated<br />

ULN decreased, presumably as a result of therapy. By 3-6<br />

months, 61% of patients had ALP < 1.5xULN. Subsequent<br />

trends did not show further improvement in ALP – as of 24<br />

months post diagnosis, 69% had elevated ALP, including 35%<br />

with ALP > 1.5xULN. Conclusion: This largest epidemiological<br />

survey for PBC in US to date suggests that annually more than<br />

5000 Americans meet the diagnosis of probable PBC. Two<br />

years after the diagnosis, more than a third continue to have<br />

significantly elevated ALP.<br />

Disclosures:<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

Tonya Marmon - Employment: Intercept Pharmaceuticals, Inc; Stock Shareholder:<br />

Intercept Pharmaceuticals, Inc<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: Keith D. Lindor


246A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

76<br />

Gender and IBD phenotype are independent predictors<br />

of death or transplantation and of malignancy in Primary<br />

Sclerosing Cholangitis – a multicenter retrospective<br />

study of the International PSC Study Group (IPSCSG)<br />

Tobias J. Weismuller 1 , Bettina E. Hansen 2 , Palak J. Trivedi 3 , Mohamad<br />

Imam 6 , Henrike Lenzen 5 , Cyriel Y. Ponsioen 6 , Kristian Holm 7 ,<br />

Ulrich Beuers 6 , Annika Bergquist 8 , Daniel Gotthardt 9 , Hanns-Ulrich<br />

Marschall 10 , Douglas Thorburn 11 , Rinse K. Weersma 12 , Johan<br />

Fevery 13 , Tobias Mueller 14 , Olivier Chazouillères 15 , Christoph<br />

Schramm 16 , Konstantinos Lazaridis 4 , Brian D. Juran 4 , Martti A.<br />

Färkkilä 19 , Stephen P. Pereira 18 , Sven Almer 17,8 , Cynthia Levy 20 ,<br />

Andrew Mason 21 , Christopher L. Bowlus 22 , Annarosa Floreani 23 ,<br />

Emina Halilbasic 24 , Michael Trauner 24 , Kidist K. Yimam 25 , Piotr<br />

Milkiewicz 26,27 , Dep K. Huynh 28 , Albert Pares 29 , Christine N.<br />

Manser 30 , George N. Dalekos 31 , Bertus Eksteen 32 , Gabi I. Kirchner<br />

38 , Christoph Sarrazin 37 , Vincent Zimmer 36 , Luca Fabris 35 ,<br />

Pietro Invernizzi 34 , Felix Braun 33 , Marco Marzioni 39 , Christoph<br />

P. Berg 40 , Ansgar W. Lohse 16 , Gideon Hirschfield 3 , Christian<br />

P. Strassburg 1 , Michael P. Manns 5 , Keith D. Lindor 4 , Tom H.<br />

Karlsen 7 , Kirsten M. Boberg 7 ; 1 Department of Internal Medicine I,<br />

University of Bonn, Bonn, Germany; 2 Department of Gastroenterology<br />

and Hepatology, Erasmus University Medical Center, Rotterdam,<br />

Netherlands; 3 National Institute for Health Research (NIHR)<br />

Birmingham Liver Biomedical Research Unit (BRU) and Centre for<br />

Liver Research, University of Birmingham, Birmingham, United<br />

Kingdom; 4 Department of Gastroenterology and Hepatology,<br />

Mayo Clinic, Rochester, MN; 5 Department of Gastroenterology,<br />

Hepatology and Endocrinology, Hannover Medical School, Hannover,<br />

Germany; 6 Department of Gastroenterology & Hepatology,<br />

Academic Medical Center, Amsterdam, Netherlands; 7 Norwegian<br />

PSC Research Center and Section for Gastroenterology, Department<br />

of Transplantation Medicine, Division of Cancer Medicine,<br />

Surgery and Transplantation, Oslo University Hospital, Rikshospitalet,<br />

Oslo, Norway; 8 Center for Digestive Diseases, Division of<br />

Hepatology, Karolinska University Hospital and Karolinska Institutet,<br />

Stockholm, Sweden; 9 Dept. of Gastroenterology, Infectious<br />

Diseases and Intoxications, University Hospital Heidelberg, Heidelberg,<br />

Germany; 10 Department of Molecular and Clinical Medicine,<br />

Sahlgrenska Academy, University of Gothenburg, Gothenburg,<br />

Sweden; 11 The Sheila Sherlock Liver Centre and UCL Institute for<br />

Liver and Digestive Health, Royal Free Hospital, London, United<br />

Kingdom; 12 Department of Gastroenterology and Hepatology, University<br />

of Groningen and University Medical Center Groningen,<br />

Groningen, Netherlands; 13 Department of Hepatology, University<br />

Hospital Gasthuisberg, Leuven, Belgium; 14 Department of Internal<br />

Medicine, Hepatology and Gastroenterology, Charité Universitätsmedizin,<br />

Berlin, Germany; 15 Service d’Hépatologie, Hôpital Saint<br />

Antoine, Assistance Publique-Hôpitaux de Paris,Faculté de Médecine<br />

Pierre et Marie Curie, Paris, France; 16 1st Department of Medicine,<br />

University Medical Center Hamburg Eppendorf, Hamburg,<br />

Germany; 17 Gastroenterology & Hepatology, Linköping University,<br />

Linköping, Sweden; 18 Institute for Liver and Digestive Health, UCL,<br />

London, United Kingdom; 19 Division of Gastroenterology, Department<br />

of Medicine, Helsinki University Central Hospital, Helsinki,<br />

Finland; 20 University of Miami, Miami, FL; 21 Division of Gastroenterology<br />

and Hepatology, University of Alberta, Edmonton, AB,<br />

Canada; 22 Division of Gastroenterology and Hepatology, University<br />

of California Davis, Davis, CA; 23 Department of Surgery,<br />

Oncology and Gastroenterology, University of Padova, Padova,<br />

Italy; 24 Division of Gastroenterology and Hepatology, Department<br />

of Internal Medicine III, Medical University of Vienna, Vienna,<br />

Austria; 25 Department of Hepatology and Liver Transplantation,<br />

California Pacific Medical Center, San Francisco, CA; 26 Department<br />

of Clinical and Molecular Biochemistry, Pomeranian Medical<br />

University, Szczecin, Poland; 27 Liver and Internal Medicine Unit,<br />

Department of General, Transplant and Liver Surgery, Medical<br />

University of Warsaw, Warsaw, Poland; 28 Department of Gastroenterology<br />

and Hepatology, Royal Adelaide Hospital, Adelaide,<br />

SA, Australia; 29 Liver Unit, Hospital Clinic, IDIBAPS, CIBERehd,<br />

University of Barcelona, Barcelona, Spain; 30 Division for Gastroenterology<br />

and Hepatology, University Hospital Zurich (USZ), Zurich,<br />

Switzerland; 31 Department of Medicine and Research Laboratory<br />

of Internal Medicine, School of Medicine, University of Thessaly,<br />

Larissa, Greece; 32 University of Calgary, Snyder Institute for<br />

Chronic Diseases, Alberta, AB, Canada; 33 Campus Kiel, UKSH,<br />

Kiel, Germany; 34 Humanitas Clinical and Research Center, Center<br />

for Autoimmune Liver Diseases, Rozzano (MI), Italy; 35 Department<br />

of Molecular Medicine, University of Padua School of Medicine,<br />

Padua, Italy; 36 Department of Medicine II, Saarland University<br />

Medical Center, Homburg, Germany; 37 Department of Internal<br />

Medicine 1, Johann Wolfgang Goethe-University Hospital, Frankfurt,<br />

Germany; 38 Department of Internal Medicine 1, University<br />

Hospital of Regensburg, Regensburg, Germany; 39 Department of<br />

Gastroenterology, Università Politecnica delle Marche, Ancona,<br />

Italy; 40 Department of Gastroenterology, Hepatology, and Infectiology,<br />

Medical Clinic, University of Tübingen, Tübingen, Germany<br />

Primary sclerosing cholangitis (PSC) is a male predominant<br />

disease bearing a strong association with inflammatory bowel<br />

disease (IBD). The aims of this study were to address impact<br />

of gender and IBD phenotype on liver-related outcomes across<br />

a large, internationally representative multi-centre registry.<br />

All patients diagnosed with PSC between 1/1/1980 and<br />

12/31/2010 in 37 institutions from 17 countries were included<br />

in the study. Data on clinical presentation, survival, liver transplantation<br />

(LT), IBD, colorectal neoplasia and hepatobiliary<br />

malignancy (HBM) were collected. Cox regression analyses,<br />

stratified by region and adjusted for age, year of diagnosis,<br />

diagnose (PSC/small duct PSC/ with AIH overlap) with IBD<br />

as a time dependent covariate were used to study the effect of<br />

gender on the risk of HBM, colorectal neoplasia and the risk<br />

of LT and death. 7119 patients were included (65.5% male,<br />

mean age at diagnosis 37.4 years in males and 40.4 years<br />

in females (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 247A<br />

Disclosures:<br />

Palak J. Trivedi - Grant/Research Support: Wellcome Trust<br />

Cyriel Y. Ponsioen - Advisory Committees or Review Panels: Takeda; Consulting:<br />

AbbVIE; Grant/Research Support: AbbVIE, Schering Plough, Dr. Falk Pharma,<br />

Tramedico Netherlands, Takeda<br />

Ulrich Beuers - Consulting: Intercept via University of Amsterdam, Novartis via<br />

University of Amsterdam; Grant/Research Support: Falk, Zambon; Speaking and<br />

Teaching: Falk Foundation, Gilead, Roche, Shire<br />

Hanns-Ulrich Marschall - Consulting: Albireo<br />

Olivier Chazouillères - Consulting: APTALIS, MAYOLY-SPINDLER<br />

Andrew Mason - Advisory Committees or Review Panels: AbbVie, Novartis,<br />

Intercept; Grant/Research Support: Abbvie, Gilead, Astellas<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

Gideon Hirschfield - Advisory Committees or Review Panels: Intercept Pharma;<br />

Consulting: Dignity Sciences, GSK, NGM Bio, Lumena, J & J; Grant/Research<br />

Support: BioTie; Speaking and Teaching: Falk Pharma<br />

Christian P. Strassburg - Advisory Committees or Review Panels: Novartis, Roche;<br />

Speaking and Teaching: Novartis, Merz, MSD, Falk Pharma, BMS, Abbvie<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

The following authors have nothing to disclose: Tobias J. Weismuller, Bettina<br />

E. Hansen, Mohamad Imam, Henrike Lenzen, Kristian Holm, Annika Bergquist,<br />

Daniel Gotthardt, Douglas Thorburn, Rinse K. Weersma, Johan Fevery,<br />

Tobias Mueller, Christoph Schramm, Konstantinos Lazaridis, Brian D. Juran,<br />

Martti A. Färkkilä, Stephen P. Pereira, Sven Almer, Cynthia Levy, Annarosa<br />

Floreani, Emina Halilbasic, Kidist K. Yimam, Piotr Milkiewicz, Dep K. Huynh,<br />

Albert Pares, Christine N. Manser, George N. Dalekos, Bertus Eksteen, Gabi<br />

I. Kirchner, Vincent Zimmer, Luca Fabris, Pietro Invernizzi, Felix Braun, Marco<br />

Marzioni, Christoph P. Berg, Ansgar W. Lohse, Keith D. Lindor, Tom H. Karlsen,<br />

Kirsten M. Boberg<br />

77<br />

Ig Repertoire of Autoantigen-Specific B cells in Primary<br />

Biliary Cirrhosis<br />

Weici Zhang 1 , Patrick S. Leung 1 , Ying Sun 1,2 , Thomas P. Kenny 1 ,<br />

Guo-Xiang Yang 1 , Xiao-Song He 1 , Christopher L. Bowlus 3 , Ross<br />

L. Coppel 4 , Ignacio Sanz 6 , Aftab A. Ansari 5 , M. Eric Gershwin 1 ;<br />

1 Internal Medicine, University of California, Davis, CA; 2 Hepatology,<br />

Center for Non-Infectious Liver Diseases, Beijing 302 Hospital,<br />

Beijing, China; 3 Division of Gastroenterology and Hepatology,<br />

University of California at Davis, Sacramento, CA; 4 Department<br />

of Microbiology, Monash University, Melbourne, VIC, Australia;<br />

5 Department of Pathology, Emory University School of Medicine,<br />

Atlanta, GA; 6 Department of Medicine, Division of Rheumatology<br />

and Lowance Center for Human Immunology, Emory University,<br />

Atlanta, GA<br />

Background: Autoantibodies to the E2 subunit of pyruvate<br />

dehydrogenase (PDC-E2) are the serological hallmark of PBC.<br />

Interestingly, a subpopulation of anti-PDC-E2 autoantibodies<br />

are cross-reactive to chemical xenobiotics including 2-octynoic<br />

acid (2OA) and 6, 8-bis (acetylthio) octanoic acid (SAc), both<br />

putative etiological agents of PBC. Recent <strong>studies</strong> suggest that<br />

anti-PDC-E2 is derived from de novo activated autoantigen-specific<br />

B cells, or from B cells previously primed by xenobiotics<br />

when self-tolerance was originally compromised. Examination<br />

of antigen specific plasmablasts from PBC demonstrate an<br />

extraordinary frequency of circulating plasmablasts specific<br />

for PDC-E2 without significant detection of reactivity to xenobiotic<br />

antigens; this is in striking contrast to comparable sera<br />

analysis. We propose that circulating PDC-E2 plasmablasts<br />

result from positive selection to xenobiotics, i.e. the “original<br />

sin” was the creation of a neoantigen between a chemical<br />

xenobiotic and a self peptide. Our aim herein was to identify<br />

Ig sequences from plasmablasts directed to PDC-E2. Methods:<br />

1,680 single plasmablast cells (CD3 – CD19 + CD20 lo/– CD27 hi C-<br />

D38 hi ) were obtained from PBC with high PDC-E2-specific<br />

plasmablasts (>30%). Thereafter, Ig heavy chain (IgH) and Igk<br />

or Igl light chain (IgL) gene transcripts were amplified using a<br />

IgG, IgA and IgM H and L chain gene specific primer cocktail<br />

and sequenced. A total of 420 functional paired IgH and<br />

Igk or Igl gene sequences were analyzed. CDR3 length and<br />

frequency of positively and negatively charged residues in<br />

various V gene libraries were analyzed. Differential analysis<br />

was performed comparing IgM versus IgG or IgA gene libraries.<br />

Results: Our data revealed a decreased diversity of IgH<br />

and IgL usage with preferential usage of IGHV3-23, IGHV3-<br />

30, IGHV3-7, IGHV4-39, IGHV3-15, IGHV1-18, IGHV3-74<br />

and IGKV3-20, IGKV2-28. Moreover, the usage of IGHV3-<br />

23 and IGHV3-30 increased by two-fold in PBC compared to<br />

controls. Thirty-two clonal V region sequences in the repertoire<br />

were identified and shared among PBC patients. Interestingly,<br />

the unique rearrangements accounted for 0.84% of the total<br />

sequenced IgH rearrangements and the majority of sequences<br />

are highly mutated with an average of 17 replacement mutations<br />

per sequence. Conclusion: We suggest that the original<br />

breach of tolerance was secondary to molecular mimicry in<br />

a genetically susceptible host to a neoantigen that includes a<br />

chemical xenobiotic of PDC-E2; subsequently, through determinant<br />

spreading and autoantibody maturation, the predominant<br />

hallmark of PBC arises, autoantibodies directed at the inner<br />

lipoyl domain of PDC-E2.<br />

Disclosures:<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

Ignacio Sanz - Advisory Committees or Review Panels: Pfizer; Consulting: Genentech,<br />

Sanofi; Grant/Research Support: MedImmune<br />

The following authors have nothing to disclose: Weici Zhang, Patrick S. Leung,<br />

Ying Sun, Thomas P. Kenny, Guo-Xiang Yang, Xiao-Song He, Ross L. Coppel,<br />

Aftab A. Ansari, M. Eric Gershwin


248A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

78<br />

Intrahepatic Cholestasis of Pregnancy (ICP) in U.S. Latinas<br />

and Chileans: Ancestry Analysis, Admixture Mapping,<br />

and Biochemical Features<br />

Laura Bull 1,2 , Donglei Hu 2,3 , Sohela Shah 1,2 , Luisa Temple 1,2 ,<br />

Karla Silva 4,5 , Scott Huntsman 2,3 , Jennifer Melgar 1,2 , Mary<br />

Geiser 1,2 , Ukina Sanford 1,2 , Juan Ortiz 5,6 , Richard Lee 7 , Juan P.<br />

Kusanovic 4,5 , Elad Ziv 3,2 , Juan Vargas 8,9 ; 1 Medicine, San Francisco<br />

General Hospital, UCSF, San Francisco, CA; 2 Institute for<br />

Human Genetics, UCSF, San Francisco, CA; 3 Medicine, UCSF,<br />

San Francisco, CA; 4 Center for Research and Innovation in Maternal-Fetal<br />

Medicine, Hospital Dr. Sótero del Río, Puente Alto, Chile;<br />

5 Obstetrics and Gynecology, Hospital Dr. Sótero del Río, Puente<br />

Alto, Chile; 6 Obstetrics and Gynecology, Clínica Santa María,<br />

Santiago, Chile; 7 Obstetrics and Gynecology, Los Angeles County<br />

+ University of Southern California Medical Center, Los Angeles,<br />

CA; 8 Obstetrics, Gynecology and Reproductive Sciences, UCSF,<br />

San Francisco, CA; 9 Radiology, UCSF, San Francisco, CA<br />

Purpose: In the Americas, women with Indigenous American<br />

ancestry are at increased risk of intrahepatic cholestasis of<br />

pregnancy (ICP). We hypothesized that ancestry-related<br />

genetic factors contribute to this increased risk, and that the<br />

statistically powerful genetic admixture mapping approach<br />

could allow mapping of an ICP susceptibility locus. We considered<br />

that diagnosis, features, and management of ICP might<br />

differ between U.S. and Chilean clinical sites. Methods: We<br />

collected patient data and performed biochemical assays and<br />

SNP genotyping on samples from U.S. Latinas and Chilean<br />

women, with and without ICP. The study sample included 198<br />

women with ICP (90 from the U.S. and 108 from Chile) and<br />

174 pregnant control women (69 from the U.S. and 105 from<br />

Chile). We compared overall genetic ancestry between cases<br />

and controls, and performed genome-wide admixture mapping,<br />

taking country of enrollment into account, to screen for<br />

ICP susceptibility loci. We characterized features of ICP at the<br />

U.S. and Chilean clinical sites. Results: Cases had a greater<br />

proportion of Indigenous American ancestry than did controls<br />

(p=0.034); when U.S. and Chilean study samples were analyzed<br />

separately, this difference was apparent in the U.S., but<br />

not the Chilean, study sample, potentially due to differences in<br />

population history. Admixture mapping allowed identification<br />

of one locus for which Native American ancestry was associated<br />

with increased risk of ICP at a genome-wide level of<br />

significance (P = 3.1 x 10 -5 , P corrected<br />

= 0.035). This locus, on<br />

chromosome 2, has an odds ratio of 4.48 (95% CI: 2.21-9.06)<br />

for 2 versus zero Indigenous American chromosomes; the 10<br />

Mb 95% confidence interval does not contain any previously<br />

identified hereditary ‘cholestasis genes.’ ICP appears biochemically<br />

more severe in the U.S., than Chilean, study sample.<br />

Differences in management include 1) ursodeoxycholate was<br />

prescribed to a greater proportion of U.S. than Chilean patients<br />

(48% vs 1%); 2) most U.S. cases were delivered by 37 weeks<br />

gestation (or at time of diagnosis if later), while induction of<br />

labor is not usually performed until 40 weeks at the Chilean<br />

site; and 3) cesarean birth is more common in Chilean cases<br />

than controls, while rates are similar between U.S. cases and<br />

controls. Conclusions: Our results indicate that genetic factors<br />

contribute to risk of developing ICP in the Americas, and support<br />

the utility of clinical and genetic <strong>studies</strong> of ethnically mixed<br />

populations for increasing our understanding of ICP. Greater<br />

characterization of diagnosis, management and outcomes of<br />

ICP in different countries may improve our understanding of<br />

ICP, and its potential subtypes.<br />

Disclosures:<br />

Sohela Shah - Employment: Qiagen Bioinformatics<br />

The following authors have nothing to disclose: Laura Bull, Donglei Hu, Luisa Temple,<br />

Karla Silva, Scott Huntsman, Jennifer Melgar, Mary Geiser, Ukina Sanford,<br />

Juan Ortiz, Richard Lee, Juan P. Kusanovic, Elad Ziv, Juan Vargas<br />

79<br />

The effects of apoptosis inhibition on pulmonary alveolar<br />

type 2 epithelial cell alterations in experimental<br />

hepatopulmonary syndrome after common bile duct<br />

ligation<br />

Wenli Yang, Bingqian Hu, Michael B. Fallon, Junlan Zhang; The<br />

university of Texas Health Science Center, Houston, TX<br />

Introduction: Abnormal oxygenation due to lung ventilation-perfusion<br />

mismatch drives the development of hepatopulmonary<br />

syndrome (HPS). In addition to alterations in microvascular<br />

perfusion, dysfunction in type II alveolar epithelial cells (AT2),<br />

which play a critical role in maintaining alveolar ventilation<br />

through surfactant protein (SPs) production, develops in experimental<br />

HPS. Specifically, AT2 cell apoptosis and decreased<br />

SP production associated with increased circulating bile acid<br />

levels and bile acid nuclear receptor (FXRα) activation, reduce<br />

alveolar airspace and impair ventilation. However, the relative<br />

effects of FXRα activation on AT2 cell apoptosis and SP<br />

expression and whether apoptosis inhibition influences AT2<br />

cells and the development of HPS are unknown. Aim: To evaluate<br />

the effects of apoptosis inhibition on FXRα activation in<br />

AT2 cells and in experimental HPS. Methods: A murine AT2<br />

cell line (MLE-12) was treated with a FXRα agonist (GW4064,<br />

1-10mM) in the presence or absence of a pan caspase inhibitor<br />

(Z-VAD-FMK). SD rats underwent common bile duct ligation<br />

(CBDL) or sham operation and were treated with Z-VAD-FMK<br />

(1.0mg/kg/day, IP, for 3 weeks). Lung and cell apoptosis was<br />

determined by TUNEL staining or cleaved caspase-3 levels.<br />

Co-staining of TUNEL and SP-C was performed to assess lung<br />

AT2 cell apoptosis. SP levels were assessed by westerns or qRT-<br />

PCR. Lung morphology (alveolar mean chord length, Lm) and<br />

gas exchange (PO 2<br />

) were measured to evaluate HPS. Results:<br />

GW4064 treatment in MLE-12 cells caused a dose dependent<br />

increase in cleaved caspase-3 protein expression and reduction<br />

of SP-C and SP-B mRNAs. Z-VAD-FMK pretreatment attenuated<br />

GW4064 induced cleaved caspase-3 protein increases<br />

(76.0±6.1% reduction, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 249A<br />

80<br />

Statins activate the canonical hedgehog-signaling and<br />

aggravate non-cirrhotic portal hypertension, but inhibit<br />

the non-canonical hedgehog signaling and cirrhotic portal<br />

hypertension.<br />

Frank E. Uschner 1 , Ganesh Ranabhat 1 , Michaela Granzow 1 , Steve<br />

S. Choi 2 , Sabine Klein 1 , Robert Schierwagen 1 , Esther Raskopf 1 ,<br />

Sebastian Gautsch 1 , Peter F. Van der Ven 3 , Christian P. Strassburg<br />

1 , Dieter O. Fürst 3 , Kanishka Hiththetiya 4 , Tilman Sauerbruch 1 ,<br />

Anna Mae Diehl 2 , Jonel Trebicka 1 ; 1 Medical Clinic I, University of<br />

Bonn, Bonn, Germany; 2 Department of Medicine, Duke University,<br />

Durham, NC; 3 Department of Molecular Cell Biology, University of<br />

Bonn, Bonn, Germany; 4 Institute of Pathology, University of Bonn,<br />

Bonn, Germany<br />

Background: Portal hypertension is a source of complications in<br />

patients with liver cirrhosis, as well as in patients with portal vein<br />

obstruction. In both conditions, increased outflow resistance<br />

and hyperperfusion of splanchnic vessels trigger angiogenesis.<br />

Extrahepatic angiogenesis exacerbates hyperperfusion of<br />

the splanchnic vessels and aggravates portal hypertension. In<br />

cirrhosis, intrahepatic angiogenesis is involved in progression<br />

of fibrosis since activated hepatic stellate cells (HSC) produce<br />

collagen and serve as pericytes. HSC activation and activity<br />

are promoted by Hedgehog (Hh) pathway, which further regulates<br />

vascular integrity and angiogenesis in different tissues.<br />

Statins not only blunt liver fibrosis and decreases portal pressure,<br />

but also interacts with the Hh. This study investigated the<br />

differences of angiogenesis in cirrhotic and non-cirrhotic portal<br />

hypertension with special emphasis on the canonical (Shh/Gli)<br />

and non-canonical (Shh/RhoA) pathway. Methods: Cirrhotic<br />

(BCL; CCl 4<br />

) and non-cirrhotic (partial portal vein ligation/<br />

PPVL) rats received either atorvastatin (15mg/kg, 7d) or control<br />

chow. Invasive hemodynamics were assessed by coloured<br />

microspheres. Implantation of matrigel (s.c. and i.p.) assessed<br />

angiogenesis in vivo. In vitro angiogenesis was analyzed in<br />

primary rat HSC and LX-2 cells, after atorvastatin incubation<br />

and transfection with RhoA plasmids. Expression was analyzed<br />

using RT-PCR and Western blot. Results: Atorvastatin decreased<br />

portal pressure, shunt flow and angiogenesis in cirrhotic rats,<br />

whereas atorvastatin increased these parameters in PPVL rats.<br />

Hh was upregulated in experimental and human liver cirrhosis,<br />

but was blunted by atorvastatin. The transfection of LX-2 cells<br />

with a RhoA plasmid increased Hh protein levels and inhibition<br />

of the Hh-pathway reduced profibrotic markers, which clearly<br />

indicate a non-canonical activation of Hh-pathway in HSC.<br />

Moreover, atorvastatin blocked this non-canonical Hh-pathway<br />

RhoA-dependently in activated HSCs. Interestingly, hepatic and<br />

extrahepatic Hh-pathway was enhanced in PPVL rats, which<br />

resulted in increased angiogenesis. Discussion: In summary,<br />

statins caused contrary effects in cirrhotic and non-cirrhotic<br />

portal hypertension. Atorvastatin inhibited the non-canonical<br />

Hh-pathway and angiogenesis in cirrhosis. In portal vein<br />

obstruction, statins enhanced the canonical Hh-pathway and<br />

aggravated PHT and angiogenesis. In conclusion, the broad<br />

use of statins with its beneficial effects in liver cirrhosis still<br />

requires caution in therapy of portal hypertension without cirrhosis.<br />

Disclosures:<br />

Christian P. Strassburg - Advisory Committees or Review Panels: Novartis, Roche;<br />

Speaking and Teaching: Novartis, Merz, MSD, Falk Pharma, BMS, Abbvie<br />

The following authors have nothing to disclose: Frank E. Uschner, Ganesh<br />

Ranabhat, Michaela Granzow, Steve S. Choi, Sabine Klein, Robert Schierwagen,<br />

Esther Raskopf, Sebastian Gautsch, Peter F. Van der Ven, Dieter O. Fürst,<br />

Kanishka Hiththetiya, Tilman Sauerbruch, Anna Mae Diehl, Jonel Trebicka<br />

81<br />

Hepatocyte Tissue Factor Contributes to the Hypercoagulable<br />

State in a Mouse Model of Chronic Liver Injury<br />

Pierre-Emmanuel Rautou 1,2 , Kohei Tatsumi 3 , Silvio Antoniak 3 ,<br />

Albert P. Owens III 3 , Erica Sparkenbaugh 3 , Anna K. Kopec 4 , Rafal<br />

Pawlinski 3 , James P. Luyendyk 4 , Nigel Mackman 3 ; 1 Hepatology,<br />

Hôpital Beaujon - APHP, Clichy, France; 2 Inserm U970; Paris cardiovascular<br />

Research center @ HEGP, Paris, France; 3 Department<br />

of Medicine, Division of Hematology and Oncology, McAllister<br />

Heart Institute, University of North Carolina at Chapel Hill, Chapel<br />

Hill, NC; 4 Department of Pathobiology and Diagnostic Investigation,<br />

Michigan State University, East Lansing, MI<br />

Background & Aims: Patients with chronic liver disease and cirrhosis<br />

have a dysregulated coagulation system and are prone<br />

to thrombosis. The basis for this so-called procoagulant imbalance<br />

is not completely understood. Tissue factor (TF) is the<br />

primary initiator of coagulation in vivo. Patients with cirrhosis<br />

have increased TF activity in white blood cells and circulating<br />

microparticles. The aim of our study was to determine the role<br />

of TF and the contribution of different cellular sources of TF to<br />

the hypercoagulable state in a mouse model of chronic liver<br />

injury. Methods: We measured levels of TF activity in the liver,<br />

white blood cells and microparticles, and a marker of activation<br />

of coagulation [thrombin-antithrombin complexes (TATc)]<br />

in the plasma of mice subjected to bile duct ligation for 12<br />

days. We used wild-type mice, mice with a global TF deficiency<br />

(low TF mice), and mice deficient for TF in either myeloid<br />

cells (TF flox/flox ,LysM Cre mice) or in hepatocytes (TF flox/flox /Albumin<br />

Cre ). Data are presented as median (interquartile range).<br />

Results: Wild-type mice with liver injury had increased levels<br />

of TATc [9.7 (7.6-10.5) vs. 6.0 (5.7-7.0) ng/mL; p


250A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

82<br />

The association of non-selective beta-blocker use and<br />

acute kidney injury in patients with decompensated<br />

cirrhosis admitted into hospital - A Study from the North<br />

American Consortium for the Study of End Stage Liver<br />

Diseae (NACSELD).<br />

Florence Wong 1 , Jacqueline G. O’Leary 2 , K. Rajender Reddy 3 ,<br />

Guadalupe Garcia-Tsao 4 , Scott W. Biggins 5 , Michael B. Fallon 6 ,<br />

Puneeta Tandon 7 , Ram M. Subramanian 8 , Benedict Maliakkal 9 ,<br />

Paul J. Thuluvath 10 , Hugo E. Vargas 11 , Patrick S. Kamath 12 , Leroy<br />

Thacker 13 , Jasmohan S. Bajaj 14,15 ; 1 Medicine, University of<br />

Toronto, Toronto, ON, Canada; 2 Medicine, Hepatology, Baylor<br />

University Medical Center, Dallas, TX; 3 Gastroenterology & Hepatology,<br />

University of Pennsylvania, Philadelphia, PA; 4 Digestive<br />

Diseases, Yale University School of Medicine, New Haven, CT;<br />

5 Gastroenterology, University of Colorado, Denver, CO; 6 Gastroenterology,<br />

Hepatology & Nutrition, University of Texas Health Science<br />

Center, Houston, TX; 7 Medicine, Gastroenterology, University<br />

of Alberta, Edmonton, AB, Canada; 8 Gastroenterology, Emory<br />

University Medical Centre, Atlanta, GA; 9 Gastroenterology, University<br />

of Rochester, Rochester, NY; 10 Institute of Digestive Health<br />

and Liver Disease, Mercy Hospital, Baltimore, MD; 11 Gastroenterology<br />

& Hepatology, Mayo Clinic, Scottsdale, AZ; 12 Internal<br />

Medicine, Gastroenterology & Hepatology, Mayo Clinic, Rochester,<br />

MN; 13 Biostatistics, Virginia Commonwealth University, Richmond,<br />

VA; 14 Gastroenterology, Hepatology & Nutrition, Virginia<br />

Commonwealth University, Richmond, VA; 15 Gastroenterology,<br />

McQuire VA Medical Center, Richmond, VA<br />

Non-selective beta-blocker (NSBB) use may be detrimental<br />

in patients with cirrhosis & refractory ascites, although this<br />

remains controversial. Theoretically, NSBB use in decompensated<br />

cirrhosis (DC) may decrease renal perfusion and predispose<br />

patients to renal impairment. Aim: To determine the<br />

impact of NSBB use on the risk i) for acute kidney injury (AKI)<br />

in hospitalized patients with DC, and ii) of infection to potentially<br />

accentuate this risk. Methods: NACSELD is a 16-center<br />

consortium that prospectively enrols non-elective cirrhotic inpatients.<br />

The presence of AKI (defined by Angeli et al, Gut 2015)<br />

at admission or its development during admission was evaluated<br />

and compared between those taking NSBB (NSBB+)<br />

vs those not (NSBB-) at admission. Further comparison was<br />

made between NSBB+ patients ± an infection during admission.<br />

Results: 981 cirrhotics (57±10 years, men: 64.6%,<br />

MELD: 19.5±7.3) with mostly alcoholic (43%) and hepatitis<br />

C (22.4%) cirrhosis were enrolled. Significantly more NSBB+<br />

patients had a history of variceal bleeding and diabetes<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 251A<br />

of AKI. Peak AKIN stage reached was stage III:II:I in 80(66%):<br />

31(25%):11(9%) patients, respectively. Presence of any E-OF<br />

was strongly associated with both development of new AKI<br />

(p=0.004, OR 6.7, 95% CI 1.86-24.1) and progression of<br />

AKI (p=0.04, OR 2.28, 95%CI 1.01-5.1). This risk further<br />

increased with increase in the number of E-OFs (p=0.001, OR<br />

1.41, 95%CI 1.2-1.7). Further, AKI at admission alone was not<br />

associated with mortality, but predicted mortality together with<br />

any E-OF (p150000) and liver stiffness<br />

(LS) ( 60% and an IQR/<br />

Median < 30% were considered. SS was US guided. Results:<br />

99 patients were included: 63% women; age 59 (± 6.6) years;<br />

BMI 28 (± 4.8); MELD 9.7 (± 3.4); 80% Child-Pugh A. LS was<br />

invalid in 2 patients and SS in 19. Platelets was 136.000 (±<br />

54.000), LS was 24 kPa (± 12) and SS was 53 kPa (± 18).<br />

Esophageal varices were detected in 54% of patients, 40%<br />

small and 14% large size. Gastric varices were present in 5%<br />

of patients, and GEV in 55%. Platelets (120.000 ± 45.000<br />

vs 155.000 ± 57.000), LS (28.4 ± 13.1 vs 18.4 ± 7.6 kPa)<br />

and SS (59.5 ± 17.0 vs 43.3 ± 17.9 kPA) were significantly<br />

different among patients with or without GEV (p<br />

6 months, n=83); and G4, past HCV infection (positive anti-<br />

HCV and negative HCV RNA PCR, n=15). The serum samples<br />

were tested using our HCV-Ags EIA before and after denaturation.<br />

RESULTS. All 38 patients in G1 tested negative for<br />

HCV-Ags, indicating 100% specificity of our HCV-Ags EIA. All<br />

85 patients with acute (G2, n=2) or chronic (G3, n=83) HCV<br />

infection (genotypes 1-6), and serum HCV RNA ranged from<br />

1,124 to 14,400,000 IU/mL quantified using PCR, tested positive<br />

for HCV-Ags with or without serum sample denaturation,<br />

indicating 100% sensitivity of the HCV-Ags EIA. The mean<br />

coefficient of variation (CVs) was 9.8% and 6.4%, respectively,<br />

for intra- precision assay and inter-assay reproducibility.<br />

In 2 cases with acute HCV infection, HCV-Ags were detectable<br />

59-65 days before anti-HCV test became positive. To further<br />

assess sensitivity, serum samples with low HCV RNA levels<br />

(214-1160 IU/mL) were serially diluted and tested. The lower<br />

limits of detection of the HCV-Ags assay using denatured or<br />

undenatured sera were equivalent to serum HCV RNA levels<br />

of 150 IU/mL and 250 IU/mL, respectively. Using denatured<br />

serum samples, 6/15 patients in G4 tested positive; however,<br />

none tested positive for HCV-Ags using undenatured serum<br />

samples. Additional <strong>studies</strong> showed that denaturation of these<br />

serum samples released HCV-Ags sequestered in HCV-ICs in<br />

these patients that may result in false positivity for active HCV<br />

infection. CONCLUSIONS. The highly specific and sensitive<br />

HCV-Ags EIA has a lower limit of detection equivalent to serum


252A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

HCV RNA levels of 150-250 IU/mL. Testing undenatured, but<br />

not denatured serum samples reliably differentiated active from<br />

past HCV infection and accomplished screening and diagnosis<br />

of HCV infection in one step.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ke-Qin Hu, Wei Cui<br />

86<br />

Hepatitis C Virus (HCV) Mortality Patterns in the British<br />

Columbia Hepatitis Testers Cohort (BC-HTC)<br />

Mel Krajden 1,2 , Amanda Yu 1 , Darrel Cook 1 , Margot E. Kuo 1 ,<br />

Maria Alvarez 1 , Mark Tyndall 3,2 , Naveed Z. Janjua 1,2 ; 1 Clinical<br />

Prevention Services, BC Centre for Disease Control, Vancouver,<br />

BC, Canada; 2 University of British Columbia, Vancouver, BC, Canada;<br />

3 BC Centre for Disease Control, Vancouver, BC, Canada<br />

Objective: To describe the population-level impacts of HCV<br />

acquisition risks vs. chronic infection on mortality for HCV<br />

negative and positive testers. Methods: The British Columbia<br />

Hepatitis Testers Cohort (BC-HTC) includes 1,135,947 individuals<br />

tested for HCV or reported to public health as a HCV<br />

case from 1990-2013, linked to their corresponding healthcare<br />

administrative data. Cohort entry was delayed one year<br />

for each individual and deaths within one year of HCV diagnosis<br />

were excluded, i.e., one year lagging. We computed<br />

age-adjusted annual all-cause, liver- and drug-related mortality<br />

rates/100,000 population and age-, sex- and year-adjusted<br />

standardized mortality ratios (SMRs) for: 1) seroconverters,<br />

those with known HCV acquisition timeframes and risk activities;<br />

2) anti-HCV positive on initial testing, the majority of whom<br />

were chronically infected and most of whom no longer engage<br />

in acquisition risk activities; and 3) HCV negatives. Results:<br />

Of 1,028,227 individuals included in this analysis, 64,876<br />

(6.3%) were HCV positive. Overall, 16.3% (10,572/64,876)<br />

of HCV positive vs. 6.4% (61,623/963,351) of HCV negative<br />

individuals died. For HCV positives, age-adjusted allcause<br />

mortality increased from 0.11/100,000 in 1993 to<br />

15.0 in 2005 and stayed in this range thereafter. Liver-related<br />

mortality increased from 0.03/100,000 in 1993 to 2.5 in<br />

2005 to 4.7 in 2013. Drug-related mortality increased from<br />

0.03/100,000 in 1993, peaked at 3.2 in 2005 and declined<br />

to 1.3 by 2013. All-cause, liver- and drug-related mortality<br />

rates were consistently higher for males vs. females. All-cause<br />

mortality for seroconverters was significantly higher than those<br />

with chronic HCV and HCV negatives (SMRs: 4.8, 3.1 and<br />

1.0). Liver-related mortality was significantly lower for seroconverters<br />

than for chronic HCV and was lowest for HCV negatives<br />

(SMRs: 10.4, 17.0 and 1.4). Similarly, liver cancer mortality<br />

was significantly lower for seroconverters than for chronic HCV<br />

and was lowest for HCV negatives (SMRs: 3.8, 17.5 and 1.3).<br />

However, drug-related mortality was highest for seroconverters<br />

compared to chronic HCV and HCV negatives (SMRs: 17.0,<br />

11.4 and 0.9). Conclusions: The BC-HTC enables comprehensive<br />

assessment of the mortality impacts of HCV infection and<br />

creates an opportunity to differentiate acquisition risk mortality<br />

from the impacts of chronic infection. These data illustrate that<br />

the excess mortality related to HCV acquisition risks is distinct<br />

from viral sequelae for a significant proportion of the HCV<br />

infected population. Therefore antiviral treatment on its own<br />

will not optimize mortality reductions in the affected populations.<br />

Disclosures:<br />

Mel Krajden - Grant/Research Support: Roche, Merck, Siemens, Boerhinger<br />

Ingelheim, Hologic<br />

The following authors have nothing to disclose: Amanda Yu, Darrel Cook, Margot<br />

E. Kuo, Maria Alvarez, Mark Tyndall, Naveed Z. Janjua<br />

87<br />

The Association of Sustained Virological Response and<br />

All-cause Mortality After Interferon-based Therapy for<br />

Chronic Hepatitis C (HCV) in a Large U.S. Community-based<br />

Health Care Delivery System<br />

Lisa M. Nyberg 1 , Xia Li 1 , Su-Jau Yang 1 , Kevin Chiang 1 , T. Craig<br />

Cheetham 1 , Susan Caparosa 1 , Jose R. Pio 1 , Zobair M. Younossi 2 ,<br />

Anders H. Nyberg 1 ; 1 Kaiser Permanente, San Diego, CA; 2 Department<br />

of Medicine, Inova Fairfax Medical Campus, Fairfax, VA<br />

Background: Previous <strong>studies</strong> performed at tertiary care centers<br />

have shown that sustained virological response (SVR) to interferon-based<br />

therapy for HCV is associated with a reduction<br />

of liver-related and all-cause mortality (1-2). The aim of this<br />

study was to determine if SVR to interferon-based therapy in<br />

a community-based health care setting is associated with a<br />

reduction in all-cause mortality and to examine for differences<br />

in those with and without cirrhosis. Methods: A retrospective<br />

cohort study at Kaiser Permanente, Southern California (KPSC),<br />

a large community-based integrated health care system including<br />

3.5 million members. Inclusion criteria: a diagnosis code<br />

and/or positive lab test for HCV RNA (index date) 1/1/02-<br />

12/31/13; age ≥ 18 years at index date, ≥ 12 months continuous<br />

membership before and after index date. Exclusion<br />

criteria: HCV diagnosis after 1/1/13 and liver transplant or<br />

HCC on or before index date. SVR was determined for patients<br />

treated for HCV with interferon-based therapy and stratified for<br />

cirrhosis vs non-cirrhosis. Mortality data were obtained from the<br />

California Department of Public Health Vital Statistics of California<br />

Report. Results: Total study cohort: 24,968 with chronic<br />

HCV; of those 10,449 (42%) had cirrhosis. Overall mortality<br />

during the study period was 18.5% (4,608/24,968). Mortality<br />

among those with cirrhosis was 33.2% (3,470/10,449) vs<br />

7.8% (1,138/14,519) among those without cirrhosis. Among<br />

patients treated for HCV, 45.1% (2348/5203) achieved SVR;<br />

mortality for those who achieved SVR was 5.4% (127/2348)<br />

vs 18.9% (540/2855) for those without SVR. Mortality in<br />

treated patients, with and without cirrhosis, is shown in Table<br />

1. Conclusions: An approximate 3-fold reduction in all-cause<br />

mortality is seen in patients with HCV who are treated and<br />

achieve SVR compared to those without SVR. Cirrhotics who<br />

are treated and achieve SVR have a > 3-fold reduction in mortality<br />

compared to the mortality in treated cirrhotics who do not<br />

achieve SVR. New, more potent, HCV treatment regimens with<br />

higher SVR have the potential of significantly reducing future<br />

mortality due to HCV. Further study is ongoing with matching<br />

for severity of liver disease to more specifically examine the<br />

effect of SVR on mortality compared to that due to the natural<br />

history of HCV/cirrhosis. 1. van der Meer, et al. JAMA 2012<br />

Dec26;308(24):2584-93. 2. Berenguer, et al. Hepatology<br />

2009 Aug;50(2):407-13<br />

Mortality in HCV patients treated with interferon-based therapy.<br />

Cirrhosis: N=2603; No Cirrhosis: N=2600<br />

Disclosures:<br />

Lisa M. Nyberg - Grant/Research Support: Merck, Gilead, Abbvie<br />

T. Craig Cheetham - Grant/Research Support: Gilead, BMS<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Xia Li, Su-Jau Yang, Kevin Chiang,<br />

Susan Caparosa, Jose R. Pio, Anders H. Nyberg


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 253A<br />

88<br />

Increasing Prevalence of Cirrhosis among US Adults<br />

with Chronic Hepatitis C Virus Infection: Results from<br />

NHANES 1988-1994 and 1999-2012<br />

Prowpanga Udompap, Ajitha Mannalithara, Nae-Yun Heo, Donghee<br />

Kim, W. Ray Kim; Division of Gastroenterology and Hepatology,<br />

Stanford University School of Medicine, Stanford, CA<br />

Background & Aims: Hepatitis C virus(HCV) is a major cause<br />

of end stage liver disease and hepatocellular carcinoma worldwide.<br />

While cirrhosis underlies these complications, the prevalence<br />

of HCV cirrhosis in the US remains unknown. We aim<br />

to determine the prevalence of advanced fibrosis/cirrhosis and<br />

to identify factors associated with advanced fibrosis/cirrhosis<br />

in people with chronic hepatitis C in the US population. Methods:<br />

Using the National Health and Nutrition Examination Survey(NHANES)<br />

data, we identified participants with detectable<br />

serum HCV RNA and assessed liver fibrosis using validated<br />

surrogate indicators, including APRI and FIB-4 scores. The prevalence<br />

of cirrhosis was determined for survey participants for<br />

Era 1 (1988-94), Era 2 (1999-2006) and Era 3 (2007-12).<br />

Results: Out of 52,644 participants with age ≥20, 736 (1.4%)<br />

had HCV. Based on APRI score, 6.6% (95%CI:2.2-11.0) of<br />

US adults with HCV infection in Era 1, 7.6% (95%CI:3.4-<br />

11.8) in Era 2 and 17.0% (95%CI:8.0-26.0) in Era 3 were<br />

estimated to have cirrhosis (Ishak stage 5-6). The prevalence<br />

of advanced fibrosis (Ishak stage 4-6) based on FIB-4 was<br />

8.6%, 10.1%, and 16.0% for Eras 1, 2, and 3, respectively.<br />

These data project to 370,500 Americans with cirrhosis and<br />

347,800 with advanced fibrosis in the most recent era. The<br />

higher prevalence of cirrhosis (determined by APRI) in the<br />

recent era was associated with increasing age (OR=1.04,<br />

95%CI:1.02-1.07), diabetes (OR=2.33, 95%CI:1.01-5.40)<br />

and obesity (OR=2.96, 95%CI:1.15-7.57). When FIB-4 score<br />

was used, advanced fibrosis was associated with increasing<br />

age (OR=1.08, 95%CI:1.05-1.11) and diabetes (OR=3.37,<br />

95%CI:1.24-9.5). Conclusion: Nearly one in five US adults<br />

with HCV infection have cirrhosis. The prevalence of cirrhosis<br />

is rising over time, which is in part attributable to the increasing<br />

age of the birth cohort with HCV infection. In addition, comorbidities<br />

such as diabetes and obesity accelerate liver fibrosis.<br />

These data further underscore the current recommendations<br />

for HCV screening in asymptomatic individuals and highlights<br />

the need for systematic assessment for liver fibrosis and comprehensive<br />

medical management in those with HCV infection.<br />

Characteristics of participants<br />

89<br />

Rectal Shedding of HCV in HCV/HIV Co-infected Men<br />

Andrew L. Foster 1,2 , Michael Gaisa 1 , Rosanne M. Hijdra 1,3 , Karen<br />

B. Jacobson 1 , Samuel Turner 1,2 , Tristan Morey 1,2 , Daniel S. Fierer 1 ;<br />

1 Infectious Diseases, Mount Sinai School of Medicine, New York,<br />

NY; 2 James Cook University, Cairns, QLD, Australia; 3 Amsterdam<br />

Medical Center, Amsterdam, Netherlands<br />

Introduction An epidemic of hepatitis C virus (HCV) infection<br />

is occurring among HIV-infected men who have sex with men<br />

(MSM). Epidemiological <strong>studies</strong> suggest that sexual transmission<br />

is fueling the epidemic. Blood and semen have been considered<br />

as potential mechanisms of transmission, but there are<br />

still transmission circumstances that remain unexplained. We<br />

hypothesized that HCV may be shed into rectal fluid of HCV/<br />

HIV co-infected MSM. Methods Written informed consent was<br />

obtained from 45 HIV-infected MSM with HCV infection and<br />

paired blood and rectal fluid specimens were obtained. Rectal<br />

fluid was collected using a moistened polyester-tipped swab<br />

that was gently inserted into the rectum. The swab was placed<br />

into transport medium, vortexed, and the supernatant analysed<br />

for HCV using COBAS AMPLICOR (Roche Diagnostics), lower<br />

limit of quantification 43 IU/mL, lower limit of detection 7 IU/<br />

mL. The experimentally-determined efficiency of HCV absorption<br />

to and elution from the swabs was used to calculate the<br />

rectal HCV viral load (VL). Rectal sexually-transmitted infection<br />

(STI) and blood syphilis testing were also performed. Results<br />

Visual inspection of the swabs showed no visible blood on<br />

any. HCV was detected in 20 (47%) of 43 specimens. Among<br />

those samples with HCV detected, the median rectal VL was<br />

2.92 log 10<br />

IU/mL (IQR 2.92, 4.27). Rectal HCV detection was<br />

associated with serum HCV VL >5 log 10<br />

IU/mL (p=0.011),<br />

and there was a high correlation between the magnitude of<br />

blood VL and rectal HCV VL (correlation coefficient 0.688,<br />

p


254A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

patients in care with hepatitis C virus infection (HCV) will<br />

achieve this status in the near future. However, the long term<br />

prognosis in terms of risk or predictors of developing hepatocellular<br />

carcinoma (HCC) among patients with SVR remains<br />

unclear. Methods: A retrospective cohort study using data from<br />

the national VA HCV Clinical Case Registry in patients with<br />

HCV diagnosed (positive HCV RNA) between 10/1999 and<br />

1/2009 who had at-least one year of follow up in the VA.<br />

HCV treatment (pegylated interferon with or without ribavirin)<br />

and SVR status (RNA test negative at least 12 weeks after<br />

the end of treatment) were ascertained. Cox proportional hazards<br />

models were used to examine the effect of demographic,<br />

virological and clinical factors on time to HCC among those<br />

who achieved SVR through end of 2011 while adjusting for<br />

medical comorbidity. We excluded those with HCC diagnostic<br />

codes before SVR date. Results: We had 33,005 HCV-infected<br />

individuals who received HCV treatment during the study timeframe.<br />

Of these, 10,817 patients achieved SVR and 10765<br />

had no HCC before SVR. Their mean age was 53.1 (sd, 6.3),<br />

12.4% were 60 years or older, and 12.9% were Black, 64.4%<br />

non-Hispanic white, and 3.4% Hispanic. Majority (95.3%)<br />

were males; and 53.6% had genotype 1, 25.0% genotype<br />

2, 13.5% genotype 3, 0.7% genotype 4, and 7.2% had no<br />

genotype test available. A total of 124 patients developed<br />

HCC during a total follow of 38,394 person year follow up<br />

(median duration of follow up after SVR was 2.7 years) at<br />

an incidence rate of 0.32% per year. In the cohort with SVR,<br />

patients 60 years and older (HR=6.59 c/w younger age),<br />

Hispanics (HR=2.16, 1.03-4.50 c/w non-Hispanic whites) and<br />

those with cirrhosis (HR=5.58, 3.21 - 9.68) or diabetes (1.82<br />

(0.92 - 3.61) were at an increased risk of developing HCC<br />

compared to their counterparts. Conclusions: The risk of HCC<br />

after HCV cure remains elevated (0.3% per year). Older age<br />

at the time of SVR, Hispanic ethnicity, and presence of cirrhosis<br />

and diabetes are associated with an increased HCC risk<br />

among those who achieve SVR.<br />

Disclosures:<br />

Hashem B. El-Serag - Grant/Research Support: WAKO, Gilead<br />

The following authors have nothing to disclose: Peter Richardson, Fasiha Kanwal<br />

91<br />

Prevalence of Pre-Treatment NS5A Resistance Associated<br />

Variants in Genotype 1 Patients Across Different<br />

Regions Using Deep Sequencing and Effect on Treatment<br />

Outcome with LDV/SOF<br />

Stefan Zeuzem 2 , Masashi Mizokami 3 , Stephen Pianko 4 , Alessandra<br />

Mangia 5 , Kwang-Hyub Han 6 , Ross Martin 1 , Evguenia<br />

S. Svarovskaia 1 , Hadas Dvory-Sobol 1 , Brian Doehle 1 , Phillip S.<br />

Pang 1 , Steven J. Knox 1 , John G. McHutchison 1 , Diana M. Brainard<br />

1 , Michael D. Miller 1 , Hongmei Mo 1 , Wan-Long Chuang 7 ,<br />

Ira M. Jacobson 8 , Gregory Dore 9 , Mark S. Sulkowski 10 ; 1 Gilead<br />

Sciences Inc, Foster City, CA; 2 Department of Internal Medicine,<br />

J.W. Goethe University Hospital, Frankfurt am Main, Germany;<br />

3 National Center for Global Health and Medicine, Tokyo, Japan;<br />

4 Department of Gastroenterology, Monash Medical Centre, Victoria,<br />

VIC, Australia; 5 Liver Unit, IRCCS-Ospedale Casa Sollievo<br />

della Sofferenza, San Giovanni Rotondo, Italy; 6 Department of<br />

Internal Medicine, Yonsei University College of Medicine, Seoul,<br />

Korea (the Republic of); 7 Internal Medicine, Kaohsiung Medical<br />

University Hospital, Kaohsiung, Taiwan; 8 Medicine, Mt. Sinai Beth<br />

Israel, New York, NY; 9 Kirby Institute, UNSW Australia, Sydney,<br />

NSW, Australia; 10 School of Medicine, Johns Hopkins University,<br />

Baltimore, MD<br />

Background: The efficacy of some HCV regimens is influenced<br />

by the presence of certain NS5A resistance associated variants<br />

(RAVs). To investigate if the pretreatment prevalence of<br />

NS5A RAVs in genotypes (GT) 1a and 1b varied by country,<br />

race or ethnicity, comprehensive analyses were performed<br />

using deep sequencing of NS5A from more than 5000 patients<br />

from 17 countries. Methods: NS5A deep sequencing analysis<br />

with a 1% cut-off was performed. GT1a RAVs were defined<br />

as K24G/N/R, M28A/G/T/V, Q30H/G/R/E/K, L31any,<br />

P32L, S38F, H58D, A92K/T, Y93any, and GT1b RAVs were<br />

L31any, P32L, P58D, A92K, Y93any. Results: Pretreatment<br />

samples from 5046 patients were analyzed. Prevalence of<br />

NS5A RAVs was similar between the different regions for both<br />

GT1a and GT1b HCV infected patients (Table). In addition, no<br />

significant differences were observed in NS5A RAV prevalence<br />

between patients with different race or ethnicity. In the subset of<br />

patients who were treated with LDV/SOF for 12 weeks, SVR12<br />

rates were similar in GT1b patients with and without pretreatment<br />

NS5A RAVs across all regions. In GT1a patients with<br />

pretreatment RAVs, SVR12 rates in North America were 91%<br />

(75/82) compared to 98% without NS5A RAVs. All seven<br />

GT1a patients who relapsed had pretreatment NS5A RAVs<br />

conferring >1000-fold reduced susceptibility to LDV (H58D,<br />

Y93H/N/F or multiple RAV combinations). However, 18 GT1a<br />

patients with similar RAVs conferring >1000-fold reduced<br />

susceptibility achieved SVR12 after LDV/SOF for 12 weeks.<br />

Conclusions: No significant difference in prevalence of NS5A<br />

pretreatment RAVs was observed between different regions,<br />

races or ethnicities. Overall, high SVR12 rates (91-100%) were<br />

observed in patients with or without NS5A RAVs with LDV/<br />

SOF. In GT1a patients, lower SVR rate (72%) was observed<br />

in patients with pretreatment NS5A RAVs conferring high level<br />

(>1000-fold) resistance to NS5A inhibitors.<br />

NA: Too few patients treated with LDV/SOF<br />

Disclosures:<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Stephen Pianko - Advisory Committees or Review Panels: Roche, Novartis, GIL-<br />

EAD, Roche, Novartis; Consulting: GILEAD; Speaking and Teaching: JANSSEN<br />

Alessandra Mangia - Advisory Committees or Review Panels: ROCHE, Janssen,<br />

MSD, ROCHE, Janssen, MSD, Boheringer ; Consulting: Gilead; Grant/Research<br />

Support: Shering-Plough, Shering-Plough<br />

Ross Martin - Employment: Gilead Sciences<br />

Evguenia S. Svarovskaia - Employment: Gilead Sciences Inc; Stock Shareholder:<br />

Gilead Sciences Inc<br />

Hadas Dvory-Sobol - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Brian Doehle - Employment: Gilead Sciences<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Steven J. Knox - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Michael D. Miller - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Hongmei Mo - Employment: Gilead Science Inc


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 255A<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Ira M. Jacobson - Consulting: AbbVie, Achillion, Alnylam, Bristol Myers Squibb,<br />

Enanta, Gilead, Janssen, Merck; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead, Janssen, Merck, Tobira; Speaking and Teaching: AbbVie, Bristol<br />

Myers Squibb, Gilead, Janssen<br />

Gregory Dore - Board Membership: Gilead, Merck, Abbvie, Bristol-Myers<br />

Squibb; Grant/Research Support: Gilead, Merck, Abbvie, Bristol-Myers Squibb;<br />

Speaking and Teaching: Gilead, Merck, Abbvie, Bristol-Myers Squibb<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

The following authors have nothing to disclose: Masashi Mizokami, Kwang-Hyub<br />

Han<br />

92<br />

Highly Successful Retreatment with Ledipasvir (LDV) and<br />

Sofosbuvir (SOF) in HCV GT-1 Patients Who Failed Initial<br />

Short Course Therapy with Combination DAA Regimens<br />

(NIH SYNERGY Trial)<br />

Eleanor Wilson 1,2 , Sarah Kattakuzhy 1 , Zayani Sims 2 , Mary<br />

McLaughlin 3 , Angie Price 1 , Hongmei Mo 4 , Anu Osinusi 4 , Henry<br />

Masur 2 , Anita Kohli 5 , Shyam Kottilil 1 ; 1 Infectious Disease, Institute<br />

for Human Virology/University of Maryland School of Medicine,<br />

Rockville, MD; 2 Critical Care Medicine Department, National Institutes<br />

of Health, Bethesda, MD; 3 National Institute of Allergy and<br />

Infectious Diseases, National Institutes of Health, Bethesda, MD;<br />

4 Gilead Sciences, Foster City, CA; 5 Hepatology, St. Joseph’s Hospital<br />

and Medical Center, Phoenix, AZ<br />

Introduction: Directly-acting antivirals (DAAs) dramatically<br />

improved HCV therapy but retreatment options for patients<br />

who have failed therapy with combination DAAs have not<br />

been studied. Our study aim is to determine if HCV genotype-1<br />

patients who failed short course combination therapy with 3<br />

or 4 DAAs can be successfully treated with 12 weeks of LDV/<br />

SOF. Material and Methods: In this single-center, open-label,<br />

phase 2a trial, HCV mono-infected persons with early stage<br />

(F0-F2) liver fibrosis and previous exposure to combination<br />

DAA therapy only (LDV/SOF with GS-9451 +/- GS-9669)<br />

were eligible to enroll and receive 12 weeks of LDV/SOF.<br />

HCV RNA was measured with the Abbott assay, lower level<br />

of quantitation (LLOQ) 12 IU/ml. The primary endpoint was<br />

defined as HCV viral load (VL) < LLOQ 12 weeks after end<br />

of therapy (SVR12). Deep sequencing of NS5B and NS5A<br />

regions was performed at baseline by Illumina next generation<br />

sequencing technology. Results: The study enrolled 34 persons,<br />

and 32 (94%) completed therapy with 12 weeks of LDV/<br />

SOF. Two patients withdrew consent after Day 0. Participants<br />

were predominantly male (82%) and black (85%), median<br />

age 60.5 years (IQR 57.0 – 63.8) and BMI 26.8 kg/m 2 (IQR<br />

25.3 – 29.3). Baseline HCV VL was 1.3 x 10^6 IU/mL (IQR<br />

5.8x10^5 – 3.9x10^6), 73.5% (25/34) were infected with<br />

HCV genotype 1a, and median Metavir fibrosis stage was<br />

1. Time from relapse to retreatment was 22 weeks (IQR 18 –<br />

23). SVR12 rates were 91% (31/34; ITT) and 97% (31/32,<br />

per protocol). For SVR rates by original treatment group (per<br />

protocol), see Figure. At baseline, 28/33 patients (85%) had<br />

resistance-associated variants (RAVs) consistent with >25 fold<br />

resistance in NS5A. Of all patients completing therapy, 1<br />

patient with NS5A RAVs relapsed. Conclusions: For the first<br />

time, we demonstrate a high SVR rate following retreatment<br />

with DAAs in patients who have previously failed DAA-only<br />

therapy.<br />

Disclosures:<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Anu Osinusi - Employment: gilead sciences<br />

The following authors have nothing to disclose: Eleanor Wilson, Sarah Kattakuzhy,<br />

Zayani Sims, Mary McLaughlin, Angie Price, Henry Masur, Anita Kohli,<br />

Shyam Kottilil<br />

93<br />

Effectiveness of Ledipasvir/Sofosbuvir in Treatment<br />

Naïve Genotype 1 Patients Treated in Routine Medical<br />

Practice<br />

Lisa I. Backus 1,2 , Pamela S. Belperio 1 , Troy Shahoumian 1 , Timothy<br />

P. Loomis 1 , Larry A. Mole 2 ; 1 Office of Public Health/Population<br />

Health, Veterans Affairs Palo Alto Health Care System, Palo Alto,<br />

CA; 2 Department of Medicine, VA Palo Alto Healthcare System,<br />

Palo Alto, CA<br />

Aim: Assess the effectiveness of ledipasvir/sofosbuvir±ribavirin<br />

(LDV/SOF±RBV) in treatment naïve genotype 1 (GT1) hepatitis<br />

C virus (HCV)-infected veterans treated in routine medical practice.<br />

Methods: This observational, intent-to-treat cohort analysis<br />

used the Veterans Affairs’ Clinical Case Registry to identify all<br />

treatment naïve GT1 HCV-infected veterans initiating 8 or 12<br />

weeks of LDV/SOF±RBV by 31 December 2014. Patients were<br />

excluded for liver transplantation or baseline HCV RNA


256A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ics with baseline HCV RNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 257A<br />

95<br />

Improving liver function and delisting of patients awaiting<br />

liver transplantation for HCV cirrhosis: do we ask<br />

too much to DAA?<br />

Audrey Coilly 1,2 , Georges-Philippe Pageaux 22 , Pauline Houssel-Debry<br />

7 , Christophe Duvoux 3 , Sylvie Radenne 6 , Victor de<br />

Ledinghen 15 , Danielle Botta-Fridlund 11 , Anaïs Vallet-Pichard 13 ,<br />

Rodolphe Anty 9 , Vincent Di Martino 5 , Filomena Conti 10 , Marie Line<br />

Debette-Gratien 20 , Laurent Alric 16 , Armando Abergel 17 , Camille<br />

Besch 19 , Helene Montialoux 18 , Pascal Lebray 10 , Sebastien Dharancy<br />

4 , Francois Durand 12 , Louis d’Alteroche 21 , Florian Charier 8 ,<br />

Olivier Chazouillères 14 , Jérôme Dumortier 24 , Vincent Leroy 23 , Jean-<br />

Charles Duclos-Vallee 1,2 ; 1 Centre Hepato-Biliaire, AP-HP Hopital<br />

Paul-Brousse, Villejuif, France; 2 Unit 1193, INSERM, Villejuif,<br />

France; 3 AP-HP Hopital Henri Mondor, Creteil, France; 4 CHRU<br />

de Lille, Lille, France; 5 CHU Jean Minjoz, Besançon, France; 6 HCL<br />

- Hopital Croix Rousse, Lyon, France; 7 CHU de Rennes - Hopital<br />

Pontchaillou, Renes, France; 8 CHU de Poitiers, Poitiers, France;<br />

9 CHU de Nice, Nice, France; 10 AP-HP Hopital Pitie-Salpetrière,<br />

Paris, France; 11 AP-HM Hopital La Conception, Marseille, France;<br />

12 AP-HP Hopital Beaujon, Clichy, France; 13 AP-HP Hopital Cochin,<br />

Paris, France; 14 AP-HP Hopital Saint-Antoine, Paris, France; 15 CHU<br />

de Bordeaux - Hopital Haut-Leveque, Pessac, France; 16 CHU de<br />

Toulouse - Hopital Purpan, Toulouse, France; 17 CHU de Clermont-Ferrand,<br />

Clermont-Ferrand, France; 18 CHU de Rouen, Rouen,<br />

France; 19 CHU de Strasbourg - Hopital Hautepierre, Strasbourg,<br />

France; 20 CHU de Limoges, Limoges, France; 21 CHU de Tours<br />

- Hopital Trousseau, Tours, France; 22 CHU de Montpellier - Hopital<br />

Saint-Eloi, Montpellier, France; 23 CHU de Grenoble - Hopital<br />

Michallon, Grenoble, France; 24 HCL - Hopital Edouard Herriot,<br />

Lyon, France<br />

Background: Combinations of DAA have shown excellent<br />

results to treat HCV-infection in cirrhotic patients (pts). But some<br />

issues remain unresolved regarding efficacy in pts awaiting<br />

liver transplantation (LT) and impact on access to LT. Methods:<br />

This cohort study enrolled 151 pts registered in 23 centers<br />

on French LT list (male: 79%, age 56±7 years), treated with<br />

sofosbuvir ± ribavirin (n=84) ± daclatasvir (n=90) or simeprevir<br />

(n=9) or ledipasvir (n=17). Meantime between listing and<br />

starting therapy was 23±57 weeks (wks). All pts were cirrhotic.<br />

LT indication was HCC in 85(56%) pts. 112(74%) pts were<br />

treatment experienced including 46% of non-responders, failures<br />

to 1rst generation protease inhibitors in 32 or sofosbuvir<br />

in 16 pts. Majority of pts were genotype 1 (56%) and 3 (24%).<br />

Results: Mean follow-up was 44±19 wks [12-79]. Mean baseline<br />

MELD score (MELD), platelet count, bilirubin and albumin<br />

levels were 10±5 [6-32], 89±50 G/L[23-347], 39±42μmol/L<br />

[5-405], 33±7 g/L [15-47]. 73 pts (48%) were decompensated<br />

(bilirubin > 50μmol/L in 34(23%) ± ascites in 53(35%)<br />

± hepatic encephalopathy (HE) in 29(19%)). Indetectability<br />

of HCV RNA was obtained at 7±4 wks [1-24]. Among 117<br />

pts, 103(88%) achieved SVR12. Among 73 decompensated<br />

pts, a complete response (normal bilirubin level, no ascites,<br />

no HE) was observed in 14(19%) after 12 wks and 31(42%)<br />

at 12 wks post-treatment. When achieving SVR12, ascites or<br />

HE persisted in 12/44 (27%) and 7/14 (50%). Variations of<br />

MELD are shown in Table 1. Among patients with baseline<br />

MELD score≥20, 83% still an indication of LT at the end of treatment<br />

(MELD≥15). 45(30%) pts were transplanted. Ten (6%) pts<br />

were delisted because of liver function improvement. A serious<br />

adverse event occurred in 24(19%) pts, mainly liver events<br />

(37%), anemia (25%) and sepsis (25%). Conclusion: Among<br />

151 cirrhotic pts awaiting LT treated with IFN-free regimens,<br />

88% achieved SVR12. Only 6% of pts were delisted due to<br />

liver function improvement. A complete clinical and biological<br />

response was observed in 42%, raising hopes that more pts<br />

could be delisted. However, major clinical improvement was<br />

observed in a minority of patients especially when MELD score<br />

was ≥20, suggesting that antiviral treatment might be postponed<br />

after LT in this population. Final results will be provided<br />

in a larger population.<br />

Variations of MELD score in 66 pts with decompensated condition<br />

from baseline to the end of treatment<br />

Disclosures:<br />

Audrey Coilly - Consulting: Novartis, Astellas, Janssen, Bristol-Myers-Squibb,<br />

Merck Sharp & Dohme, Gilead, Roche<br />

Georges-Philippe Pageaux - Advisory Committees or Review Panels: Roche,<br />

Roche, Roche, Roche; Board Membership: Astellas, Astellas, Astellas, Astellas<br />

Pauline Houssel-Debry - Speaking and Teaching: NOVARTIS, ASTELLAS, GILEAD<br />

Christophe Duvoux - Advisory Committees or Review Panels: Novartis, Roche,<br />

Novartis, Roche, Novartis, Roche, Novartis, Roche; Speaking and Teaching:<br />

Astellas, Astellas, Astellas, Astellas<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

Danielle Botta-Fridlund - Consulting: GILEAD, ABBVIE, BMS, MSD<br />

Anaïs Vallet-Pichard - Independent Contractor: Schering Plough, Gilead, BMS,<br />

Roche<br />

Vincent Di Martino - Advisory Committees or Review Panels: Gilead, France,<br />

Abbvie, BMS France; Board Membership: MSD France; Consulting: Gilead,<br />

France; Speaking and Teaching: Janssen, BMS France, Gilead France<br />

Laurent Alric - Board Membership: Schering Plough, Schering Plough, Schering<br />

Plough, Schering Plough; Consulting: MSD; Speaking and Teaching: Roches,<br />

BMS, Gilead, Roches, BMS, Gilead, Roches, BMS, Gilead, Roches, BMS, Gilead,<br />

MSD, Abbvie<br />

Armando Abergel - Consulting: gilead, msd, bms; Speaking and Teaching: abbvie<br />

Pascal Lebray - Grant/Research Support: Schering-Plough, Schering-Plough, Schering-Plough,<br />

Schering-Plough; Speaking and Teaching: Gilead, Gilead, Gilead,<br />

Gilead<br />

Sebastien Dharancy - Advisory Committees or Review Panels: CHIESI; Board<br />

Membership: NOVARTIS; Speaking and Teaching: ASTELLAS<br />

Francois Durand - Advisory Committees or Review Panels: Astellas, Novartis,<br />

BMS; Speaking and Teaching: Gilead<br />

Olivier Chazouillères - Consulting: APTALIS, MAYOLY-SPINDLER<br />

Jérôme Dumortier - Board Membership: Novartis, Astellas, Roche; Consulting:<br />

Novartis; Grant/Research Support: Novartis, Astellas, Roche, MSD, GSK<br />

Vincent Leroy - Board Membership: Abbvie, BMS, Gilead; Consulting: Janssen,<br />

MSD; Speaking and Teaching: Abbvie, BMS, Gilead, Janssen, MSD<br />

The following authors have nothing to disclose: Sylvie Radenne, Rodolphe Anty,<br />

Filomena Conti, Marie Line Debette-Gratien, Camille Besch, Helene Montialoux,<br />

Louis d’Alteroche, Florian Charier, Jean-Charles Duclos-Vallee<br />

96<br />

Detecting Drug-Induced Liver Injury in Patients with<br />

Decompensated Chronic Hepatitis C: A Review of the<br />

SOLAR-1 and SOLAR-2 Studies<br />

Andrew J. Muir 2 , Michael R. Charlton 5 , Phillip S. Pang 1 , Luisa<br />

M. Stamm 1 , Diana M. Brainard 1 , John G. McHutchison 1 , Nezam<br />

H. Afdhal 3 , Paul B. Watkins 4 ; 1 Gilead Sciences, Foster City, CA;<br />

2 Duke University Medical Center, Durham, NC; 3 Division of Gastroenterology,<br />

BIDMC, Boston, MA; 4 The Hamner-UNC Institute for<br />

Drug Safety Sciences, Chapel Hill, NC; 5 Intermountain Medical<br />

Center, Murray, UT<br />

Background: The identification of drug-induced liver injury<br />

(DILI) is challenging in patients with underlying liver disease,<br />

particularly in the setting of decompensated cirrhosis which has<br />

recently become a target population for novel therapies. Traditional<br />

laboratory criteria (e.g., elevation in serum ALT > 5x<br />

upper limit of normal) may not be appropriate for patients with


258A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

advanced liver disease. The aim of this analysis was to develop<br />

new criteria for identifying DILI associated with direct-acting<br />

antivirals in end-stage liver disease patients. Methods: The<br />

SOLAR-1 and SOLAR-2 <strong>studies</strong> assessed the safety and efficacy<br />

of ledipasvir/sofosbuvir plus ribavirin (LDV/SOF+RBV) for 12<br />

or 24 weeks in patients with advanced liver disease pre- or<br />

post-liver transplant. Laboratory data, including total and direct<br />

bilirubin and ALT and AST relative to baseline, from patients<br />

with CPT B and C cirrhosis (N=328) were used to determine<br />

criteria to screen for possible cases of DILI. Cases of interest,<br />

including clinical outcome data, were further evaluated for DILI<br />

by an expert panel. DILI was classified as possible or unlikely,<br />

where unlikely required the identification of a clear non-drug<br />

related etiology of the decompensation or laboratory abnormalities.<br />

Results: Overall, on-treatment elevations above the<br />

upper limit of normal were common for total bilirubin (90%,<br />

n=295), ALT (48%, n=157) and AST (90%, n=294). Further<br />

analysis of total bilirubin identified only 11 of 295 patients in<br />

whom the direct bilirubin had a treatment-emergent elevation of<br />

greater than 1 mg/dL. Adjudication of these 11 cases revealed<br />

1 possible case of DILI. The use of ALT and AST criteria, including<br />

a cutoff of 5x nadir, did not identify further cases of interest<br />

beyond those identified by the >1 mg/dL increase in direct<br />

bilirubin. Median ALT and AST at baseline were 62 IU/ml and<br />

100 IU/ml, respectively; only 3 subjects reached a maximum<br />

ALT or AST > 2x baseline. ALT and AST improved (on-treatment<br />

ALT maximum < baseline) in 91% and 92% of patients,<br />

respectively. Conclusions: In patients with CPT B or C cirrhosis<br />

and HCV treated with LDV/SOF+RBV, elevation from baseline<br />

in direct bilirubin of > 1mg/dL was a clinically sensitive and<br />

conservative cutoff that identified patients who should be further<br />

evaluated for the possibility of DILI. No additional possible<br />

cases of DILI were identified by using ALT or AST criteria. An<br />

adjudication of all biochemical and clinical events of interest<br />

(including deaths and liver transplantations) of all patients in<br />

the SOLAR-1 and SOLAR-2 <strong>studies</strong> will be presented.<br />

Disclosures:<br />

Andrew J. Muir - Advisory Committees or Review Panels: BMS, Gilead, Janssen,<br />

Merck; Consulting: Theravance; Grant/Research Support: Abbvie, Abbvie, BMS,<br />

Gilead, Janssen, Merck, Achillion, Lumena<br />

Michael R. Charlton - Grant/Research Support: GIlead Sciences, Merck, Janssen,<br />

AbbVie, Novartis<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Luisa M. Stamm - Employment: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Nezam H. Afdhal - Advisory Committees or Review Panels: Trio Helath Care;<br />

Board Membership: Journal Viral hepatitis; Consulting: Merck, EchoSens, BMS,<br />

Achillion, GlaxoSmithKline, Springbank, Gilead, AbbVie; Grant/Research Support:<br />

Gilead; Stock Shareholder: Springbank<br />

Paul B. Watkins - Consulting: Abbott, Actelion, Boerringer-Ingelheim, Cempra,<br />

Alecra, Roche, Merck, Reservlogix, Intercept, Janssen, Novartis, Otsuka, Pfizer,<br />

Sanolfi, Takeda, UCB, Bristol-Myers Squibb, GSK<br />

the underlying mechanism driving apoptosis remains unclear.<br />

In this study we investigated the role of autophagy in regulating<br />

apoptosis in NAFLD. Methods and Results: Administration of<br />

palmitic acid (PA) to HepG2 cells increased Annexin V-positive/7-AAD-negative<br />

cells and induced apoptosis from 8 hours,<br />

which was found to be mediated by activation of IRE1-JNK and<br />

PERK-CHOP pathways. PA increased the expression level of<br />

LC3-II and decreased the autophagic flux index, which can be<br />

measured by monitoring LC3-II turnover using bafilomycin A1,<br />

from 3 hours. These findings suggest that PA inhibits the autophagic<br />

process after autophagosome formation. In PA-treated<br />

HepG2 cells, mTOR signaling, a negative regulator of autophagy,<br />

was inhibited and the expressions levels of Beclin1,<br />

Atg7 and Atg5 were unaltered. In contrast, the expression<br />

level of Rubicon, a negative regulator of autophagosome-lysosome<br />

fusion, was clearly upregulated by PA treatment from<br />

2 hours. siRNA-mediated Rubicon knockdown clearly abolished<br />

PA-induced inhibition of autophagic flux, reduced ER<br />

stress and ameliorated PA-induced apoptosis. While PA treatment<br />

did not change the levels of Rubicon mRNA expression,<br />

a pulse chase assay revealed that Rubicon protein stability<br />

was enhanced by PA treatment. Consistent with in vitro findings,<br />

mice on high fat diet (HFD) showed increased Rubicon<br />

expression from 1 month, autophagy inhibition from 2 months<br />

and increased apoptosis from 3 months in the liver. Electron<br />

microscopy revealed that isolation membranes and autophagosomes<br />

in the liver of mice fed HFD were increased, suggesting<br />

the inhibition of autophagy after autophagosome maturation.<br />

Hepatocyte-specific Rubicon knockout (KO) mice on HFD for 4<br />

months improved autophagy inhibition, ER stress and hepatocyte<br />

apoptosis compared with wild-type littermates. They also<br />

showed the significant reduction of liver volume to approximately<br />

normal level, which was accompanied by significant<br />

and substantial reduction of triglycerides in their livers. Finally,<br />

we investigated Rubicon expression levels in human liver samples<br />

by western blotting under approval of institutional ethics<br />

committee. Rubicon expression levels in steatotic livers were<br />

higher than those in non-steatotic livers. Conclusion: Increased<br />

expression of Rubicon induced by PA and HFD suppresses<br />

autophagic flux and promotes hepatocyte apoptosis by increasing<br />

ER stress. Enhancement of autophagy by inhibiting Rubicon<br />

may provide new approaches for treatment of NAFLD.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Satoshi Tanaka, Sadatsugu<br />

Sakane, Yasutoshi Nozaki, Yugo Kai, Tasuku Nakabori, Yoshinobu Saito,<br />

Ryotaro Sakamori, Tomohide Tatsumi<br />

97<br />

Enhanced expression of Rubicon inhibits autophagy and<br />

promotes apoptosis by increasing endoplasmic reticulum<br />

stress in nonalcoholic fatty liver disease<br />

Satoshi Tanaka, Hayato Hikita, Sadatsugu Sakane, Yasutoshi<br />

Nozaki, Yugo Kai, Tasuku Nakabori, Yoshinobu Saito, Ryotaro<br />

Sakamori, Tomohide Tatsumi, Tetsuo Takehara; Gastroenterology<br />

& Hepatology, Osaka University, Suita, Japan<br />

Background and Aim: Hepatocyte apoptosis is a characteristic<br />

feature of nonalcoholic fatty liver disease (NAFLD). However,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 259A<br />

98<br />

In vivo reprogramming of myofibroblasts into hepatocytes<br />

as a therapy for liver fibrosis<br />

Milad Rezvani 1 , Regina Español-Suñer 1 , Yann Malato 1 , Laure<br />

Dumont 1 , Andrew A. Grimm 2 , Eike Kienle 3 , Julia Bindmann 1 , Ellen<br />

Wiedtke 3 , Syed J. Naqvi 1 , S. C. Derderian 4 , Robert F. Schwabe 5 ,<br />

Dirk Grimm 3 , Holger Willenbring 1,6 ; 1 Institute of Regeneration<br />

Medicine, University of California San Francisco, San Francisco,<br />

CA; 2 Department of Pediatrics, Division of Gastroenterology,<br />

Hepatology, and Nutrition, University of California San Francisco,<br />

San Francisco, CA; 3 Department of Infectious Diseases, Cluster of<br />

Excellence CellNetworks, BioQuant BQ0030, Heidelberg University<br />

Hospital, Heidelberg, Germany; 4 4Department of Surgery,<br />

Division of Pediatric Surgery, University of California San Francisco,<br />

San Francisco, CA; 5 Department of Medicine,, Columbia<br />

University, New York, NC; 6 Department of Surgery, Division of<br />

Transplantation, University of California San Francisco, San Francisco,<br />

CA<br />

Repeated hepatocyte loss in chronic liver injury can exceed<br />

the regenerative capabilities of hepatocytes and lead to liver<br />

fibrosis, a form of scarring characterized by replacement of<br />

hepatocytes by collagen produced by myofibroblasts (MFs).<br />

The structural and molecular changes associated with liver<br />

fibrosis further impair liver function, leading to liver failure.<br />

The only cure for advanced liver fibrosis is liver transplantation,<br />

but donor organs are scarce. As a strategy to replenish<br />

the hepatocyte mass and limit collagen deposition in the<br />

chronically injured liver, we developed in vivo reprogramming<br />

of MFs into hepatocytes by overexpression of hepatic transcription<br />

factors. To facilitate clinical translation, we delivered<br />

these genes to MFs using adeno-associated viral (AAV) vectors,<br />

which are not toxic and do not integrate into the genome. We<br />

first investigated the feasibility of generating induced hepatocytes<br />

from primary MFs (MF-iHeps) using AAV vectors expressing<br />

FOXA1, FOXA2, FOXA3, GATA4, HNF1α and HNF4α in<br />

vitro. We found that AAV vectors were effective in generating<br />

expandable MF-iHeps resembling previously reported iHeps<br />

generated from fibroblasts with integrating retroviral or lentiviral<br />

vectors. Specifically, MF-iHeps lost most of their original<br />

cell identity and acquired hepatocyte gene and protein expression<br />

and functions like low-density lipoprotein uptake, glycogen<br />

storage and cytochrome P450 activity. Next, we established<br />

hepatic reprogramming of MFs in vivo. For this we used a<br />

lineage-tracing mouse model in which MFs and their progeny<br />

are constitutively labeled. To induce liver fibrosis, we treated<br />

these mice with carbon tetrachloride. After intravenous injection<br />

of the AAV vectors we observed the formation of MF-iHeps<br />

and reduced liver fibrosis. Like primary hepatocytes MF-iHeps<br />

proliferated in response to liver injury in both CCl4-treated<br />

and fumarylacetoacetate hydrolase-deficient mice. To assess<br />

hepatocyte differentiation of MF-iHeps, we isolated them by<br />

laser-capture microscopy and analyzed their gene expression<br />

with microarrays. MF-iHeps closely resembled primary hepatocytes,<br />

with the exception of minimal residual MF identity. We<br />

confirmed these results using functional assays, including analysis<br />

of hepatic glucose metabolism, and ascertained stability<br />

of MF-iHep differentiation in mice followed for 6 months. Our<br />

results establish repurposing of MFs as a potential new therapy<br />

for liver fibrosis that not only reduces fibrosis but also increases<br />

the functional hepatocyte mass. By using AAV vectors, which<br />

proved effective and safe in liver-directed human gene therapy,<br />

our strategy lends itself well to clinical translation.<br />

Disclosures:<br />

Robert F. Schwabe - Consulting: Merck<br />

The following authors have nothing to disclose: Milad Rezvani, Regina Español-<br />

Suñer, Yann Malato, Laure Dumont, Andrew A. Grimm, Eike Kienle, Julia Bindmann,<br />

Ellen Wiedtke, Syed J. Naqvi, S. C. Derderian, Dirk Grimm, Holger<br />

Willenbring<br />

99<br />

Protective role of miR-122 against acetaminophen toxicity<br />

is due to suppression of Cyp2e1 and Cyp1a2<br />

Vivek K. Chowdhary 3 , Huban Kutay 3 , Laura James 2 , William M.<br />

Lee 1 , Kalpana Ghoshal 3 ; 1 UT Southwestern Medical Center, Dallas,<br />

TX; 2 University of Arkansas for Medical Science, Little Rock,<br />

AR; 3 Pathology, Comprehensive Cancer Center, The Ohio State<br />

University College of Medicine, Columbus, OH<br />

Approximately 2000 cases of acute liver failure occur annually<br />

in the United States and acetaminophen (APAP) accounts for<br />

nearly 50% of cases. miR-122 is the most abundant, conserved<br />

liver-specific microRNA that maintains metabolic homeostasis<br />

and functions as a tumor suppressor. Although circulating<br />

miR-122 is a sensitive biomarker of APAP toxicity in humans<br />

and rodents, its role in this injury has not been elucidated. To<br />

investigate whether miR-122 has any protective role against<br />

APAP toxicity, we gave APAP (500mg/Kg) intraperitoneally<br />

to the Mir122 fl/fl (WT) and liver-specific miR-122 knockout<br />

(Mir122 fl/fl ; Alb-Cre) (aka LKO) mice generated in our lab<br />

(PMID: 22820288), and monitored their survival. Mortality<br />

rate of LKO mice was significantly higher than that of WT mice<br />

(P


260A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

100<br />

CX3CR1 is a gatekeeper for intestinal barrier integrity:<br />

Limiting steatohepatitis by maintaining intestinal<br />

homeostasis<br />

Kai M. Schneider 1 , Veerle Bieghs 1 , Felix Heymann 1 , Wei Hu 1 ,<br />

Daniela Dreymueller 2 , Lijun Liao 1 , Mick Frissen 1 , Andreas Ludwig 2 ,<br />

Nikolaus Gassler 3 , Oliver Pabst 4 , Eicke Latz 5,6 , Gernot Sellge 1 ,<br />

John Penders 7 , Frank Tacke 1 , Christian Trautwein 1 ; 1 Department<br />

of Internal Medicine III, RWTH Aachen University, Aachen, Germany;<br />

2 Institute of Pharmacology and Toxicology, Medical Faculty,<br />

University Hospital RWTH Aachen, Aachen, Germany; 3 Institute<br />

of Pathology, University Hospital RWTH Aachen, Aachen, Germany;<br />

4 Institute of Molecular Medicine, University Hospital RWTH<br />

Aachen, Aachen, Germany; 5 Institute of Innate Immunity, University<br />

Hospital, University of Bonn, Bonn, Germany; 6 Department of<br />

Medicine, University of Massachusetts Medical School, Worcester,<br />

MA; 7 Department of Medical Microbiology, School of Nutrition<br />

and Translational Research in Metabolism, Maastricht University<br />

Medical Center, Maastricht, Netherlands<br />

Background: Non-alcoholic fatty liver disease (NAFLD) represents<br />

the most common liver disease in Western societies<br />

and is regarded as the hepatic manifestation of the metabolic<br />

syndrome. The G-Protein-coupled chemokine receptor Cx3cr1<br />

has been shown to play a central role in many metabolic diseases<br />

including type-2 diabetes, atherosclerosis and obesity.<br />

However, the role of Cx3cr1 for NAFLD progression is unclear.<br />

Recent data indicate that intestinal dysbiosis drives NASH<br />

development and Cx3cr1 is essential for intestinal homeostasis.<br />

Based on these findings, we hypothesized that Cx3cr1<br />

plays a role in regulating the gut-liver axis and therefore has<br />

implications for NASH progression. Methods: Male wildtype<br />

(WT) and Cx3cr1 -/- mice were fed either a normal chow diet<br />

(NCD), high-fat diet (HFD) or methionine-choline-deficient diet<br />

(MCD) to induce steatohepatitis. For eradicating the intestinal<br />

microbiota, the drinking water of the mice was supplemented<br />

with broad-spectrum non-resorbable antibiotics. Cx3cr1-signaling<br />

was studied using Bone-marrow-derived macrophages<br />

(BMDMs). Results: On HFD or MCD Cx3cr1 -/- mice showed<br />

more severe hepatic steatosis and inflammation, as well as systemic<br />

glucose intolerance compared to WT controls. Mechanistically,<br />

in-vitro experiments with BMDMs identified Cx3cr1 as<br />

a regulator of macrophage homeostasis by modulating inflammasome<br />

activation. Accordingly, Cx3cr1 deficiency in mice<br />

was associated with significantly altered intestinal microbiota<br />

composition, which was linked to an impaired intestinal barrier<br />

including thinner colonic mucus layers, reduction of tight junction<br />

expression and a decrease in colonic resident phagocytic<br />

macrophages. Concomitantly, these intestinal changes led to<br />

an increased translocation of endotoxin and stronger activation<br />

of pathogen recognition receptors (PRRs) in livers of Cx3cr1 -/-<br />

mice, thereby triggering an enhanced inflammatory response<br />

in the liver. Strikingly, depletion of the intestinal microbiota<br />

by administration of broad-spectrum antibiotics (AB) did not<br />

only suppress the number of infiltrating macrophages, but also<br />

promoted a restorative phenotype of liver macrophages. Consequently,<br />

AB-treated mice demonstrated a marked improvement<br />

of steatohepatitis and glucose tolerance. Conclusion: Our<br />

data demonstrate that microbiota-mediated activation of the<br />

innate immune responses via Cx3cr1 is crucial for controlling<br />

steatohepatitis progression, thereby recognizing CX3CR1 as<br />

an essential gatekeeper in this scenario.<br />

Disclosures:<br />

Frank Tacke - Advisory Committees or Review Panels: Tobira; Grant/Research<br />

Support: Novartis, Noxxon; Speaking and Teaching: BMS, Gilead, Falk, MSD,<br />

Janssen, Abbvie<br />

Christian Trautwein - Grant/Research Support: BMS, Novartis, BMS, Novartis;<br />

Speaking and Teaching: Roche, BMS, Roche, BMS<br />

The following authors have nothing to disclose: Kai M. Schneider, Veerle Bieghs,<br />

Felix Heymann, Wei Hu, Daniela Dreymueller, Lijun Liao, Mick Frissen, Andreas<br />

Ludwig, Nikolaus Gassler, Oliver Pabst, Eicke Latz, Gernot Sellge, John Penders<br />

101<br />

Clonal analysis of the human liver reveals adaptable<br />

stem cell dynamics<br />

Malcolm Alison, Biancastella Cereser, Hemant Kocher, Stuart A.<br />

McDonald; Centre for Tumour Biology, Barts Cancer Institute, London,<br />

United Kingdom<br />

The study of cell lineages through heritable genetic lineage<br />

tracing is well established in experimental animals. However,<br />

using such techniques it is still unclear whether the liver conforms<br />

to a stem cell and lineage system, as seen for example<br />

in the bone marrow and gut. We have sought to resolve<br />

this question in the human liver using a variety of investigative<br />

techniques. The mitochondrial genome is highly prone<br />

to non-pathogenic mutations resulting in a deficiency in cytochrome<br />

c oxidase (CCO). Such mutations can be used as markers<br />

of clonal expansion and understanding the phylogenetics<br />

of these mutations will help in identifying stem cell derived<br />

clonal populations and their dynamics. Analysis of many hundreds<br />

of periportal and hepatic venous regions revealed that<br />

the majority of CCO-deficient cell patches abutted portal tracts,<br />

and when adjacent to hepatic veins the CCO-deficient patch<br />

typically extended from the portal tract. Furthermore, CCO-deficient<br />

patches were associated with trifurcating terminal portal<br />

tracts, but not with conducting portal tracts. Secondly we analyzed<br />

the methylation patterns of CpG islands in the promoter<br />

regions of non-expressed genes in CCO-deficient cell patches<br />

with respect to the portal tract-hepatic vein axis. Patterns from<br />

microdissected areas within CCO-deficient patches became<br />

diverse a short distance away from portal tracts, suggesting<br />

that most periportal and centrilobular hepatocytes are of similar<br />

mitotic age and only hepatocytes in the immediate periportal<br />

zone revealed patterns that were proportional to the size<br />

of the clone. Next generation sequencing (NGS) of human<br />

mtDNA was also employed to determine the clonal dynamics in<br />

CCO-deficient patches. Some mutations were common (clonal)<br />

to all cells within a patch, while some were confined (private)<br />

to cells at different points through the patch. Overall the methylation<br />

and NGS data suggest that CCO patches developed<br />

initially as a rapid expansion (perhaps through acute damage)<br />

which then ceased and allowed cell by cell diversity to accrue.<br />

Interestingly, the centrilobular regions revealed a higher mutation<br />

burden, indicative of these being older cells having had<br />

more time for subclonal expansion of mutations. Our <strong>studies</strong><br />

employing CCO activity, promoter methylation and NGS are<br />

consistent with the human liver being a lineage system, presumably<br />

emanating from periportal hepatic progenitor cells,<br />

the so-called streaming liver. However, this process appears to<br />

be very slow in normal homeostasis and rapid perhaps during<br />

periods of acute damage and suggests an adaptable stem cell<br />

organisation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Malcolm Alison, Biancastella<br />

Cereser, Hemant Kocher, Stuart A. McDonald


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 261A<br />

102<br />

MicroRNA-21 is a Potential Link between Non-alcoholic<br />

Fatty Liver Disease and Hepatocellular Carcinoma via<br />

Modulation of the HBP1-p53-Srebp1c Pathway<br />

Heng Wu, Guisheng Song; Medicine, University of Minnesota,<br />

Minneapolis, MN<br />

Background: Nonalcoholic fatty liver disease (NAFLD) is a<br />

major risk factor for hepatocellular carcinoma (HCC). However,<br />

the mechanistic pathways that link both disorders are<br />

essentially unknown. Our study was designed to investigate<br />

the role of microRNA-21 in the pathogenesis of NAFLD and its<br />

potential involvement in HCC. Methods: Wild-type mice maintained<br />

on a high fat diet (HFD) received tail-vein injections of<br />

microRNA-21-ASO (anti-sense oligonucleotide) or miR-21 mismatched<br />

ASO for 4 or 8 weeks. Livers were collected after that<br />

time period for lipid content and gene expression analysis. The<br />

role of microRNA-21 in carcinogenesis was analyzed by softagar<br />

colony formation, cell cycle analysis, and xenograft tumor<br />

assay. Results: The expression of microRNA-21 was increased<br />

in the livers of HFD-treated mice and human HepG2 cells incubated<br />

with fatty acid. MicroRNA-21 knockdown in those mice<br />

and HepG2 cells impaired lipid accumulation and growth of<br />

xenograft tumor. Further <strong>studies</strong> revealed that Hbp1 was a<br />

novel target of microRNA-21 and a transcriptional activator of<br />

p53. It is well-established that p53 is a tumor suppressor and<br />

an inhibitor of lipogenesis by inhibiting Srebp1c. As expected,<br />

microRNA-21 knockdown led to increased HBP1 and p53 and<br />

subsequently reduced lipogenesis and delayed G1/S transition,<br />

and the additional treatment of HBP1-siRNA antagonized<br />

the effect of microRNA-21-ASO, suggesting that HBP1<br />

mediated the inhibitory effects of microRNA-21-ASO on both<br />

hepatic lipid accumulation and hepatocarcinogenesis. Mechanistically,<br />

microRNA-21 knockdown induced p53 transcription,<br />

which subsequently reduced expression of genes controlling<br />

lipogenesis and cell cycle transition. In contrast, the opposite<br />

result was observed with overexpression of microRNA-21,<br />

which prevented p53 transcription. Conclusion: Our findings<br />

reveal a novel mechanism by which microRNA-21, in part,<br />

promotes hepatic lipid accumulation and cancer progression<br />

by interacting with the Hbp1-p53-Srebp1c pathway, and suggest<br />

the potential therapeutic value of microRNA-21-ASO for<br />

both disorders.<br />

Disclosures:<br />

The following authors have nothing to disclose: Heng Wu, Guisheng Song<br />

103<br />

Comparison And Outcomes Of 5% Albumin Vs 0.9%<br />

Normal Saline Fluid Resuscitation In Cirrhotics Presenting<br />

With Sepsis Induced Hypotension – A Randomized<br />

Controlled Trial - Fluid Resuscitation In Septic Shock In<br />

Cirrhosis (FRISC Protocol)<br />

Cyriac A. Philips 1 , Ashok K. Choudhury 1 , Amrish Sahney 1 , Rakhi<br />

Maiwall 1 , Lalita G. Mitra 2 , Shiv K. Sarin 1 ; 1 Hepatology and Transplant<br />

Medicine, Institute of Liver and Biliary Sciences, New Delhi,<br />

India; 2 Anesthesia and Critical Care, Institute of Liver and Biliary<br />

Sciences, New Delhi, India<br />

Background&Aims:Fluid resuscitation and choice of fluid for<br />

use in sepsis induced hypotension (SIH)is generally done as per<br />

Surviving Sepsis Guidelines (SSG). There is however, no data<br />

on fluid resuscitation protocols in cirrhosis patients. We investigated<br />

the efficacy, hemodynamic and perfusion effects of<br />

human albumin (HA) versus 0.9% normal saline(NS)in cirrhotics<br />

presenting with SIH. Patients&Methods: 308 patients of cirrhosis<br />

with SIH were randomized to receive either HA(5%,250ml<br />

bolus over 15 mins,n=154) or NS(30mL/kg over 30 mins);primary<br />

end-point (PE) being increase in mean arterial pressure<br />

(MAP)>65 mm Hg at 3 hr; secondary end-points of effects<br />

on heart rate(HR),lactate [delta lactate (dLAC), lactate clearance<br />

(cLAC)], urine output (UO) from baseline, at 1, 2 and<br />

3 hr and survival at 1 wk. Results:154 patients each, in HA<br />

(males,76%,mean age 49.7 yr) or NS group (79.2%,47 yr);(-<br />

MAP-53.8±6.2/54.4±6 mmHg,HR–93.6±22.6/93.4±16.8/<br />

bpm,LAC–6.94±2.2/6.83±2.6 mmol/L) and severity scores<br />

(CTP11.8±1.9/12.2±1.9,MELD 32±8.6/30±8.3,SOFA-<br />

9.99±2.5/10.4±2.8) had comparable baseline parameters.<br />

MAP>65mmHg at 1hr and sustenance of MAP at 3hr were in<br />

higher(p


262A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

idly. Our objective was to project the number of HCV patients<br />

needing treatment in 2015 and beyond, and the long-term<br />

health outcomes under different treatment penetration rates.<br />

Method: We used our previously published Hepatitis C Disease<br />

Burden Simulation model (HEP-SIM), which projected the<br />

changing prevalence of HCV in the United States. The HEP-SIM<br />

model was validated with NHANES <strong>studies</strong> and other published<br />

data. We simulated the current clinical management of<br />

HCV from 2001 onwards, which included risk-based screening<br />

until 2013 and addition of birth-cohort screening afterwards.<br />

We modeled antiviral treatment in different waves starting with<br />

peginterferon-ribavirin (PEG-RBV) until 2011, followed by the<br />

launch of boceprevir/telaprevir in 2012, sofosbuvir/simeprevir<br />

in 2014, and finally oral DAAs in 2015. We also implemented<br />

changes in insurance status because of the Affordable Care<br />

Act. We projected the number of patients needing treatment<br />

in 2015 and beyond under varying treatment capacity/penetration<br />

scenarios. We also projected patients’ fibrosis score,<br />

HCV awareness status, and access to insurance. Results: We<br />

estimated that in 2015, 2 million noninstitutionalized patients<br />

would be chronically infected and viremic. Among these<br />

patients, 1.1 million would be potential DAA candidates, i.e.,<br />

aware of their HCV status and insured. At the current annual<br />

treatment capacity of 200,000 patients, it will take at least 10<br />

years to treat the majority of the patients. By 2025, the number<br />

of treatment candidates would decline to fewer than 50,000<br />

patients. Even in the DAA era, 320,000 patients would die<br />

because of HCV, 32,000 patients would get liver transplants,<br />

202,000 would develop decompensated cirrhosis (DC) and<br />

156,000 would progress to hepatocellular carcinoma (HCC)<br />

by 2050. Doubling the annual treatment capacity could avoid<br />

8,000 deaths, 700 transplants, 7,000 DC and 4,000 HCC<br />

cases by 2050. Conclusions: Though the majority of patients<br />

aware of their HCV will be treated in the next 10 years, the<br />

HCV burden would still remain substantial unless aggressive<br />

screening and treatment policies are implemented. Increasing<br />

HCV treatment capacity is essential to decrease disease burden<br />

and improve health outcomes of HCV patients in the US, and<br />

decrease health resource utilization.<br />

Disclosures:<br />

Jagpreet Chhatwal - Consulting: Merck & Co., Inc., Gilead, Complete HEOR<br />

Solutions; Grant/Research Support: NIH/National Center for Advancing Translational<br />

Sciences<br />

The following authors have nothing to disclose: Xiaojie Wang, Fasiha Kanwal,<br />

Mina Kabiri, Turgay Ayer, Julie M. Donohue, Mark S. Roberts<br />

105<br />

An international, phase 2 randomized controlled trial of<br />

the dual PPAR α-δ agonist GFT505 in adult patients with<br />

NASH<br />

Vlad Ratziu 1 , Stephen A. Harrison 2 , Sven M. Francque 3 , Pierre<br />

Bedossa 4 , Lawrence Serfaty 5 , Manuel Romero-Gomez 6 , Paul<br />

Cales 7 , Manal F. Abdelmalek 8 , Stephen H. Caldwell 9 , Joost<br />

Drenth 10 , Quentin M. Anstee 11 , Dean W. Hum 12 , Rémy Hanf 12 ,<br />

Alice Roudot 12 , Sophie Megnien 12 , Bart Staels 13 , Arun J. Sanyal 14 ;<br />

1 Hepatology, Hopital Pitie Salpetriere, Paris, France; 2 Department<br />

of Medicine, Gastroenterology & Hepatology Service,, Brooke<br />

Army Medical Center, Fort Sam Houston, TX; 3 Department of Gastroenterology<br />

& Hepatology, Antwerp University Hospital, University<br />

of Antwerp, Antwerp, Belgium; 4 Department of Pathology,<br />

Hôpital Beaujon, University Paris-Denis Diderot, Clichy, France;<br />

5 Service d’Hépatologie,, Hôpital Saint-Antoine, APHP, UPMC Paris<br />

6, Paris, France; 6 Hospital Universitario de Valme, Unit for the<br />

Clinical Management of Digestive Diseases and CIBERehd, Sevilla,<br />

Spain; 7 Hepatology Department, University Hospital & LUNAM<br />

University, Angers, France., Angers, France; 8 Duke University,<br />

Duke university, Durham, NC; 9 University of Virginia, Gastroenterology<br />

& hepatology Division, Charlottesville, VA; 10 Radboud<br />

University Medical Center, Department of Gastroenterology and<br />

Hepatology, Nijmegen, Netherlands; 11 Institute of Cellular Medicine,,<br />

Faculty of Medical Sciences, Newcastle University, Newcastle<br />

upon Tyne, United Kingdom; 12 Genfit SA, Loos, France;<br />

13 Université Lille 2, INSERM U1011, European Genomic Institute<br />

for Diabetes (EGID), Institut Pasteur de Lille, Lille, France, Lille,<br />

France; 14 Virginia Commonwealth University, Richmond, VA<br />

Peroxisome proliferator-activated receptor α-δ dual agonists,<br />

such as GFT505, are a promising therapy for NASH as they<br />

improve hepatic insulin sensitivity, glucose homeostasis, lipid<br />

metabolism, and inflammation. Methods. In this randomized<br />

controlled trial (56 European and US centers) 274 patients (pts)<br />

(full analysis set, FAS) with histologically-defined non-cirrhotic<br />

NASH received GFT505 80 mg or 120 mg QD vs placebo (PLB)<br />

for one year. The primary outcome was resolution of NASH<br />

without worsening of fibrosis. Data were analyzed according<br />

to baseline severity (histological NAS score) and center effect.<br />

Biopsies were read by a single pathologist. Results. 237 pts<br />

had entry and end-of-treatment biopsies (ITT population). While<br />

the a priori primary endpoint did not meet significance, after<br />

controlling for baseline severity and center effect, pts in the<br />

120 mg arm had a 1.94 (CI 1.08-3.48, p=0.027) higher relative<br />

risk (RR) of achieving the primary end-point compared to<br />

PLB, while the RR was 1.68 (0.92-3.05, p=0.091) for the 80<br />

mg arm. Results were similar in the FAS where pts missing the<br />

second biopsy were counted as failures. In pts with moderate<br />

activity (NAS 4 or 5) the response rate was 27.5% in the 120<br />

mg arm vs. 19.5% for the PLB arm. In those with severe activity<br />

(NAS>5) it was 14.8% vs. 0%, respectively. In the 120 pts<br />

with NAS>4 from centers that recruited >1 patient/arm, the<br />

response rate was 29% and 5% in the 120 mg and PLB arms,<br />

respectively, p=0.01. A >2 point NAS reduction was obtained<br />

in 48% and 21% of patients respectively, p=0.01. Compared<br />

to PLB, the 120 mg arm improved ballooning (45% vs. 23%,<br />

p=0.02), inflammation (55% vs. 33%, p=0.05) and steatosis<br />

(35.5% vs. 18%, NS). In the 120 mg arm, resolution of NASH,<br />

resulted in a significant improvement in fibrosis (mean change<br />

-0.67 vs. +0.09 in non-responders, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 263A<br />

as HbA1c and FFA in diabetic pts, all on top of standard of<br />

care therapies. Tolerability was excellent without weight gain,<br />

cardiac events or safety signal. Conclusion. In NASH patients<br />

120 mg daily of GFT505 induced histological improvement<br />

and resolution of NASH, significantly more often than PLB.<br />

The excellent safety and tolerability and the improvement in<br />

cardiometabolic risk profile makes GFT505 an ideal drug candidate<br />

to be tested in phase 3 trials<br />

Disclosures:<br />

Vlad Ratziu - Advisory Committees or Review Panels: GalMed, Abbott, Genfit,<br />

Enterome, Gilead; Consulting: Tobira, Intercept, Exalenz, Sanofi-Synthelabo,<br />

Boehringer-Ingelheim<br />

Stephen A. Harrison - Advisory Committees or Review Panels: Merck, Nimbus<br />

Discovery, Fibrogen, RuiYi, CLDF; Consulting: NGM Biopharmaceuticals; Speaking<br />

and Teaching: Gilead, Abbvie, Janssen, CLDF<br />

Lawrence Serfaty - Board Membership: BMS, Gilead; Consulting: Merck; Speaking<br />

and Teaching: Roche, Janssen, Merck, Janssen, BMS, Gilead<br />

Manuel Romero-Gomez - Advisory Committees or Review Panels: Roche Farma,SA.,<br />

MSD, S.A., Janssen, S.A., Abbott, S.A.; Grant/Research Support: Ferrer,<br />

S.A.<br />

Paul Cales - Consulting: BioLiveScale<br />

Manal F. Abdelmalek - Consulting: Islet Sciences; Grant/Research Support:<br />

Tobira, Gilead Sciences, NIH/NIDDK, Synageva, Genfit Pharmaceuticals,<br />

Immuron, Galmed, TaiwanJ Pharma, Intercept, NGM Pharmaceuticals<br />

Stephen H. Caldwell - Advisory Committees or Review Panels: Vital Therapy;<br />

Grant/Research Support: Genfit, Gilead Sciences, Immuron, Hyperion, Immuron,<br />

NGM<br />

Quentin M. Anstee - Advisory Committees or Review Panels: Genfit, Intercept,<br />

Raptor; Grant/Research Support: GSK; Speaking and Teaching: Abbott Laboratories<br />

Dean W. Hum - Management Position: Genfit<br />

Rémy Hanf - Management Position: GENFIT<br />

Alice Roudot - Employment: GENFIT<br />

Sophie Megnien - Employment: GENFIT<br />

Bart Staels - Advisory Committees or Review Panels: MSD; Consulting: Genfit<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: Sven M. Francque, Pierre<br />

Bedossa, Joost Drenth<br />

106<br />

NGM282, A Novel Variant of FGF-19, Demonstrates<br />

Biologic Activity in Primary Biliary Cirrhosis Patients<br />

with an Incomplete Response to Ursodeoxycholic Acid:<br />

Results of a Phase 2 Multicenter, Randomized, Double<br />

Blinded, Placebo Controlled Trial<br />

Marlyn J. Mayo 3 , Alan J. Wigg 4 , Stuart K. Roberts 5 , Hays<br />

Arnold 19 , Tarek I. Hassanein 7 , Barbara A. Leggett 8 , John P. Bate 9 ,<br />

Martin Weltman 10 , Elizabeth J. Carey 11 , Andrew J. Muir 12 , Geoff<br />

McCaughan 14 , Steven J. Bollipo 15 , Stuart C. Gordon 16 , Peter W.<br />

Angus 13 , Stephen Riordan 17 , Mitchell L. Shiffman 18 , Elisa Young 6 ,<br />

Lei Ling 1 , Jian Luo 1 , Michael Elliott 1 , Stephen Rossi 1 , Alex M.<br />

DePaoli 1 , Alex J. Thompson 2 ; 1 NGM Biopharmaceuticals, Inc.,<br />

South San Francisco, CA; 2 St. Vincent’s Hospital, Melbourne, VIC,<br />

Australia; 3 UT Southwestern, Dallas, TX; 4 Flinders Medical Center,<br />

Adelaide, SA, Australia; 5 Alfred Hospital, Melbourne, VIC,<br />

Australia; 6 Novotech, Sydney, NSW, Australia; 7 Southern California<br />

Liver Research Center, Coronado, CA; 8 Royal Brisbane and<br />

Women’s Hospital, Brisbane, QLD, Australia; 9 Royal Adelaide<br />

Hospital, Adelaide, SA, Australia; 10 Nepean Hospital, Sydney,<br />

NSW, Australia; 11 Mayo Clinic - Arizona, Scottsdale, AZ; 12 Duke<br />

University, Durham, NC; 13 Austin Hospital, Heidelberg, VIC, Australia;<br />

14 University of Sydney, Sydney, NSW, Australia; 15 John<br />

Hunter Hospital, New Lambton, NSW, Australia; 16 Henry Ford<br />

Hospital, Detroit, MI; 17 Prince of Wales Hospital, Sydney, NSW,<br />

Australia; 18 Liver Institute of Virginia, Richmond, VA; 19 Digestive<br />

Research Center, San Antonio, TX<br />

Background: PBC patients with an incomplete biochemical<br />

response to ursodeoxycholic acid (UDCA) are at increased<br />

risk for disease progression and with limited treatment options.<br />

NGM282 is a novel engineered variant of FGF-19 that inhibits<br />

the CYP7A1-mediated bile acid (BA) synthesis in animals and<br />

healthy volunteers. The biologic activity was therefore evaluated<br />

in PBC patients. Methods: 45 subjects with a baseline<br />

(BL) alkaline phosphatase (ALP) >1.67xULN after 1yr of UDCA<br />

were randomized to NGM282 0.3 or 3mg vs placebo (PBO)<br />

as a daily SC injection for 28d. Change from BL ALP was the<br />

primary endpoint, with key secondary efficacy endpoints of<br />

liver chemistries and BA synthesis (7a-hydroxy-4-cholesten-3-<br />

one or C4). Pruritus was measured at all study visits with the<br />

5D Itch and Visual Analogue Score (VAS). Post-hoc sub-analyses<br />

evaluated ALP response by BL ALP< or >3xULN. Results:<br />

All study arms were balanced for typical PBC patient characteristics<br />

(female=91%, mean age=56y, mean BL ALP=297<br />

IU/L, mean UDCA dose=15mg/kg/d). Pruritus at BL was 54%,<br />

64% and 67% of PBO, 0.3mg and 3mg arms, respectively.<br />

NGM282-treated subjects had a significant, dose-dependent<br />

reduction in ALP vs PBO (p3xULN (Table 1). Significant reductions were also<br />

seen in liver chemistries (p90% supporting a potent biologic effect. NGM282<br />

was safe and well tolerated with no observed safety signals.<br />

Adverse events occurring >10% of both NGM282 treatment<br />

arms were diarrhea (PBO=6.7%, 0.3mg=21.4%, 3mg=25%),<br />

headache (PBO=6.7%, 0.3mg=14.3%, 3mg=25%) and nausea<br />

(PBO=6.7%, 0.3mg=14.3%, 3 mg=12.5%), the majority<br />

of which were mild. There was no clinically significant evidence<br />

of drug-induced pruritus by either VAS or 5D itch. Conclusions:<br />

These data demonstrate the biologic activity of NGM282 in<br />

PBC and support a therapeutic potential in other BA-mediated<br />

diseases. Studies are ongoing of longer duration, increased<br />

dose and BL predictors of response to identify optimal PBC<br />

patient populations for potential treatment with NGM282.


264A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Table 1 Mean change in liver and BA Parameters<br />

Disclosures:<br />

Marlyn J. Mayo - Grant/Research Support: Intercept, Salix, NGM, Lumena,<br />

Gilead<br />

Stuart K. Roberts - Board Membership: AbbVie, Gilead<br />

Tarek I. Hassanein - Advisory Committees or Review Panels: AbbVie, Bristol-Myers<br />

Squibb; Grant/Research Support: AbbVie Pharmaceuticals, Obalon, Bristol-Myers<br />

Squibb, Eiasi Pharmaceuticals, Gilead Sciences, Janssen R&D, Idenix<br />

Pharmaceuticals, Ikaria Therapeutics, Merck Sharp & Dohme, NGM BioPharmaceuticals,<br />

Ocera Therapeutics, Salix Pharmaceuticals, Sundise, TaiGen Biotechnology,<br />

Takeda Pharmaceuticals, Vital Therapies, Tobria; Speaking and<br />

Teaching: Baxter, Bristol-Myers Squibb, Gilead, Salix, AbbVie<br />

Barbara A. Leggett - Advisory Committees or Review Panels: MSD, MSD, MSD,<br />

MSD; Speaking and Teaching: Roche, Roche, Roche, Roche, Gilead<br />

Andrew J. Muir - Advisory Committees or Review Panels: BMS, Gilead, Janssen,<br />

Merck; Consulting: Theravance; Grant/Research Support: Abbvie, Abbvie, BMS,<br />

Gilead, Janssen, Merck, Achillion, Lumena<br />

Geoff McCaughan - Advisory Committees or Review Panels: Gilead<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

Peter W. Angus - Advisory Committees or Review Panels: Gilead Sciences, BMS;<br />

Grant/Research Support: Gilead sciences<br />

Mitchell L. Shiffman - Advisory Committees or Review Panels: Merck, Gilead,<br />

Boehringer-Ingelheim, Bristol-Myers-Squibb, Abbvie, Janssen, Acchillion; Consulting:<br />

Roche/Genentech; Grant/Research Support: Merck, Gilead, Boehringer-Ingelheim,<br />

Bristol-Myers-Squibb, Abbvie, Beckman-Coulter, Achillion, Lumena,<br />

Intercept, Novartis, Gen-Probe; Speaking and Teaching: Roche/Genentech,<br />

Merck, Gilead, Abbvie, Janssen, Bayer<br />

Lei Ling - Employment: NGM Biopharmaceuticals, Inc.<br />

Jian Luo - Employment: NGM Biopharmaceuticals<br />

Michael Elliott - Employment: NGM; Stock Shareholder: NGM<br />

Stephen Rossi - Employment: NGM Biopharmaceuticals, Inc; Stock Shareholder:<br />

NGM Biopharmaceuticals, Gilead Sciences<br />

Alex M. DePaoli - Employment: NGM Biopharmaceuticals<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

The following authors have nothing to disclose: Alan J. Wigg, Hays Arnold, John<br />

P. Bate, Martin Weltman, Elizabeth J. Carey, Steven J. Bollipo, Stephen Riordan,<br />

Elisa Young<br />

in non-diabetic NASH derived from the PIVENS vitamin E and<br />

placebo groups and from non-diabetics in the FLINT placebo<br />

group. Methods: Two efficacy measures from FLINT were<br />

applied to our pooled data: histologic improvement, defined<br />

as ≥ 2 point improvement in NAS with no worsening of fibrosis<br />

or NASH resolution. Safety estimates paralleled those used in<br />

FLINT and included incidence of cardiac events and changes in<br />

lipid levels. Logistic regression models were used to summarize<br />

the odds ratio (OR) effects, confidence limits, and p-values on<br />

the pooled vitamin E treatment vs no vitamin E treatment efficacy<br />

estimates; Fisher’s exact test was used to assess cardiac<br />

events. Results: A total of 250 patients were randomized to<br />

vitamin E (n=80) or placebo (n=72) in PIVENS or to placebo<br />

(n=98) in FLINT and had both baseline and end-treatment liver<br />

biopsies; 53 had diabetes (21%) and 197 were non-diabetic<br />

(79%); 105 (42%) received vitamin E and 145 (58%) did not<br />

in the PIVENS or FLINT trials. Vitamin E use was associated<br />

with histologic improvement in diabetic (OR 4.4, 95% CI 1.1,<br />

18.0, p=0.04) and non-diabetic patients (OR 3.1, 95% CI<br />

1.7, 5.8, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 265A<br />

decompensation, reduced risk for hepatocellular carcinoma,<br />

and reduced liver-related mortality. We propose that cure of<br />

CHC (defined as SVR of at least 12 weeks) can also lead to<br />

regression of advanced fibrosis and/or cirrhosis and that Fibroscan<br />

elastography may be used to confirm reversal of fibrosis.<br />

METHODS: We conducted a retrospective chart review to<br />

identify patients treated for CHC with concomitant advanced<br />

fibrosis or cirrhosis based on clinical features, Fibroscan scores<br />

and/or biopsy. Advanced fibrosis was defined as FibroScan<br />

score 11-13.9Kpa and cirrhosis was defined as Fibroscan<br />

score > 14KPa. This was followed by prospective and retrospective<br />

Fibroscan and clinical data collection at 6 month intervals,<br />

up to 14 years. RESULTS: 201 subjects were enrolled, 58<br />

are currently eligible for analysis with at least 2 measurements<br />

of fibrosis. Mean age was 60.7 years, 70.7% were male, and<br />

most were Non-Hispanic (81%) white (91.4%). At baseline<br />

25.9% had hypertension, 12.1 % had diabetes, 12.1% had<br />

hyperlipidemia and the average BMI was 27.7 (SD 5). Of<br />

the 34 subjects who had cirrhosis at baseline, 18 (52.9%)<br />

demonstrated improvement by Fibroscan, with a median time<br />

to improvement of 2.8 years (IQR 1.0-3.5). Improvement was<br />

defined as a change in stage of fibrosis rather than a change in<br />

Fibroscan score. Of the 17 subjects who had advanced fibrosis,<br />

16 (66.7%) demonstrated improvement, with a median time of<br />

2.0 years (IQR 1.2-2.5). Gender, BMI, age, individual medical<br />

problems and HCV genotype were not found to be associated<br />

with improvement or worsening of baseline liver disease. ALT,<br />

albumin and INR changes did not correlate with fibrosis regression.<br />

However, platelet count change (increased counts in subjects<br />

who demonstrated improved liver disease) approached<br />

statistical significance with a p=0.06 in subjects with cirrhosis.<br />

CONCLUSION: The majority of our subjects, 34/58 (58.6%),<br />

with advanced fibrosis or cirrhosis demonstrated improvement<br />

by Fibroscan scores after a median follow-up (after treatment<br />

of CHC) of 2.5 years (range: 0.5-14 years, IQR: 1- 3.5 years).<br />

We have yet been not able to establish factors that may predict<br />

improvement of cirrhosis or advanced fibrosis. Most enrolled<br />

patients are still awaiting follow-up Fibroscan test results anticipated<br />

to be completed in 2015.<br />

Disclosures:<br />

Catherine T. Frenette - Speaking and Teaching: Bayer, Salix, Gilead; Stock<br />

Shareholder: Gilead<br />

Paul J. Pockros - Advisory Committees or Review Panels: Janssen, Merck, BMS,<br />

Gilead, AbbVie; Consulting: Lumena, Beckman Coulter; Grant/Research Support:<br />

Intercept, Janssen, BMS, Gilead, Lumena, Beckman Coulter, AbbVie, RMS,<br />

Merck; Speaking and Teaching: AbbVie, Janssen, Gilead<br />

The following authors have nothing to disclose: Ana Maria Crissien, William B.<br />

Minteer, Jason J. Pan<br />

109<br />

Circulating extracellular vesicles and their microRNA<br />

cargos are potential novel biomarkers with a functional<br />

role of miR-122 in monocyte activation in alcoholic hepatitis<br />

Banishree Saha, Fatemah Momen-Heravi, Shashi Bala, Karen<br />

Kodys, Gyongyi Szabo; Medicine, UMASS Med School, Worcester,<br />

MA<br />

Purpose: Alcohol and its metabolites induce hepatocyte damage<br />

and recruitment of inflammatory monocytes/macrophages<br />

and neutrophils leading to alcoholic hepatitis. Currently there<br />

is no reliable biomarker of alcoholic hepatitis. Extracellular<br />

vesicles (EVs) found in the circulation carry mRNA, microRNA<br />

(miRs) and proteins. EVs can serve as biomarkers and mediate<br />

intercellular communication. miR-122 is abundantly expressed<br />

in hepatocytes, not in immune cells and increased levels of circulating<br />

miR-122 were found in liver injury. We hypothesized<br />

that EV-associated miRs can serve as biomarkers and modulate<br />

intercellular signaling between hepatocytes and immune<br />

cells in alcoholic hepatitis. Methods: EVs were isolated from<br />

chronic alcohol-fed (5 weeks of Lieber DeCarli diet) or pair-fed<br />

mice sera and from serum of patients with alcoholic hepatitis.<br />

EVs were characterized by transmission electron microscopy,<br />

western blot, nanoparticle tracking analysis system and<br />

miRNA analysis. Results: The total number of circulating EVs<br />

was significantly increased in alcohol-fed mice as compared<br />

to control mice. Exosomes (40-150nm) represented most of<br />

the EVs (~80%). MicroRNA array of circulating EVs revealed<br />

a significant increase of 7 inflammatory miRs including: miR-<br />

192, 122, 30a, 744, 1246, 30b and miR-130a in alcohol-fed<br />

mice compared to controls. The ROC analyses indicated excellent<br />

diagnostic value of miR-192, 122, and 30a to identify<br />

alcohol-induced liver injury. In patients with acute alcoholic<br />

hepatitis, we found a significant increase in the number of<br />

circulating EVs compared to normal controls and miR-192 and<br />

miR-30a were significantly increased in the EVs from alcoholic<br />

hepatitis patients. Serum miR-122 was increased after alcohol<br />

binge drinking. In the liver, miR-122 is abundantly expressed<br />

in hepatocytes and monocytes/macrophages have low levels.<br />

In vitro experiments revealed that exosomes derived from<br />

ethanol-treated human hepatocytes were taken up by monocytes<br />

and transferred mature miR-122 into monocytes. This horizontally<br />

transferred miR-122 inhibited the hemeoxygenase-1<br />

expression, a target of miR-122 and sensitized monocytes to<br />

LPS stimulation to increase production of pro-inflammatory cytokines,<br />

TNF-α and IL-1β; all of these effects were inhibited by<br />

exosome-mediated delivery of a miR-122 inhibitor in monocytes.<br />

Conclusion: Elevated levels of EVs and their miR signature<br />

could serve as biomarkers of alcoholic hepatitis. This study<br />

reveals a novel EV-mediated mechanism of alcohol-induced<br />

communication between hepatocytes and monocytes by transferring<br />

hepatocyte-derived miR-122 that reprograms monocytes<br />

promoting inflammation in alcoholic hepatitis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Banishree Saha, Fatemah<br />

Momen-Heravi, Shashi Bala, Karen Kodys, Gyongyi Szabo<br />

110<br />

Fat-specific Protein 27/CIDEC Promotes Alcoholic Steatohepatitis<br />

in Mice and Humans<br />

Ming-Jiang Xu 1 , Yan Cai 1 , Hua Wang 1 , José T. Altamirano 2 ,<br />

Gemma Odena 3 , Frank J. Gonzalez 4 , Ramon Bataller 3 , Bin Gao 1 ;<br />

1 NIAAA, NIH, Rockville, MD; 2 Vall d’Hebron Institut de Recerca,<br />

Barcelona, Spain; 3 University of North Carolina, Chapel Hill, NC;<br />

4 National Institutes of Health, Bethesda, MD<br />

Objectives: Alcoholic steatohepatitis (ASH) is the progressive<br />

form of alcoholic liver disease that leads to cirrhosis and<br />

hepatocellular carcinoma. There is an urgent need to identify<br />

molecular drivers in order to develop targeted therapies.<br />

Here the functions of mouse fat-specific protein 27 (Fsp27)/<br />

human cell death activator CIDEC (the human homologue of<br />

Fsp27) were investigated. Methods and results: We developed<br />

a mouse model with chronic (8 weeks)-plus-binge ethanol feeding,<br />

which mimics the drinking patterns of alcoholic hepatitis<br />

(AH) patients, produced severe ASH and mild fibrosis. Microarray<br />

analyses revealed that mouse Fsp27/human CIDEC gene<br />

was increased in this animal model and human ASH samples.<br />

Fsp27 is expressed at high levels in adipose tissues but at<br />

very low levels in normal liver. Chronic-plus-binge ethanol feeding<br />

markedly upregulated hepatic Fsp27 mRNA and protein<br />

expression. Silencing the Fsp27 gene by shRNA or genetic<br />

disruption ameliorated chronic-plus-binge ethanol-induced


266A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ASH. Inhibition of peroxisome proliferator-activated receptor<br />

g (PPARg) or cyclic-AMP-responsive-element binding protein H<br />

(CREBH) attenuated chronic-plus-binge ethanol-induced elevation<br />

of Fsp27a and FSP27b mRNA, respectively, and subsequently<br />

ameliorated liver injury. Overexpression of Fsp27 and<br />

ethanol exposure synergistically induced mitochondrial reactive<br />

oxygen species production and hepatocyte injury in vivo and<br />

in vitro, which is likely owing to the mitochondrial location of<br />

FSP27 leading to the decreased mitochondrial complex I activity.<br />

Finally, hepatic expression of CIDEC mRNA was elevated<br />

and positively correlated with hepatic steatosis, disease severity,<br />

and mortality in AH patients. Conclusion: FSP27/CIDEC<br />

gene product promotes ASH in chronic-plus-binge ethanol-fed<br />

mice and in human AH. Targeting CIDEC gene is likely to be a<br />

novel therapeutic targets for the treatment of ASH.<br />

overgrowth was similar to WT mice, alcohol-fed Reg3b and<br />

Reg3g deficient mice showed a significantly higher number of<br />

mucosa-associated bacteria in the small intestine as compared<br />

with their respective WT littermates. This was accompanied<br />

by a significantly increased number of bacteria translocating<br />

through intestinal epithelial cells and more positive mesenteric<br />

lymph node cultures in alcohol-fed Reg3b -/- and Reg3g -/- mice.<br />

Interestingly, absence of Reg3b or Reg3g did not affect the<br />

intestinal paracellular barrier function as determined by systemic<br />

LPS level and fecal albumin. Most importantly, Reg3b<br />

and Reg3g deficient mice showed more alcoholic liver disease<br />

as assessed by higher plasma ALT levels, increased hepatic<br />

triglycerides and inflammation. To corroborate our data, we<br />

generated transgenic mice overexpressing Reg3g under the<br />

control of the intestine specific villin promoter. Consistent with<br />

our results, Reg3g transgenic mice showed significantly less<br />

mucosa-associated bacteria than WT littermates after alcohol<br />

feeding. Overexpression of intestinal Reg3g suppressed<br />

translocation of bacteria through intestinal epithelial cells to<br />

mesenteric lymph nodes, while paracellular permeability and<br />

systemic LPS levels were not changed after alcohol feeding.<br />

Reg3g transgenic mice were protected from alcohol-induced<br />

liver injury, steatosis and inflammation. Conclusion: Intestinal<br />

Reg3b and Reg3g prevent bacterial colonization of epithelial<br />

surfaces, which reduces translocation of viable bacteria and<br />

prevents alcoholic liver disease. Reg3 proteins are novel and<br />

promising targets in the treatment of alcohol-induced liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Lirui Wang, Peng Chen, Lora V.<br />

Hooper, Bernd Schnabl<br />

Disclosures:<br />

Ramon Bataller - Advisory Committees or Review Panels: Sandhill; Consulting:<br />

VTI, Oncozyme Pharma<br />

The following authors have nothing to disclose: Ming-Jiang Xu, Yan Cai, Hua<br />

Wang, José T. Altamirano, Gemma Odena, Frank J. Gonzalez, Bin Gao<br />

111<br />

Antimicrobial proteins Reg3b and Reg3g protect mice<br />

from alcoholic liver disease by preventing bacterial<br />

translocation<br />

Lirui Wang 1 , Peng Chen 1 , Lora V. Hooper 2 , Bernd Schnabl 1 ;<br />

1 University of California San Diego, La Jolla, CA; 2 Department of<br />

Immunology, The University of Texas Southwestern Medical Center,<br />

Dallas, TX<br />

Background: Antimicrobial proteins are secreted by intestinal<br />

epithelial cells and Paneth cells. They represent the first<br />

line of defense against pathogens and maintain homeostasis<br />

with commensal bacteria. We have previously shown that the<br />

expression of the C-type lectin regenerating islet derived-3<br />

(Reg3) is suppressed in the small intestine after chronic alcohol<br />

use in mice and humans, yet functional consequences of<br />

lower Reg3 expression on the intestinal microbiota, bacterial<br />

translocation and alcoholic liver disease are unknown. The<br />

aim of our study was to investigate the role of Reg3b and<br />

Reg3g in chronic alcoholic liver disease. Methods and Results:<br />

A Lieber-DeCarli model was used to induce intestinal dysbiosis,<br />

bacterial translocation and liver disease in mice. After alcohol<br />

feeding for 8 weeks, wild-type (WT) littermate mice showed<br />

bacterial overgrowth of the luminal and the mucosa-associated<br />

microbiota in the small intestine. While luminal bacterial<br />

112<br />

Hepatocytes from mice on intragastric feeding model of<br />

alcoholic steatohepatitis release extracellular vesicles<br />

with specific microRNA cargo that modulate hepatic<br />

stellate cell and macrophage phenotype<br />

Akiko Eguchi 1 , Jihoon Kim 1 , Lucila Ohno-Machado 1 , Hidekazu<br />

Tsukamoto 2 , Ariel E. Feldstein 1 ; 1 UCSD, La Jolla, CA; 2 USC, Los<br />

Angeles, CA<br />

Liver inflammation and fibrosis are key histological features<br />

associated with prognosis in patients with alcoholic steatohepatitis<br />

(ASH). Extracellular vesicles (EVs) are released during cell<br />

stress or demise, can contain a barcode of the cell of origin<br />

including specific microRNAs and are growingly recognized<br />

as key cell-to-cell communicators. Here we tested the hypothesis<br />

that during ASH development, hepatocyte damage release EVs<br />

with a microRNA signature that can fuse with hepatic stellate<br />

cells (HSC) and Kupffer cells (KC) to regulate their phenotype.<br />

Methods: C57/B6 mice were placed on intragastric feeding<br />

model of continuous ethanol infusion or control diet for 4 weeks<br />

to reproduce a physiologically relevant model of ASH. The<br />

extent of steatosis, inflammation, and fibrosis was assessed<br />

by histological and molecular analyses on liver specimens.<br />

Isolated hepatocytes (HC) or KC from ASH or control mice<br />

were incubated in medium-EV free serum and HC- or KC-derived<br />

EVs were isolated by ultracentrifugation from culture<br />

medium. A complete characterization of EVs was performed<br />

by FACS, electron microscopy, dynamic light scattering. HC-EV<br />

biological function was investigated in isolated primary HSCs<br />

or KCs derived from C57/B6 mice by cell morphology or<br />

qPCR. Comprehensive encapsulated miRNA in HC-EVs were<br />

assessed via transcriptome using illumina miRNA-seq. Results:<br />

We observed highly significant differences in the levels of HCor<br />

KC-EVs between control group and ASH (1.3x10 6 HC-EV/


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 267A<br />

ml of medium in ASH versus 1.4x10 5 HC-EV/ml of medium<br />

in control, p


268A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ing significant reduction in certain long- (hexadecanoic and<br />

heptadecanoic), medium (hexanoic and octanoic), and short<br />

(butanoic) free fatty acids. Remarkably, the levels of octanoic<br />

acid (known to have antimicrobial properties) were dramatically<br />

reduced (~48 fold) in mice fed USF+EtOH compared to<br />

SF+EtOH fed animals. A decline in certain fecal amino acids<br />

(e.g. serine and glycine) was observed in USF+EtOH fed animals.<br />

Conclusions: These data support an important role of<br />

dietary lipids in ALD pathogenesis, and provide insight into<br />

mechanisms of ALD development. A diet enriched in USF not<br />

only enhanced alcohol-induced liver injury, but also caused<br />

major fecal metagenomic and metabolomic changes that may<br />

play an etiologic role in observed liver injury. Characterization<br />

of both microbiota composition and function is an important<br />

approach to investigate host–microbial interaction. Our data<br />

suggest that dietary lipids can potentially serve as interventions<br />

for the prevention/treatment of ALD.<br />

Disclosures:<br />

Shirish Barve - Speaking and Teaching: Abbott<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Irina Kirpich, Wenke Feng, Xinmin<br />

Yin, Xiaoli Wei, Xiang Zhang<br />

115<br />

PEGylated TRAIL treatment ameliorates liver cirrhosis in<br />

rats by targeting activated hepatic stellate cells<br />

Ogyi Park 1,2 , Yumin Oh 1,2 , Magdalena Swierczewska 1,2 , Jong<br />

Sung Park 1,2 , Justin Hanes 2 , Martin Pomper 1 , Bin Gao 3 , Seulki<br />

Lee 1,2 ; 1 The Russell H. Morgan Department of Radiology and<br />

Radiological Sciences, Johns Hopkins University, Baltimore, MD;<br />

2 The Center for Nanomedicine at the Wilmer Eye Institute, Johns<br />

Hopkins University School of Medicine, Baltimore, MD; 3 Laboratory<br />

of Liver Diseases, National Institute of Alcohol Abuse and<br />

Alcoholism, National Institutes of Health, Bethesda, MD<br />

Background: Liver fibrosis is a common outcome of chronic<br />

liver disease and leads to liver failure and cancer. Still, no targeted<br />

anti-fibrotic therapy exists at this time. Activated hepatic<br />

stellate cells (aHSCs) are the originators of liver fibrosis; therefore,<br />

eradication of aHSCs is a logical strategy to stop and/<br />

or reverse liver fibrosis. However, there are no effective strategies<br />

to specifically deplete apoptosis-resistant aHSCs during<br />

fibrosis without systemic toxicity. TRAIL is known to selectively<br />

induce apoptosis in cancer cells or transformed cells by binding<br />

to its death receptors (DRs) while sparing normal tissue.<br />

But because recombinant TRAIL suffers from a short half-life<br />

and low potency, it failed to show efficacy in cancer clinical<br />

trials. Here we introduce a longer-acting, more stable form<br />

of TRAIL that shows striking anti-fibrotic activity in rat fibrosis<br />

and cirrhosis models with no detected hepatotoxicity. Methods:<br />

We developed PEGylated TRAIL (TRAIL PEG<br />

) by stabilizing a<br />

potent homotrimer TRAIL of iLZ-TRAIL with a 5kDa PEG. In in<br />

vivo <strong>studies</strong>, SD rats were induced liver fibrosis and cirrhosis<br />

by CCl 4<br />

and treated with 4 mg/kg of TRAIL PEG<br />

daily for 10<br />

days for fibrosis and two weeks for cirrhosis. In in vitro <strong>studies</strong>,<br />

primary human HSCs were culture-activated and TRAIL PEG<br />

effects were investigated on quiescent and activated HSCs.<br />

Primary human hepatocytes were used to determine hepatotoxicity<br />

of TRAIL PEG<br />

. Results: Intravenously injected TRAIL PEG<br />

has a markedly extended half-life in non-human primates and<br />

no toxicity in primary human hepatocytes. We found DRs are<br />

upregulated in human cirrhotic livers because of aHSCs. In<br />

vitro primary human aHSCs, but not quiescent HSCs, become<br />

sensitive to TRAIL PEG<br />

-induced apoptosis via DR- and DISC-activated<br />

caspase-8-dependent mechanisms. Intravenous TRAIL PEG<br />

directly induced apoptosis of aHSCs and ameliorated CCl 4<br />

-induced<br />

fibrosis/cirrhosis in rats by simultaneously down-regulating<br />

multiple key fibrogenic molecules in vivo. TRAIL PEG<br />

also<br />

caused selective toxicity to hepatocellular carcinoma (HCC)<br />

cells. Conclusions: TRAIL-based therapies could serve as firstin-class<br />

therapeutics for liver fibrosis/cirrhosis, cirrhosis-associated<br />

HCC and possibly other fibrotic diseases. Our results not<br />

only introduce the unique activity that TRAIL-based compounds<br />

have in liver fibrosis but also warrant clinical translation of<br />

TRAIL PEG<br />

for severe liver fibrosis therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ogyi Park, Yumin Oh, Magdalena<br />

Swierczewska, Jong Sung Park, Justin Hanes, Martin Pomper, Bin Gao,<br />

Seulki Lee<br />

116<br />

Statins, ACE inhibitors and ARB use is associated with<br />

reduced mortality and morbidity in chronic liver diseases<br />

– a nationwide cohort study in Sweden<br />

Knut Stokkeland 1,2 , Christine Takami Lageborn 3 , Anders Ekbom 4 ,<br />

Jonas Höijer 5 , Matteo Bottai 5 , Per Stål 6 , Karin Söderberg Löfdal 7 ;<br />

1 Department of Medicine, Visby Hospital, Visby, Sweden; 2 Karolinska<br />

Institutet, Department of Medicine, Gastroenterology and<br />

Hepatology, Stockholm, Sweden; 3 Karolinska Institutet, Stockholm,<br />

Sweden; 4 Department of Medical Epidemiology and Biostatistics,<br />

Karolinska Institutet, Stockholm, Sweden; 5 Unit of Biostatistics,<br />

IMM, Karolinska Institutet, Stockholm, Sweden; 6 Division of Hepatology,<br />

Karolinska Hospital, Stockholm, Sweden; 7 Division of Clinical<br />

Pharmacology, Department of Laboratory Medicine, Karolinska<br />

Institutet, Stockholm, Sweden<br />

Aim: To explore whether current medication in patients with<br />

chronic liver diseases affect overall mortality, liver-related mortality<br />

and liver-related morbidity. Methods: We performed a<br />

register-based cohort study in Sweden with patients with a<br />

first-time diagnosis of chronic liver disease between 2005 and<br />

2012. We studied the use of statins, angiotensin converting<br />

enzyme inhibitors (ACEi), angiotensin receptor blockers (ARB),<br />

acetylsalicylic acid (ASA), non-steroidal anti-inflammatory<br />

drugs (NSAID), selective seretonin re-uptake inhibitors (SSRI)<br />

and antibiotics. We measured total mortality, liver-specific<br />

mortality and liver-specific morbidity. We used register data<br />

from the Patient Register, the Prescribed Drug Register and the<br />

Death Certificate Register. Liver-specific morbidity and mortality<br />

included all liver diseases, esophageal varices and liver<br />

cancer. Results: We analyzed 70 546 patients who had been<br />

seen in hospital out-patient services or hospitalized with chronic<br />

liver disease for the first time between 2005 and 2012. We<br />

found a reduction in mortality risk for statin users (n= 8 688)<br />

with a hazard ratio range between 0.42 (95% CI: 0.28-0.63)<br />

for patients with autoimmune hepatitis and 0.91 (0.63-1.32)<br />

for primary biliary cirrhosis with a similar risk reduction for<br />

liver-specific mortality and liver-related morbidity. There was<br />

a significant mortality reduction for patients exposed to ARB<br />

(n= 6204) between 0.62 (0.53-0.72) in alcoholic liver disease<br />

and 0.91 (0.64-1.32) in primary biliary cirrhosis and a similar<br />

reduction for patients exposed to ACEi (n=9 714). There was<br />

a significant reduction of liver morbidity in patients exposed to<br />

ASA (n= 9 768) for all chronic liver diseases, even among users<br />

of NSAID (n= 14 119). An increased risk for all-cause mortality<br />

was seen for almost all liver diseases in patients exposed to<br />

SSRI (n= 11 048), with a range between 1.25 (1.16-1.35) and<br />

1.61 (1.30-1.98), and a reduction in risk of liver-specific morbidity<br />

for almost all diagnoses. There was a reduction in risk for<br />

liver morbidity for patients exposed to antibiotics (n= 27 756)<br />

between 0.55 (0.44-0.70) and 0.77 (0.73-0.80). Conclusion:


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 269A<br />

Several frequently used drug classes in patients with chronic<br />

liver disease are associated with reduced all-cause mortality,<br />

and liver-specific morbidity and mortality. Statin, ACEi and<br />

ARB use was associated with lower risk of total mortality, liver-specific<br />

mortality and liver-specific morbidity. ASA use was<br />

associated with reduced liver-specific mortality, and NSAID,<br />

SSRI and antibiotics exposure was associated with a reduction<br />

in risk for liver morbidity.<br />

Disclosures:<br />

Anders Ekbom - Advisory Committees or Review Panels: Centocor, Schering-Plough,<br />

Centocor, Schering-Plough, Centocor, Schering-Plough, Centocor,<br />

Schering-Plough; Speaking and Teaching: Astra Zeneca, Astra Zeneca, Astra<br />

Zeneca, Astra Zeneca<br />

The following authors have nothing to disclose: Knut Stokkeland, Christine<br />

Takami Lageborn, Jonas Höijer, Matteo Bottai, Per Stål, Karin Söderberg Löfdal<br />

that becomes significantly more accurate than its 2 composite<br />

tests, and solves discrepancies between these tests. The<br />

combined test that associates elastography and a blood test,<br />

used with its detailed fibrosis classification, should become the<br />

reference.<br />

Disclosures:<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

Vincent Leroy - Board Membership: Abbvie, BMS, Gilead; Consulting: Janssen,<br />

MSD; Speaking and Teaching: Abbvie, BMS, Gilead, Janssen, MSD<br />

Isabelle Fouchard-Hubert - Speaking and Teaching: JANSSEN, BMS, GILEAD,<br />

ABBVIE<br />

Paul Cales - Consulting: BioLiveScale<br />

The following authors have nothing to disclose: Jerome Boursier, Oberti Frederic<br />

117<br />

Evaluation of EASL recommendation for the non-invasive<br />

diagnosis of liver fibrosis in chronic hepatitis C<br />

Jerome Boursier 1,3 , Victor de Ledinghen 2 , Vincent Leroy 4 , Oberti<br />

Frederic 1,3 , Isabelle Fouchard-Hubert 1,3 , Paul Cales 1,3 ; 1 Hepato-gastroenterology<br />

Department, University Hospital, Angers,<br />

France; 2 Hepatology Department, University Hospital, Bordeaux,<br />

France; 3 HIFIH Laboratory EA3859, University, Angers, France;<br />

4 Hepato-Gastroenterology Department, University Hospital, Grenoble,<br />

France<br />

Introduction. The EASL has proposed the first recommendations<br />

about non-invasive testing for the evaluation of liver disease<br />

severity. For chronic hepatitis C (CHC), it is proposed to associate<br />

transient elastography (TE) with a blood test (BT), and to<br />

accept the diagnosis if both are in agreement. We aimed to<br />

validate this rule and to improve it by using a combined fibrosis<br />

test that simultaneously associates both BT and TE (BET). Methods.<br />

679 CHC patients with liver biopsy, TE, and a patented BT<br />

were included. The primary outcome was severe fibrosis (Metavir<br />

F≥3, prevalence: 31.9%). Diagnostic cut-off for this target<br />

was evaluated according to the maximum Youden index or the<br />

threshold between the 2 classes F2±1 and F3±1 of a detailed<br />

fibrosis classification including 6 fibrosis classes. Results.<br />

AUROC was: BT: 0.829, TE: 0.846, BET: 0.882 (p


270A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Hans Huber - Consulting: MiNA Therapeutics; Stock Shareholder: Merck & Co<br />

Robert Habib - Board Membership: MiNA Therapeutics Ltd; Management Position:<br />

MiNA Therapeutics Ltd; Stock Shareholder: MiNA Therapeutics Ltd<br />

Pål Sætrom - Patent Held/Filed: MiNA Therapeutics, Ltd; Stock Shareholder:<br />

MiNA Therapeutics, Ltd<br />

Nagy Habib - Board Membership: EMcision Limited, Omnicyte Limited, Apterna<br />

Limited, MiNA Therapeutics Ltd; Patent Held/Filed: EMcision Limited, Omnicyte<br />

Limited; Stock Shareholder: EMcision Limited, Omnicyte Limited, Apterna Limited,<br />

MiNA Therapeutics Ltd<br />

The following authors have nothing to disclose: Kai-Wen Huang, Anjaneyulu<br />

Muragundla, Aravindakshan Jayaprakash, Prashant Vadnal, John Rossi<br />

119<br />

Fibrosis is not just fibrosis - Extracellular matrix turnover<br />

is markedly different in two types of viral hepatitis, suggesting<br />

different fibrotic representations<br />

Mette J. Nielsen 1,2 , Morten A. Karsdal 1 , Konstantin Kazankov 3 ,<br />

Aleksander Krag 2 , Diana J. Leeming 1 , Jacob George 4 , Henning<br />

Gronbaek 3 , Detlef Schuppan 5,6 ; 1 Nordic Bioscience A/S, Herlev,<br />

Denmark; 2 Department of Gastroenterology and Hepatology,<br />

Odense University Hospital, University of Southern Denmark,<br />

Odense, Denmark; 3 Department of Medicine V, Aarhus University<br />

Hospital, Aarhus, Denmark; 4 Storr Liver Unit, Westmead Millennium<br />

Institute, Westmead Hospital and University of Sydney,<br />

Sydney, NSW, Australia; 5 Institute of Translational Immunology,<br />

University Medical Center, Mainz, Germany; 6 Division of Gastroenterology,<br />

Beth Israel Deaconess Medical Center, Harvard Medical<br />

School, Boston, MA<br />

Background and aim: Fibrosis is the result of a dysregulated<br />

tissue remodelling leading to excessive and abnormal accumulation<br />

of extracellular matrix (ECM). The developmental pattern<br />

of fibrosis depends on the underlying aetiology causing the<br />

fibrosis such as; portal-portal, portal-central, or central-central<br />

septa, albeit the molecular composition and turnover remain<br />

to be investigated. In this study we investigated the level of<br />

ECM remodelling in two seemingly similar types of viral hepatitis,<br />

namely hepatitis B and C. Methods: In a cross-sectional<br />

study design, specific protein fragments of matrix metalloprotease<br />

degraded type I, III, IV and VI collagen (C1M, C3M,<br />

C4M, C6M) and type III and IV collagen formation (Pro-C3<br />

and P4NP7S) were assessed in plasma from 403 chronic<br />

hepatitis C patients and 197 chronic hepatitis B patients by<br />

specific ELISAs. The grade of inflammation and fibrosis was<br />

scored according to the Metavir Activity and Fibrosis Scoring<br />

systems based on liver biopsy. Results: Plasma levels of<br />

P4NP7S, C3M, C4M, and C6M were significantly elevated<br />

in HBV compared to HCV (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 271A<br />

121<br />

Objective Determination of Hepatitis B Phenotype using<br />

Biochemical and Serological Markers: Immune Tolerant,<br />

Immune Active, Inactive Carrier and Indeterminant Status<br />

Adrian M. Di Bisceglie 1 , Manuel Lombardero 2 , Jeffrey Teckman 1 ,<br />

Lewis R. Roberts 3 , Harry L. Janssen 4 , Steven H. Belle 2 , Jay H. Hoofnagle<br />

5 ; 1 Saint Louis University, St. Louis, MO; 2 University of Pittsburgh,<br />

Pittsburgh, PA; 3 Mayo Clinic, Rochester, MN; 4 University of<br />

Toronto, Toronto, ON, Canada; 5 NIH, Bethesda, MD<br />

Background: HBV infection is frequent world-wide and may<br />

result in progressive liver disease or HCC. Effective therapies<br />

are available but treatment is usually reserved for patients (pts)<br />

with selected biochemical and serological profiles (phenotypes).<br />

Currently, criteria for defining HBV phenotype are not<br />

standardized and are based upon expert opinion rather than<br />

medical evidence. Aim: To determine the distribution of phenotypes<br />

in a large cohort of North American pts with chronic<br />

HBV infection using standardized definitions and calculation.<br />

Methods: Epidemiologic, demographic, biochemical and virologic<br />

features of pts enrolled into the HBRN Cohort Study at 19<br />

US and 1 Canadian site were analyzed, assigning pts into 1<br />

of 4 phenotypes: immune tolerant (IT), chronic hepatitis B with<br />

(CHB e+) or without HBeAg (CHB e-) or inactive carrier (IC)<br />

based upon baseline HBV DNA and ALTs. Cut-off HBV DNA<br />

values used to define phenotypes were 10 4 IU/ml for e-negative<br />

and 10 5 for e-positive pts. ALT cut-offs were analyzed using<br />

either fixed values (30 U/L for men, 20 U/L for women) or<br />

using each local lab ULN. Pts not fitting into a phenotype were<br />

designated “Indeterminant”. Independently, local investigators<br />

assigned a phenotype at baseline based on available laboratory<br />

values and clinical impression. Results: 1398 adults (51%<br />

male, 71% Asian) had data needed to determine phenotype.<br />

The table shows the distribution of phenotypes calculated by<br />

two methods of assessing ALT and physician assessment. Only<br />

moderate agreement was found between clinician-determined<br />

and calculated phenotype using fixed ALTs (Kappa 0.39, 95%<br />

CI 0.36-0.42). Of note, only 13% of pts were designated Indeterminant<br />

by clinicians compared to 20-40% by computer calculation.<br />

Using the Fixed ALT groups for comparison, IT was<br />

most distinct, being generally younger, more frequently Asian<br />

and female. The most frequent clinical phenotype identified<br />

was Indeterminant, the majority of whom (83%) had elevated<br />

ALT values despite low levels of HBV DNA, not apparently<br />

associated with higher BMI, alcohol consumption or HBV genotype.<br />

Clinician estimates often differed from the calculated<br />

phenotype. Conclusions: The clinical significance of HBV phenotypes<br />

using these standardized definitions requires further<br />

study based on long term follow up of these patient cohorts but<br />

there is clearly a need to re-examine categorization of pts with<br />

chronic HBV infection to more accurately predict their outcomes<br />

and need for therapy.<br />

Disclosures:<br />

Adrian M. Di Bisceglie - Advisory Committees or Review Panels: Gilead, AbbVie,<br />

Novartis, Bayer, BTG; Grant/Research Support: Gilead, AbbVie<br />

Jeffrey Teckman - Consulting: Dicerna, Isis Pharmaceuticals, Vertex, Proteostasis,<br />

Genkyotex, The Alpha-1 Project, Retrophin, RxCelerate, Velgene; Grant/<br />

Research Support: Alnylam, Arrowhead, Alpha-1 Foundation, Gilead<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Manuel Lombardero, Steven H.<br />

Belle, Jay H. Hoofnagle<br />

122<br />

Length of antiviral therapy and cirrhosis are key factors<br />

in the development of hepatocellular carcinoma among<br />

U.S. veterans with chronic hepatitis B<br />

Gina Choi, Marina Serper, David E. Kaplan, Kimberly A. Forde;<br />

Gastroenterology, University of Pennsylvania, Philadelphia, PA<br />

Background: The risk of hepatocellular carcinoma (HCC) and<br />

the effect of anti-viral therapy in U.S. populations with chronic<br />

hepatitis B (HBV) is poorly defined. Aim: To examine the incidence<br />

and impact of anti-viral therapy on the development of<br />

HCC among a national cohort of veterans with chronic HBV.<br />

Methods: A retrospective cohort study using the VA Corporate<br />

Data Warehouse was performed between 1999-2013. The<br />

primary exposure variable was antiviral therapy, defined as<br />

≥6 months of therapy with oral nucleosides. Additional covariates<br />

such as demographics and clinical variables such as cirrhosis,<br />

alcohol abuse, dyslipidemia, hepatitis C (HCV), and<br />

HIV were obtained. Multivariable Cox proportional hazard<br />

models were used to evaluate associations between exposures<br />

and the primary outcome of HCC. Results: A total of 26,704<br />

veterans had a HBsAg+ result; N=9001 were confirmed to<br />

have chronic HBV. The median follow-up time was 7.6 years<br />

(IQR 4.2–10.9). Patients were predominantly male (96%),<br />

white (45%) and middle aged (M=51, SD 11). A total of 20%<br />

had HCV, 7% had HIV, and 23% had alcohol abuse. The<br />

prevalence of dyslipidemia was 52%; 13% had cirrhosis. The<br />

overall incidence of HCC was 5.6 per 1000 person-years;<br />

26 per 1000 person-years with cirrhosis and 1.9 per 1000<br />

person-years without cirrhosis. The median exposure time of<br />

antiviral therapy was 3.6 years (IQR 1.4-7.1). A total of 38%<br />

(N=3333) received antiviral therapy for ≥6 months; 23%<br />

received ≥3 years and 16% received therapy for ≥5 years.<br />

In multivariable models adjusted for age, race, cirrhosis, alcohol<br />

abuse, and dyslipidemia, antiviral therapy ≥5 years was<br />

associated with decreased incidence of HCC (HR 0.77, 95%<br />

CI 0.60-0.98, p=0.03), whereas ≥4 years of antiviral therapy<br />

was not protective (HR 1.04, CI 1.03-1.05). Antiviral therapy<br />

for ≥5years was associated with a decreased incidence of<br />

HCC among cirrhotics (HR 0.56, CI 0.41-0.78, p


272A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

123<br />

Serum Alanine Aminotransferase (ALT) and Hepatitis<br />

B Virus (HBV) DNA Flares In Pregnant and Postpartum<br />

Women With Chronic Hepatitis B (CHB)<br />

Christine Y. Chang 1 , Natali Aziz 2 , Huy N. Trinh 3 , Mindie H.<br />

Nguyen 1 ; 1 Division of Gastroenterology and Hepatology, Stanford<br />

University Medical Center, Palo Alto, CA; 2 Obsterics and Gynecology,<br />

Stanford University School of Medicine, Stanford, CA; 3 San<br />

Jose Gastroenterology, San Jose, CA<br />

Background: During pregnancy, alterations in the immune system<br />

allow the mother to tolerate the semiallogenic fetus, but<br />

their effects on CHB activity are poorly understood. Methods:<br />

To examine flares in HBV DNA (≥1 log increase from baseline)<br />

and/or ALT (≥2xULN or baseline) during pregnancy and 6<br />

months postpartum, we performed a retrospective study of 88<br />

pregnancies in 74 CHB expectant mothers who were not on<br />

anti-HBV therapy for ≥1 year prior to pregnancy at 2 U.S.<br />

centers. Results: Most were Asian (93%), with a mean age<br />

of 32±4, median gravida of 2 (IQR: 1-3), and median parity<br />

of 0 (IQR: 0-1). Mean log HBV DNA was 2.0±0.8 IU/mL at<br />

baseline in low HBV DNA mothers (≤2000 IU/mL, n=33), and<br />

6.1±2.0 in high HBV DNA mothers (>2000 IU/mL, n=55). At<br />

baseline, low HBV DNA mothers had lower ALT (≤x2ULN: 82%<br />

vs. 47%, p=0.001) and were more likely to be HBeAg- (90%<br />

vs. 44%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 273A<br />

Disclosures:<br />

The following authors have nothing to disclose: Prabhu P. Gounder, Lisa Bulkow,<br />

Mary Snowball, Susan Negus, Philip R. Spradling, Brian J. McMahon<br />

126<br />

Relationship between hepatocellular carcinoma development<br />

and serum viral markers in chronic hepatitis B<br />

patients who achieved undetectable serum HBV DNA<br />

while on long-term nucleoside analogue therapy<br />

Ka-Shing Cheung 1 , Wai-Kay Seto 1 , Danny Wong 2 , Ching-Lung<br />

Lai 1 , Man-Fung Yuen 1 ; 1 Medicine, Queen Mary Hospital, The University<br />

of Hong Kong, Hong Kong, Hong Kong; 2 Queen Mary<br />

Hospital, Hong Kong, Hong Kong<br />

Background: Hepatitis B surface antigen (HBsAg) and hepatitis<br />

B core-related antigen (HBcrAg) are risk factors for the<br />

development of hepatocellular carcinoma (HCC). Linearized<br />

HBsAg (HQ-HBsAg) is a more sensitive assay to measure<br />

HBsAg level. However, their roles in predicting HCC development<br />

are unknown when there is profound viral suppression<br />

by nucleos(t)ide analogues (NA). Methods: Seventy-six chronic<br />

hepatitis B (CHB) patients who developed HCC in 2007-2014<br />

despite undetectable serum HBV DNA level after at least oneyear<br />

NA therapy were recruited. 152 CHB patients without<br />

HCC were recruited as controls. They were matched by age,<br />

gender, HBeAg status, cirrhosis status, undetectable serum HBV<br />

DNA and duration of NA therapy with the 76 HCC patients in<br />

a 2:1 ratio. Serum HBsAg, HQ-HBsAg and HBcrAg levels were<br />

analysed and compared between the two groups. Results: A<br />

statistically significant difference in the HBcrAg level was noted<br />

between the HCC group and non-HCC groups (10.2 and 1.7<br />

kU/mL respectively, p=0.005), but was not observed in HBsAg<br />

or HQ-HBsAg levels. The area under receiver operating characteristic<br />

curve (AUROC) was 0.61 (95% CI: 0.54-0.69) with a<br />

negative predictive value (NPV) of 77.0% when a cutoff value<br />

of HBcrAg level ≥7.8 kU/mL was used. The odds ratio of HCC<br />

development was 3.27 (95% CI: 1.84-5.80). In the subgroup<br />

analysis for non-cirrhotic patients, a statistically significant<br />

difference in the HBcrAg level was noted between the HCC<br />

group and non-HCC groups (10.2 and 1.0 kU/mL respectively,<br />

p=0.001). The AUROC was 0.70 (95% CI: 0.58-0.81) with a<br />

NPV of 80.6% when a cutoff value of HBcrAg level ≥7.9 kU/<br />

mL was used. The odds ratio of HCC development was 5.95<br />

(95% CI: 2.35-15.07). Conclusion: HBcrAg (but not HBsAg or<br />

HQ-HBsAg) can predict HCC risk in CHB patients who achieve<br />

undetectable serum HBV DNA while receiving NA therapy.<br />

Future prospective <strong>studies</strong> are warranted to further define the<br />

role of HBcrAg level in this group of patients.<br />

Disclosures:<br />

Wai-Kay Seto - Advisory Committees or Review Panels: Gilead Science; Speaking<br />

and Teaching: Bristol-Myers Squibb<br />

Ching-Lung Lai - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead Sciences Inc; Consulting: Bristol-Myers Squibb, Gilead Sciences, Inc;<br />

Speaking and Teaching: Bristol-Myers Squibb, Gilead Sciences, Inc<br />

Man-Fung Yuen - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Bristol-Myers Squibb, Pfizer, GlaxoSmithKline, Bristol-Myers Squibb, Pfizer,<br />

GlaxoSmithKline, Bristol-Myers Squibb, Pfizer, GlaxoSmithKline, Bristol-Myers<br />

Squibb, Pfizer; Grant/Research Support: Roche, Bristol-Myers Squibb,<br />

GlaxoSmithKline, Gilead Science, Roche, Bristol-Myers Squibb, GlaxoSmith-<br />

Kline, Gilead Science, Roche, Bristol-Myers Squibb, GlaxoSmithKline, Gilead<br />

Science, Roche, Bristol-Myers Squibb, GlaxoSmithKline, Gilead Science<br />

The following authors have nothing to disclose: Ka-Shing Cheung, Danny Wong<br />

127<br />

A New Clinically-Based Staging System for Distal Cholangiocarcinoma<br />

Maria E. Lozada 1 , Jonggi Choi 1 , Roongruedee Chaiteerakij 1,2 ,<br />

Ju Dong Yang 1 , Nasra H. Giama 1 , Steven Alberts 3 , Gregory J.<br />

Gores 1 , Lewis R. Roberts 1 ; 1 Division of Gastroenterology and<br />

Hepatology, Mayo Clinic College of Medicine, and Mayo Clinic<br />

Cancer Center, Rochester, MN; 2 Department of Medicine, Faculty<br />

of Medicine, Chulalongkorn University and King Chulalongkorn<br />

Memorial Hospital, Thai Red Cross Society, Bangkok, Thailand;<br />

3 Division of Medical Oncology, Mayo Clinic College of Medicine,<br />

and Mayo Clinic Cancer Center, Rochester, MN<br />

Background: Clinical predictors of survival for each subtype of<br />

cholangiocarcinoma, intrahepatic, perihilar and distal (dCCA)<br />

are limited. Currently available staging systems can be used to<br />

predict prognosis for resectable dCCA but are not relevant to<br />

patients with unresectable disease. Based on these limitations,<br />

we aimed to construct a clinical staging system that stratifies<br />

dCCA patients. Methods: We identified 170 treatment-naïve<br />

dCCA patients seen at Mayo Clinic, Rochester from January<br />

1, 2000 through May 31, 2014. Data were retrospectively<br />

abstracted from the electronic medical record. Overall median<br />

survival (MS), the primary endpoint, was estimated using the<br />

Kaplan-Meier method and compared using the log-rank test.<br />

Cox-proportional hazards analysis was used to identify predictors<br />

of survival. Performance of the staging system was assessed<br />

using concordance index. Results: Of the 170 dCCA patients,<br />

93 (58%) were male, the mean age was 67 years, and 19<br />

(12%) had primary sclerosing cholangitis. The MS of the entire<br />

cohort was 17.0 months (95% confidence interval [CI]: 14.0-<br />

20.5). Baseline ECOG performance status, post-stenting CA<br />

19-9 and disease extent were significantly associated with survival<br />

and further incorporated into a 4-tier staging system. A<br />

total of 143 patients were classified into the new staging system.<br />

Overall MS for patients in Stage I, Stage II, Stage III and<br />

Stage IV were 46.8, 14.3, 8.1, and 3.4 months, with hazard<br />

ratios (95% CI) of 1.0 (ref.), 1.9 (1.0-3.4), 4.1 (2.3-7.0), and<br />

9.7 (5.1-18.5), respectively. Concordance index for the new<br />

staging system was 0.73. Conclusion: We have developed a<br />

new staging system based on non-operative clinical variables<br />

that classifies dCCA patients into 4 distinct stages with homogeneous<br />

survival. Given the small cohort, validation to confirm<br />

these findings is crucial. However, this staging system could<br />

potentially be a key prognostic tool for better stratification of<br />

patients enrolled in clinical trials of novel therapies for dCCA.<br />

Disclosures:<br />

Gregory J. Gores - Advisory Committees or Review Panels: Conatus<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics


274A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Maria E. Lozada, Jonggi Choi,<br />

Roongruedee Chaiteerakij, Ju Dong Yang, Nasra H. Giama, Steven Alberts<br />

128<br />

Trends in the burden of cirrhosis and hepatocellular carcinoma<br />

by underlying liver disease in US Veterans from<br />

2001-2013<br />

Lauren A. Beste 2 , Steven Leipertz 2 , Pamela Green 2 , Jason A. Dominitz<br />

2,1 , David Ross 3 , George N. Ioannou 2,1 ; 1 University of Washington,<br />

Seattle, WA; 2 Veterans Affairs Puget Sound Healthcare<br />

System, Seattle, WA; 3 Office of Public Health, Department of Veterans<br />

Affairs, Washington, D.C., DC<br />

Background & Aims Cirrhosis and hepatocellular carcinoma<br />

(HCC) are predicted to increase in the U.S. but the accuracy<br />

of prior forecasts and the contributions from various etiologies<br />

remain unclear. We aimed to determine the burden of<br />

cirrhosis and HCC according to underlying etiology from<br />

2001-2013. Methods We developed a national retrospective<br />

cohort of Veterans Affairs (VA) patients diagnosed with cirrhosis<br />

(n=129,998) or HCC (n=21,326) from 2001-2013.<br />

We used laboratory results, ICD-9 codes, and biometrics to<br />

identify underlying etiologies. Results In 2013, VA provided<br />

care to 5,720,614 individuals, of whom 60,553 (1.06%) had<br />

cirrhosis and 7,670 (0.13%) had HCC. Hepatitis C virus (HCV)<br />

was present in an increasing proportion of cirrhosis and HCC<br />

between 2001-2013, reaching 48% of cirrhosis cases and<br />

deaths and 67% of HCC cases and deaths by 2013. Cirrhosis<br />

prevalence nearly doubled from 2001-2013 (664 to 1058<br />

per 100,000 enrollees), driven by HCV and nonalcoholic fatty<br />

liver disease (NAFLD). Cirrhosis incidence ranged from 159<br />

to 193 per 100,000 patient-years. Deaths in patients with<br />

cirrhosis increased from 83 to 126 per 100,000 patient-years,<br />

largely driven by HCV. HCC incidence increased 2.5-fold from<br />

17 to 45 per 100,000 patient-years. HCC mortality tripled<br />

from 13 to 37 per 100,000 patient-years, driven overwhelmingly<br />

by HCV with much smaller contributions from NAFLD and<br />

alcoholic liver disease. Conclusions Cirrhosis prevalence and<br />

mortality and HCC incidence and mortality increased from<br />

2001-2013 driven by HCV with a smaller contribution from<br />

NAFLD. If current trends continue, cirrhosis prevalence will<br />

peak in 2021. Healthcare systems will need to accommodate<br />

rising numbers of patients with cirrhosis and HCC.<br />

INCIDENCE OF HCC IN THE VETERANS AFFAIRS HEALTHCARE<br />

SYSTEM 2002-2012, BY UNDERLYING LIVER DISEASE<br />

Disclosures:<br />

Jason A. Dominitz - Employment: Department of Veterans Affairs; Grant/Research<br />

Support: Gilead Pharmaceuticals<br />

The following authors have nothing to disclose: Lauren A. Beste, Steven Leipertz,<br />

Pamela Green, David Ross, George N. Ioannou<br />

129<br />

Multi-platform analysis of Telomerase Reverse Transcriptase<br />

(TERT) gene alterations in Hepatocellular Carcinoma<br />

(HCC)<br />

Renumathy Dhanasekaran 1 , Kyle Covington 2 , Jennifer Watt 2 ,<br />

Andrew D. Cherniack 3 , Daniel R. O’Brien 4 , Ju-Seog Lee 5 , Chad<br />

Creighton 2 , Donna M. Munzy 2 , Lewis R. Roberts 1 , David A.<br />

Wheeler 2 ; 1 Gastroenterology and Hepatology, Mayo Clinic, Rochester,<br />

MN; 2 Baylor College of Medicine, Houston, TX; 3 Broad<br />

Institute of MIT and Harvard, Cambridge, MA; 4 Mayo Clinic, Rochester,<br />

MN; 5 The University of Texas MD Anderson Cancer Center,<br />

Houston, TX<br />

Background: Telomerases are enzymatic protein complexes<br />

which maintain telomere length and play an important role in<br />

cancer cell immortalization. Mechanisms of TERT gene activation<br />

and clinical associations of TERT gene alterations in<br />

HCC are not completely understood. Methods: We used data<br />

from TCGA analysis of HCC primary tumor from 192 patients.<br />

Germline DNA from blood or benign surrounding liver was<br />

used as reference. Mutational, copy number, mRNAseq and<br />

viral genomic integration analyses were performed. Results:<br />

The TERT gene was found to be significantly overexpressed<br />

in HCC tumor tissue when compared to benign surrounding<br />

liver (pA (C228T) and 1,295,250 G>A (C250T)) were found.<br />

The majority of tumors had the C288T mutation (92.1%) and<br />

six (7.9%) tumors had the C250T mutation. Patients with TERT<br />

promoter mutation were older (p=0.03) and predominantly<br />

male (p=0.02). Patients with TERT promoter mutation were<br />

more likely to have hepatitis C than no mutation (33% vs 10%;<br />

p=0.02). In contrast, patients with TERT promoter mutation<br />

were less likely to have hepatitis B than no mutation (14% vs.<br />

27%; p=0.05). Also, patients with TERT promoter mutation<br />

had worse survival than patients without the mutation (3 year<br />

survival 59% vs. 39%; p=0.02). Further, we report here the first<br />

case of a germline mutation in the TERT promoter in HCC in a<br />

30 year old male who did not have any risk factors for HCC.<br />

Copy number analyses showed TERT gene amplification in<br />

7% of tumor samples; TERT was the fourth most common gene<br />

found to be amplified in HCC. There was significant correlation<br />

between TERT gene amplification and TERT mRNA overexpression.<br />

In the subset of patients with HBV infection, the integration<br />

of HBV viral genome into TERT promoter region was identified<br />

in 13.6% (6/44) of patients and this was the most common site<br />

of recurrent HBV integrations. Conclusion: This is the first comprehensive<br />

multi-platform analysis of the various mechanisms of<br />

TERT gene alterations in HCC and their association with patient<br />

outcomes. TERT promoter mutations were found to be the most<br />

common somatic mutation in HCC and were associated with<br />

poor survival. We identify TERT promoter somatic mutations,<br />

TERT gene amplification and HBV integrations into the TERT<br />

promoter region as various mechanisms of TERT gene alterations.<br />

Also, we report here the first case of germline mutation<br />

of the TERT promoter gene in HCC, possibly resulting in an<br />

inherited predisposition to HCC.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 275A<br />

Disclosures:<br />

Andrew D. Cherniack - Grant/Research Support: Bayer AG<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Renumathy Dhanasekaran, Kyle<br />

Covington, Jennifer Watt, Daniel R. O’Brien, Ju-Seog Lee, Chad Creighton,<br />

Donna M. Munzy, David A. Wheeler<br />

130<br />

Targeting the TGF-β Pathway Increases Patient Survival<br />

for Hepatocellular Cancer<br />

Jian Chen 1 , Jiun-Sheng Chen 1 , Wilma Jogunoori 1 , Young Jin Gi 1 ,<br />

Yun Seong Jeong 1 , Xiaoping Su 1 , Asif Rashid 1 , Bibhuti Mishra 2 ,<br />

Jon White 2 , Milind Javle 1 , Marta L. Davila 1 , John R. Stroehlein 1 ,<br />

Xuemei Wang 1 , Jeffrey S. Morris 1 , Rehan Akbani 1 , Lopa Mishra 1 ;<br />

1 The University of Texas MD Anderson Cancer Center, Houston,<br />

TX; 2 Institute of Clinical Research, Veterans Affairs Medical Center,<br />

Washington DC, DC<br />

Liver cancers that include hepatocellular cancer (HCC) and<br />

intrahepatic cholangiocarcinoma are among the most challenging<br />

and lethal malignancies with a degree of complexity seldom<br />

seen in other cancers. Patients with HCC have a poor survival<br />

of 9-11 months. However, key driver pathways and identifying<br />

specific populations responsive to targeted therapeutics have<br />

remained elusive. Recent clinical <strong>studies</strong> show that targeting<br />

TGF-β in patients with high TGF-β level improves survival up<br />

to 21 months. Therefore, we hypothesized that genomic, transcriptomic<br />

profiles coupled with mouse mutants and functional<br />

analyses may identify specific patients with HCC with a high<br />

response rate to targeting TGF-β and present an attractive new<br />

treatment strategy. Methods and Results: (1) We analyzed the<br />

transcriptome of 488 hepatocellular cancers and screened for<br />

mutations in the TGF-β pathway in 202 HCCs from The Cancer<br />

Genome Atlas (TCGA). We found aberrant TGF-β superfamily<br />

in 72% of HCC with mutations in 38%, the greatest number<br />

being for the Smad adaptor β2SP (6%). (2) Increased levels of<br />

TGF-β genes were designated as an “activated” signature that<br />

is associated with hepatic fibrosis. Decreased levels of TGF-β<br />

genes were defined as an “inactivated” signature, which was<br />

associated with the loss of TGF-β tumor suppressor function.<br />

HCCs characterized by the “inactivated” TGF-β signature were<br />

associated with a significantly poorer survival particularly in<br />

early stage HCCs, compared to HCCs with the “activated”<br />

TGF-β signature (p=0.0027). (3) The TGF-β pathway is statistically<br />

significantly associated with DNA repair genes. (4)<br />

We performed (80x) whole-genome sequence analysis of eight<br />

additional HCCs to define the role of TGF-β in their development<br />

and characterize a potential novel “driver mutations” in<br />

HCV- and alcohol-associated HCC. (5) Further functional analyses<br />

were performed in mouse mutants in the TGF-β pathway<br />

that spontaneously develop HCC, patient derived xenografts,<br />

and human cell lines. Increased levels of TGF-β associated-E3<br />

ligases, including KEAP1, and PJA1 were found in 20% of<br />

HCCs. Targeting these E3 ligases revealed a strong response<br />

using patient derived xenografts. Conclusions: The molecular<br />

signatures of the TGF-β pathway we characterize here appear<br />

to have prognostic significance. The additional association<br />

with the DNA repair pathway supports new approaches to<br />

developing biomarkers. TGF-β and TGF-β-associated-E3 ligases<br />

are activated in distinct groups of liver cancer patients. It may<br />

be possible to identify subsets of patients, defined by the activation<br />

or inactivation of TGF-β signaling, who may respond more<br />

effectively to specific treatments.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jian Chen, Jiun-Sheng Chen,<br />

Wilma Jogunoori, Young Jin Gi, Yun Seong Jeong, Xiaoping Su, Asif Rashid,<br />

Bibhuti Mishra, Jon White, Milind Javle, Marta L. Davila, John R. Stroehlein,<br />

Xuemei Wang, Jeffrey S. Morris, Rehan Akbani, Lopa Mishra<br />

131<br />

A Study for the Risk Factors of Hepatocellular Carcinoma<br />

in Cirrhotic Patients with Chronic Hepatitis B<br />

Yuanqing Zhang 1 , Lijun Peng 1,2 , Yirong Cao 1 , Zhiping Zeng 1 ,<br />

Yujing Wu 1 , Hong Shi 1 , Shiyao Chen 1 , Jiyao Wang 1 , Scott L.<br />

Friedman 1,3 , John J. Sninsky 1,4 , Jinsheng Guo 1 ; 1 Division of Digestive<br />

Diseases, Zhong Shan Hospital, Fu Dan University, Shanghai,<br />

China; 2 Department of Gastroenterology, Linyi People’s Hospital,<br />

Linyi, China; 3 Division of Liver Diseases,, Mount Sinai Hospital,<br />

Icahn School of Medicine at Mount Sinai, New York, NY; 4 Celera/Quest<br />

Diagnostics, Alameda, CA<br />

Background and Aims: Chronic hepatitis B virus infection and<br />

cirrhosis are important risk factors for hepatocellular carcinoma<br />

(HCC), however the HCC risk can vary widely among individuals.<br />

The aim of this study was to identify the clinical factors<br />

and host genetic single nucleotide polymorphisms (SNPs) that<br />

are associated with HCC risk in cirrhotic patients with CHB.<br />

Methods: Data from 949 Chinese Han patients with chronic<br />

HBV infection were analyzed by single and multivariate logistic<br />

regression analysis in cirrhotic patients without (n = 281) and<br />

with HCC (n = 434) to identify the clinical factors associated<br />

with HCC risk. In a parallel study, DNA was extracted from<br />

879 chronic HBV patients including non-cirrhotic HBV carriers<br />

and CHB patients (n = 234), cirrhotic patients without (n =<br />

257) and with HCC (n = 388) for genotyping of ten candidate<br />

SNPs using polymerase chain reaction-ligase detection<br />

reaction method. The correlations between the candidate SNPs<br />

and HCC risk were analyzed by chi-square test followed by<br />

logistic regression analysis. Results: 1. Ineffective antiviral treatment<br />

(OR=5.2), fatty liver (OR=3.4), family history of HCC<br />

(OR=2.3), drinking history (OR=2.2), and age≥50 (OR=1.8)<br />

were independent risk factors for HCC in cirrhotic patients with<br />

CHB. Longer HBV infection increased the HCC risk whereas<br />

sustained virologic suppression significantly lowered HCC<br />

risk compared to those without enough response or untreated<br />

patients, especially in patients with decompensated cirrhosis.<br />

Even after achieving viral suppression, 36.5% (62/170) of cirrhotic<br />

patients with CHB still developed HCC. Fatty liver, family<br />

history of HCC and HBV infection were significantly associated<br />

with the HCC development in these patients. 2. After adjusted<br />

for influencing factors, TLR4 rs11536889 was found to be a<br />

protective factor (OR for CC vs. GG+GC or GG =0.4 and<br />

0.6, respectively) whereas SPP1 rs2853744 (OR for TT vs. GG<br />

and GT+TT vs. GG =1.9 and 3.1, respectively) and AP3S2<br />

rs2290351 (OR for GA+AA or GA vs. GG =2.5 and 2.9,<br />

respectively) were risk factors of HCC in cirrhotic patients with<br />

CHB. In cirrhotic patients with CHB and drinking history, MLEC<br />

rs7976497 (OR for TT vs. CC=2.8, CT+TT vs. CC=3.6) and<br />

SOCS3 rs4969168 (OR for GG vs. AA=2.4) were found to be<br />

risk factors of HCC. Conclusion: Ineffective antiviral treatment,<br />

fatty liver, family history of HCC, drinking history and age≥50<br />

are risk factors for HCC. Sustained suppression of HBV does<br />

not remove the risk of HCC. Specific host genetic factors may<br />

impact on HCC development in cirrhotic patients with CHB<br />

including a protective SNP in TLR4, and risk SNPs in SPP1,<br />

AP3S2, MLEC, and SOCS3.


276A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

John J. Sninsky - Consulting: Roche Diagnostics; Employment: CareDx<br />

The following authors have nothing to disclose: Yuanqing Zhang, Lijun Peng,<br />

Yirong Cao, Zhiping Zeng, Yujing Wu, Hong Shi, Shiyao Chen, Jiyao Wang,<br />

Jinsheng Guo<br />

132<br />

Diabetes, HBV infection and smoking are independent<br />

risk factors for developing hepatocellular carcinoma<br />

on non-fibrotic liver in the NoFLIC French multicenter<br />

case-control study.<br />

Jean-Frédéric Blanc 1 , Adelaide Doussau 2 , Marie-Quitterie Picat 2 ,<br />

Aline Maillard 2 , Caroline Bouyssou 1 , Charlotte E. Costentin 3 , Olivier<br />

Rosmorduc 4 , Isabelle Rosa 5 , Luc Letenneur 2 , Karen Leffondre 6 ,<br />

Jessica Zucman-Rossi 6 , Marianne Ziol 7 ; 1 Hepatology and Digestive<br />

Oncology Unit, CHU bordeaux, Bordeaux, France; 2 ISPED,<br />

Bordeaux, France; 3 Hepatology, Hopital Henri Mondor, Creteil,<br />

France; 4 Heptology, Hopital saint-Antoine, Paris, France; 5 Hepatology,<br />

CHI Créteil, Créteil, France; 6 INSERM U1162, Paris, France;<br />

7 Pathologie, Hopital Jean Verdier, Bondy, France<br />

risk factors for developing nfHCC. Conclusion NoFLIC is the<br />

first case-control study dedicated to the study of factors associated<br />

with the occurrence of HCC in non-fibrotic liver (F0/F1).<br />

The first analysis shows a significant association between metabolic<br />

syndrome, diabetes, HBV infection and smoking with the<br />

occurrence of HCC in non-fibrotic liver suggesting an important<br />

role of these factors in liver carcinogenesis.<br />

Disclosures:<br />

Jean-Frédéric Blanc - Advisory Committees or Review Panels: Lilly / Imclone,<br />

Merck Serono, Sanofi; Speaking and Teaching: Bayer, BMS, Roche, Abbvie<br />

Olivier Rosmorduc - Advisory Committees or Review Panels: Syrtex, IPSEN;<br />

Speaking and Teaching: Bayer<br />

Jessica Zucman-Rossi - Advisory Committees or Review Panels: Astellas, Celgene;<br />

Consulting: Pfizer; Grant/Research Support: Integragen; Speaking and Teaching:<br />

Bayer, Lilly<br />

Marianne Ziol - Grant/Research Support: Echosens<br />

The following authors have nothing to disclose: Adelaide Doussau, Marie-Quitterie<br />

Picat, Aline Maillard, Caroline Bouyssou, Charlotte E. Costentin, Isabelle<br />

Rosa, Luc Letenneur, Karen Leffondre<br />

Introduction The occurrence of HCC in non-fibrotic liver (nfHCC)<br />

is a rare event (5% of HCC patients) and factors contributing to<br />

their onset remain poorly understood. The aim of the case-control<br />

exploratory study NoFLIC was to identify factors involved<br />

in hepatocarcinogenesis in non-fibrotic liver. Patients and<br />

methods NoFLIC is a French multicenter case-control study that<br />

included consecutive patients who developed HCC in a non-fibrotic<br />

liver, F0 or F1, proven by histological analysis. Controls<br />

were matched for age and gender, they were recruited in<br />

people with gastro-intestinal screening and normal endoscopy,<br />

or in rheumatology departments. Clinical data were collected<br />

including the parameters of the metabolic syndrome defined<br />

according to the NCEP-ATP3 criteria, viral serology, quantification<br />

of alcohol and tobacco use. Data were taken from the<br />

patient’s medical record, the results of blood test performed for<br />

this study and a specific epidemiologic survey. Logistic regressions<br />

adjusted for age, sex and geographical origin, were performed<br />

to determine the association of each of the risk factors<br />

with the status of the patient (case / control). A multivariate<br />

analysis was performed to study the effect of significant risk<br />

factors. Results A total of 109 nfHCC (cases) and 163 controls<br />

were included in 12 hospitals between 2010 and 2013.<br />

Mean age of the nfHCC patients was 64.4 years, with 68.8%<br />

male. Active HBV infection was detected in 6 cases (OR 11.4,<br />

95% 1.18-110.84, p=0.035). History of alcohol abuse was<br />

more frequent in nfHCC (OR 2.13, 95% 1.09-4.16 p=0.026).<br />

A metabolic syndrome was identified in 28% of the patients<br />

developing nfHCC (OR 3.3; 95% CI 1.62-6.82 p=0.001).<br />

Among the components of metabolic syndrome, treated diabetes<br />

or fasting glucose ≥ 6.1mmol/l was significantly associated<br />

with nfHCC occurrence (OR 3.08; 95% CI 1.67-5.69<br />

p=0.0002), while there was no significant association with<br />

BMI, hypertension and dyslipidemia. Smoking (defined by at<br />

least 100 smoked cigarettes and at least with a weekly smoking<br />

habit) was also associated with nfHCC (OR 2.18 adjusted<br />

for age and sex; 95% CI 1.23-3.83 p = 0.0067). In multivariate<br />

analysis, HBV infection (p=0.024), metabolic syndrome<br />

(P=0.0023) and tobacco use (p=0.0062) were independent


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 277A<br />

133<br />

Neurodevelopmental Outcomes in Patients with Biliary<br />

Atresia and Native Liver at Ages 1 and 2 Years: Results<br />

from ChiLDReN<br />

Vicky L. Ng 1 , Lisa G. Sorensen 16 , Estella M. Alonso 2 , Emily M.<br />

Fredericks 3 , Wen Ye 21 , Saul J. Karpen 17 , Benjamin Shneider 18 ,<br />

Jorge A. Bezerra 4 , Jean P. Molleston 5 , Karen F. Murray 6 , Philip<br />

Rosenthal 7 , Kasper S. Wang 19 , Kathleen M. Loomes 8 , Paula M.<br />

Hertel 9 , Nanda Kerkar 20,22 , Kathleen B. Schwarz 10 , Yumirle P.<br />

Turmelle 11 , Barbara A. Haber 12 , Averell H. Sherker 15 , John C.<br />

Magee 13 , Ronald J. Sokol 14 ; 1 Division of Pediatric Gastroenterology,<br />

Hepatology and Nutrition, The Hospital for Sick Children,<br />

University of Toronto, Toronto, ON, Canada; 2 Division of Pediatric<br />

Gastroenterology, Hepatology and Nutrition, Ann & Robert H.<br />

Lurie Children’s Hospital, Chicago, IL; 3 Division of Child Behavioral<br />

Health, University of Michigan and C.S. Mott Children’s<br />

Hospital, Ann Arbor, MI; 4 Division of Pediatric Gastroenterology,<br />

Hepatology and Nutrition, Cincinnat Children’s Hospital Medical,<br />

Cincinnati, OH; 5 Pediatric Gastroenterology, Hepatology<br />

and Nutrition, Indiana Universtiy School of Medicine/Riley Hospital<br />

for Children, Indianapolis, IN; 6 Division of Gastroenterology<br />

and Hepatology, University of Washington Medical Center,<br />

Seattle Children’s, Seattle, WA; 7 Division of Gastroenterology,<br />

Hepatology and Nutrition, Department of Pediatrics, University<br />

of California, San Francisco, San Francisco, CA; 8 Pediatric Gastroenterology,<br />

Hepatology and Nutrition, Children’s Hospital of<br />

Philadelphia, Philadelphia, PA; 9 Baylor College of Medicine,<br />

Houston, TX; 10 Johns Hopkins School of Medicine, Baltimore, MD;<br />

11 Pediatric Gastroenterology, Hepatology and Nutrition, Washington<br />

University School of Medicine, St. Louis, MO; 12 Infectious<br />

Disease, Clinical Research, Merck, North Wales, PA; 13 University<br />

of Michigan Medical School, Ann Arbor, MI; 14 Section of Pediatric<br />

Gastroenterology, Hepatology and Nutrition, Department of Pediatrics,<br />

University of Colorado School of Medicine, Denver, CO;<br />

15 Liver Diseases Research Branch, National Institute of Diabetes<br />

and Digestive and Kidney Diseases, National Institutes of Health,<br />

Bethesda, MD; 16 Ann & Robert H. Lurie Children’s Hospital of<br />

Chicago, Northwestern University Feinberg School of Medicine,<br />

Chicago, IL; 17 Pediatric Gastroenterology, Hepatology and Nutrition,<br />

Emory University School of Medicine/Children’s Healthcare<br />

of Atlanta, Atlanta, GA; 18 Pediatric Gastroenterology, Hepatology<br />

and Nutrition, Baylor College of Medicine, Houston, TX; 19 Division<br />

of Pediatric Surgery, Children’s Hospital Los Angeles, Los Angeles,<br />

CA; 20 Children’s Hospital of Los Angeles, Los Angeles, CA;<br />

21 Department of Biostatistics, University of Michigan, Ann Arbor,<br />

MI; 22 Mount Sinai, New York, NY<br />

Background: One of the most serious consequences of liver<br />

disease and malnutrition early in life is neurodevelopmental<br />

(ND) delay for which prompt identification and intervention<br />

may improve long-term outcomes. We hypothesized i) biliary<br />

atresia (BA) patients with their native liver at ages 1 and 2<br />

yrs will have significant ND delays and ii) that specific demographic<br />

and clinical variables will predict worse ND outcomes.<br />

Methods: Infants with BA, birth weight >2 kg, who received<br />

hepatoportoenterostomy (HPE) and were enrolled between<br />

2004-2012 in the multi-center, prospective NIDDK-funded<br />

ChiLDReN study (PROBE-NCT00061828) underwent ND testing<br />

at ages 12±2 mos (T1) and 24±2mos (T2) using either<br />

Bayley Scales of Infant Development-2(B2) or Bayley-3(B3),<br />

based on era. A subset of these infants was also enrolled in a<br />

blinded randomized placebo-controlled study of corticosteroids<br />

as adjunctive therapy to HPE (START–NCT00294684). Scores<br />

(normative mean=100±15) were categorized as ≥100, 85-99,<br />

70-84,


278A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

were included. Clinical and biochemical data were collected<br />

and compared with subsequent endoscopies to compare the<br />

efficacy and safety of prophylactic therapy versus non-treatment.<br />

Results: 273 endoscopies from 70 patients (42M) mean<br />

age 8.8 +/- 4.7y, were included. Diagnosis was portal vein<br />

thrombosis, biliary atresia, and other liver diseases, such as<br />

Wilson’s disease, autoimmune hepatitis, cystic fibrosis and<br />

congenital hepatic fibrosis in 23 (33%), 25 (35%) and 22<br />

(32%), respectively. Subjects were in the surveillance program<br />

for mean time of 4.8 years +/- 2.7 years (range, 9 months-8<br />

years). Variceal grading was as per UK national guidelines.<br />

Prophylactic treatment was administered in 177 (65%) endoscopies<br />

versus 96 (35%). Treatment group showed improvement<br />

of varices in 127 (72%) endoscopies, no change in 39<br />

(22%) and worsening of varices in 11 (6%). The non-treatment<br />

group showed improvement of varices in 16 (16%) endoscopies,<br />

no change in 38 (40%) and worsening of varices in 42<br />

(43%). Mean values for serum bilirubin, platelet count (PLT),<br />

haemoglobin (Hb) and INR at the time of endoscopy in the<br />

prophylaxis group and non-prophylaxis treatment group were<br />

18.65mg/dL and 21.11mg/dL [n.v. 3-20], 109.1x10 9 /L and<br />

103.95x10 9 /L [n.v. 180-400], 118.91g/L and 110.5g/L<br />

[n.v. 115-155], 1.28 and 1.25 [n.v. 0.9-1.2], respectively.<br />

No significant difference was found between treatment and<br />

non-treatment groups in bilirubin (p=0.4), PLT (p=0.4), Hb<br />

(p=0.38) or INR (p=0.5). Blood products prior to the procedure<br />

were required in 33 endoscopies (12%), as per protocol.<br />

Seven (10%) patients suffered a single GI bleed, all but one of<br />

which were in the no prophylactic treatment group. No other<br />

complications were recorded in either group. Conclusion: Prophylactic<br />

treatment of GI varices results in an overall reduction<br />

of varices and GI bleeds. Children with PTH could benefit from<br />

surveillance endoscopy programs with prophylactic treatment<br />

in the form of EVL or EST as it has been shown to be both safe<br />

and effective.<br />

Disclosures:<br />

The following authors have nothing to disclose: Harry Sutton, Mark Davenport,<br />

Alastair J. Baker, Anil Dhawan, Tassos Grammatikopoulos<br />

135<br />

Molecular genetic dissection of 101 neonatal/infantile<br />

intrahepatic cholestasis using targeted next-generation<br />

sequencing<br />

Takao Togawa, Tokio Sugiura, Koichi Ito, Takeshi Endo, Shinji<br />

Saitoh; Department of Pediatrics and Neonatology, Nagoya City<br />

University Graduate School of Medical Sciences, Nagoya, Japan<br />

Neonatal/infantile intrahepatic cholestasis (NIIC) is a heterogeneous<br />

disorder and caused by mutations in a number of<br />

genes, making molecular diagnosis challenging. Our aim is<br />

to clarify the molecular diagnosis for subjects with NIIC using<br />

next-generation sequencing (NGS) and bioinformatics pipeline.<br />

We further sought genotype-phenotype correlation to<br />

help comprehensive understanding of NIIC. Neonatal/infantile<br />

cholestasis was defined by the level of serum direct bilirubin;<br />

D.Bil≥1 mg/dL, if T.Bil8.6 kPa to detect METAVIR F3-F4 fibrosis.<br />

Several <strong>studies</strong> in children have used different LSM cut-points to<br />

identify fibrosis. However, validation of these LSM cut-points in<br />

children has not been reported. METHODS: Patients at Boston<br />

Children’s Hospital who underwent LSM from October 2006-<br />

May 2015 were eligible for this prospective two-phase study.<br />

All patients must have had a liver biopsy within 12 months<br />

prior to enrollment. The METAVIR system was used to stage<br />

fibrosis by pathologists blinded to LSM. Patients enrolled from<br />

October 2006- March 2011 were in the test group used to<br />

develop cut-points for discriminating F3-F4 and F4 fibrosis.<br />

Patients enrolled from April 2011-May 2015 were the validation<br />

group to test the accuracy of these cut-points. Diagnostic<br />

performance of TE was assessed by sensitivity (Se), specificity<br />

(Sp), accuracy (Ac) and area under the receiver operating<br />

characteristic (AUC). RESULTS: 263 subjects were enrolled; 97<br />

in the test group and 166 in the validation group. Clinical/<br />

demographic data of the two groups were similar. Overall,<br />

the subjects were 46% female, aged 2 mo–24 yrs (median


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 279A<br />

13y), with indications for biopsy of autoimmune LD (19%),<br />

viral infection (18%), cholestatic LD (9%), PSC (12%), non-alcoholic<br />

fatty LD (11%), and post-transplantation (2%). METAVIR<br />

stages were: F0 (20%), F1 (29%), F2 (20%), F3 (17%), F4<br />

(14%). Median time from LSM to biopsy was 50 days (IQR<br />

21-144). Fasting prior to TE was known for 23%. In the test<br />

group, the cut-point to discriminate F3-F4 was >8.6 kPa with<br />

Se=79%, Sp=83%, Ac=81% and AUC=0.83. The cut-point for<br />

F4 was >11.5 kPa (Se=74%, Sp=82%, Ac=80%, AUC=0.81).<br />

Applied to the validation group, the cut-point of >8.6 kPa to<br />

detect F3-F4 had Se=67%, Sp=69% and Ac=69%. For subjects<br />

who fasted prior to TE (N=45), the performance for this<br />

cut-point improved (Se=92%, Sp=70%, Ac=76%). To detect<br />

F4, the cut-point of >11.5 kPa had Se=78%, Sp=80% and<br />

Ac=80%. In fasting subjects (N=45), the accuracy improved<br />

(Se=86%, Sp=84%, Ac=84%). CONCLUSIONS: This is the first<br />

study in children and young adults in which LSM cut-points for<br />

F3-F4 and F4 fibrosis are defined and evaluated in separate<br />

test and validation samples. This study supports recent findings<br />

that accuracy of TE improves with fasting. TE can be an important<br />

tool for identifying advanced fibrosis in children.<br />

Disclosures:<br />

Christine K. Lee - Grant/Research Support: Echosens<br />

Maureen M. Jonas - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Gilead, Bristol Myers Squibb, Roche, Echosens<br />

The following authors have nothing to disclose: Paul D. Mitchell, Roshan Raza,<br />

Sarah Harney, Shanna M. Wiggins<br />

137<br />

Hospitalizations for Respiratory Syncytial Virus, Influenza,<br />

and Other Vaccine Preventable Illnesses in Liver<br />

Transplant Recipients at Freestanding US Children’s<br />

Hospitals<br />

Amy G. Feldman 1,4 , Brenda Beaty 2 , Shikha Sundaram 1,4 , Allison<br />

Kempe 3 ; 1 Pediatric Gastroenterology, Hepatology and Nutrition,<br />

Children’s Hospital Colorado, Aurora, CO; 2 University of Colorado,<br />

Aurora, CO; 3 Pediatrics, Children’s Hospital Colorado,<br />

Aurora, CO; 4 University of Colorado School of Medicine, Aurora,<br />

CO<br />

Objectives: Influenza and respiratory syncytial virus (RSV) are<br />

common infections requiring pediatric hospitalizations with<br />

annual rates of 0.05% and 0.5% respectively. Rates of these<br />

same illnesses in the pediatric liver transplant population are<br />

unknown. The goals of this study were to examine among liver<br />

transplant recipients at centers participating in the Pediatric<br />

Health Information System (PHIS) dataset: (1) number of hospitalizations<br />

for RSV, influenza, and other vaccine preventable<br />

illnesses (VPIs); (2) morbidity and mortality associated<br />

with these hospitalizations; and (3) costs associated with these<br />

hospitalizations. Methods: Retrospective cohort study of subjects<br />

< 18 years of age who underwent liver transplantation at<br />

a PHIS center between January 1, 2004, and December 31,<br />

2013. Subsequent hospitalizations during this time period for<br />

RSV, influenza, and 10 other VPIs were ascertained using ICD-<br />

9-CM diagnosis codes. Data were collected on clinical care,<br />

outcomes, and costs during these hospitalizations. Results:<br />

2,554 pediatric liver transplant recipients were identified;<br />

511 (20.0%) of the patients had a total of 799 cases of RSV,<br />

influenza, or other VPI. Overall, RSV, influenza, and rotavirus<br />

were the most common infections (7.2%, 6.6%, and 5.8% of<br />

cohort respectively). The mean time from transplant to infection<br />

was 1.7 years (SD 1.7 years). Ninety-two patients (3.6%)<br />

had infection during their transplant hospitalization, 8 of these<br />

had multiple infections. In the first year post-transplant, 4.0%<br />

of all patients had a hospitalization for RSV and 3.3% had<br />

a hospitalization for influenza. The death rate during these<br />

hospitalizations was 4.9% for RSV and 2.4% for influenza.<br />

Excluding those infections that occurred during initial transplant<br />

hospitalization (when all patients receive mechanical ventilation<br />

and ICU care), rejection occurred in 15.9% of patients<br />

with RSV and 13.4% of patients with influenza, ventilation in<br />

9.3% (RSV) and 5.9% (influenza), and ICU level care in 20.9%<br />

(RSV) and 12.8% (influenza). The median inflation adjusted<br />

costs for hospitalization were $ 15,934 (RSV) and $7806<br />

(influenza). Conclusions: VPIs occur in 1 of every 5 pediatric<br />

liver transplant recipients at a rate of up to 66 times higher than<br />

in the general pediatric population. RSV and influenza were<br />

the most common infections and most cases occurred in the<br />

first two years following transplant, demonstrating the importance<br />

of ensuring vaccination in all patients prior to liver transplant.<br />

Further research on clinical utility and cost-effectiveness<br />

of Palivizumab to prevent RSV in select transplant candidates/<br />

recipients is needed.<br />

Disclosures:<br />

The following authors have nothing to disclose: Amy G. Feldman, Brenda Beaty,<br />

Shikha Sundaram, Allison Kempe<br />

138<br />

Liver Fibrosis is Present in One-Third of Adults with<br />

Alpha-1 Antitrypsin Deficiency Without Overt Liver Disease<br />

Virginia C. Clark 1 , George W. Marek 2 , Tracie L. Kurtz 3 , Chen<br />

Liu 4 , Farshid Rouhani 3 , Mark Brantly 3 ; 1 Medicine, Division of Gastroenterology,<br />

Hepatology, and Nutrition, University of Florida,<br />

Gainesville, FL; 2 Medicine, University of Florida, Gainesville, FL;<br />

3 Medicine, Pulmonary Critical Care, University of Florida, Gainesville,<br />

FL; 4 Pathology, University of Florida, Gainesville, FL<br />

Rationale: Alpha-1 antitrypsin deficiency (AATD) is a risk factor<br />

for the development of cirrhosis and liver related mortality in<br />

affected individuals. We hypothesize individuals with AATD<br />

have subclinical liver injury and fibrosis not detected by routine<br />

testing. The primary aim of this study is to define the prevalence<br />

and histologic spectrum of liver disease in adults with<br />

AATD. Methods: Adults with confirmed AATD Pi*ZZ and other<br />

rare AATD alleles participated in the IRB approved study after<br />

informed consent was obtained. A percutaneous liver biopsy<br />

was performed on eligible subjects. A single liver pathologist<br />

staged the biopsy for fibrosis (Ishak stage 1-6). The findings of<br />

AATD in PAS-D globules were scored qualitatively to describe<br />

the total observed (rare 0-1; moderate 2-3; severe 4-5). Individuals<br />

with a clinical history of decompensated liver disease were<br />

excluded. Results: The predominant AATD genotype is Pi*ZZ<br />

(n=73), and two patients with rare alleles are included Pi*ZM-<br />

Malton,<br />

and Pi*null. The mean age was 57 years, range 25-71<br />

and 61% (n=46) were female. The mean age of AATD diagnosis<br />

is 45.3 years, range 8 months-68 years. Pulmonary symptoms<br />

were present at AATD diagnosis in 64% (n=48) and 28%<br />

(n=21) were diagnosed because of family member. Only 5%<br />

(n=4) were identified by abnormal liver test. Pulmonary function<br />

characteristics included a baseline FEV1 1.95L (0.37-5.69L)<br />

and % predicted FEV1 of 61.5 (14-117). The prevalence of<br />

clinically significant liver fibrosis defined by Ishak fibrosis score<br />

≥2 is 34.7% (n=25). Men were more likely to have advanced<br />

fibrosis (Ishak ≥3, 22.2%, n=6) compared to women (2.2%<br />

n=1). ALT (22.5 vs 29.3 IU/L, p=0.03) and GGT (26.3 vs<br />

51.7, p=0.0006) were higher in those with Ishak ≥2, but mean<br />

values were within normal range. PAS-D globules correlated<br />

with fibrosis stage (spearman r= 0.5224, p


280A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and risk for liver disease in AATD is needed to direct screening<br />

and novel therapies toward patients who could benefit most.<br />

In this cross section of older AATD subjects, over one-third had<br />

clinically significant liver fibrosis confirmed by biopsy. This is<br />

much higher than our previously reported prevalence of 7.9%<br />

using patient self-reported data. We also confirmed our previous<br />

finding that ALT and GGT for affected subjects fall within<br />

normal range, which limits clinical utility. The histologic spectrum<br />

was varied both in fibrosis and qualitative presence of<br />

AATD globules. Future <strong>studies</strong> will evaluate clinical risk factors<br />

and predictors for advanced liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Virginia C. Clark, George W.<br />

Marek, Tracie L. Kurtz, Chen Liu, Farshid Rouhani, Mark Brantly<br />

139<br />

Bile Acid Sequestrant Prevents NAFLD and NASH in<br />

Western Diet Fed Mice Independent of FXR<br />

Yuhuan Luo 1 , Xiaoxin Wang 1 , David J. Orlicky 2 , Moshe Levi 1 ;<br />

1 Medicine, Univ Colorado, Aurora, CO; 2 Pathology, University of<br />

Colorado, Aurora, CO<br />

Bile acid sequestrants (BAS) are orally administered nonabsorbable<br />

polymers that decrease serum cholesterol and serum glucose<br />

in patients and animal models of type 2 diabetes mellitus.<br />

We evaluated the effect of the BAS sevelamer in mice with diet<br />

induced obesity and insulin resistance that develop NAFLD and<br />

NASH. C57BL/6J male mice were fed with 1) control low fat<br />

(LF) diet or 2) western (WD) diet (high fat, high sucrose, cholesterol)<br />

for 5 months. The diets contained a) no addition or b) 2%<br />

sevelamer. Treatment with sevelamer prevented the increases<br />

in body weight, liver weight, serum triglycerides and serum<br />

cholesterol. Histopathological analysis revealed diet induced<br />

NAFLD and NASH mouse model resulted in panlobular macrosteatosis<br />

with some microvesicular steatosis, and increases in<br />

peri-portal inflammation and peri-portal and zone 1 sinusoidal<br />

fibrosis extending into zones 2 & 3 in places. Sevelamer treatment<br />

significantly decreased steatosis, inflammation and fibrosis<br />

scores. There was decreased FXR signaling in the intestine<br />

and the liver (FGF15 and SHP in the intestine as well as SHP<br />

in the liver). There was also increased liver bile acid synthesis<br />

(Cyp7a1) and expression of bile acid transporters. In addition,<br />

there were significant decreases in SREBP-1c, ACC and FAS,<br />

resulting in decreased neutral lipid accumulation. Sevelamer<br />

also prevented or reversed WD-induced upregulation of profibrotic<br />

(COL1A1, α-SMA, TGF-β) and proinflammatory (CCL2,<br />

IL-1β, and TNF-α) genes in the liver. These effects of sevelamer<br />

were independent of FXR as in FXR KO mice fed a western diet<br />

sevelamer had similar beneficial effects in improving NAFLD/<br />

NASH as in wild type mice. BAS treatment therefore prevents<br />

diet induced NAFLD and NASH. This beneficial effect is largely<br />

independent of FXR but may be mediated through activation of<br />

the TGR5 signaling pathway.<br />

Disclosures:<br />

Moshe Levi - Grant/Research Support: Intercept, Genzyme-Sanofi, Daiichi Sankyo,<br />

Merck, Novartis<br />

The following authors have nothing to disclose: Yuhuan Luo, Xiaoxin Wang,<br />

David J. Orlicky<br />

140<br />

Targeting enterohepatic bile acid signaling as a novel<br />

approach to modulate hepatic autophagic activity in<br />

maintaining cholesterol homeostasis<br />

Yifeng Wang, Yifeng Ding, Hong-Min Ni, Wen-Xing Ding, Tiangang<br />

Li; Pharmacolgy, KUMC, Kansas City, KS<br />

Liver plays a key role in whole body cholesterol homeostasis.<br />

In addition, hepatic free cholesterol accumulation results in<br />

hepatocyte injury in non-alcoholic steatohepatitis (NASH). Lysosomes<br />

play an important role in maintaining cellular cholesterol<br />

homeostasis by regulating cholesterol ester hydrolysis that<br />

affects pathways such as cholesterol uptake, secretion and bile<br />

acid synthesis. New <strong>studies</strong> revealed that autophagy mediates<br />

the cellular transport of cholesterol ester from lipid droplets to<br />

lysosomes. Goal: this study investigated the role and regulation<br />

of hepatic autophagy in cholesterol metabolism. Methods: the<br />

effect of cholesterol accumulation on autophagy flux and lysosome<br />

function was investigated in hepatocytes and in mice.<br />

The impact of autophagy impairment on hepatic and plasma<br />

cholesterol metabolism was studied in liver-specific ATG5<br />

knockout mice. The effect of cholestyramine on hepatic autophagy<br />

was studied in mice. Results: Feeding mice a cholesterol-containing<br />

diet or an atherogenic diet for 6 days increased<br />

liver LC3-II and p62 levels in response to hepatic cholesterol<br />

accumulation. Both free cholesterol loading and treatment with<br />

low density lipoprotein (LDL) increased LC3B-II and p62 levels<br />

in human hepatocytes and HepG2 cells, however, only free<br />

cholesterol loading, but not LDL treatment, increased lysosome<br />

size and decreased cathepsin B activities in HepG2 cells, suggesting<br />

free cholesterol accumulation can also impair lysosome<br />

function. Consistently, inhibition of cholesterol esterification<br />

with an ACAT inhibitor to increase cellular free cholesterol<br />

levels blocked autophagosome/lysosome fusion and caused<br />

CASPASE-3 activation in HepG2 cells. In autophagy-defective<br />

liver-specific ATG5 knockout mice, hepatic cholesterol was<br />

unchanged, but FPLC analysis showed that plasma LDL-cholesterol<br />

was significantly elevated. As previous <strong>studies</strong> suggested<br />

that autophagy was impaired in NASH, we found that cholestyramine<br />

feeding increased hepatic autophagy in both lean<br />

and ob/ob mice, which was possibly due to the removal of<br />

bile acid-mediated inhibition of autophagosome/lysosome<br />

fusion. Conclusion: this study suggests that hepatic autophagy<br />

is required for maintaining whole body cholesterol homeostasis.<br />

Hepatic free cholesterol accumulation impairs hepatic<br />

autophagy, which exacerbates hypercholesterolemia and<br />

hepatocyte-injury in NASH. As we recently showed that bile<br />

acids inhibit hepatic autophagy, targeting enterohepatic bile<br />

acid signaling by bile acid binding resin may be an effective<br />

approach to induce hepatic autophagy and attenuate cholesterol-induced<br />

liver injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yifeng Wang, Yifeng Ding, Hong-<br />

Min Ni, Wen-Xing Ding, Tiangang Li<br />

141<br />

Anti-Fibrotic Effect of Autotaxin and LPA1R Antagonists<br />

in a Rodent Model of NASH.<br />

Malavi Madireddi 1 , Gretchen Bain 3 , James Swaney 2 , Brian J.<br />

Murphy 1 , Simeon Taylor 1 ; 1 Metabolic and Fibrotic Diseases, Bristol<br />

Myers Squibb, West Windsor, NJ; 2 Inception Sciences, San<br />

Diego, CA; 3 PharmAkea, Inc., San Diego, CA<br />

Nonalcoholic steatohepatitis (NASH) is a condition where<br />

increased fat accumulation concomitant with inflammation<br />

results in liver cell damage. Patients with NASH have elevated


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 281A<br />

serum lipids and diabetes. Early disease is clinically asymptomatic,<br />

therefore often presenting only when critical hepatic<br />

function is compromised. A wound healing response in general<br />

occurs in injured liver tissue, and the persistence of this<br />

response may result in liver fibrosis. At this time, there is no<br />

treatment for NASH, and left untreated leads to hepatocellular<br />

carcimona and liver cirrhosis. Current management focuses on<br />

treating underlying cause of injury, as there are no targeted<br />

therapeutics to interrupt fibrotic response; liver transplant is<br />

the only option for late-stage cirrhosis. The enzyme Autotaxin<br />

(ATX) hydrolyzes lysophosphatidylcholine to produce lysophosphatidic<br />

acid (LPA), a multi-functional bioactive lipid mediator.<br />

ATX is a major determinant of LPA levels in circulation, and the<br />

pathophysiological function of ATX is attributed to its ability<br />

to produce LPA. There is mounting literature supporting serum<br />

ATX as an indicator of the severity of liver disease. Here we<br />

investigated two small molecule inhibitors of ATX and LPA1<br />

receptor in the STAM ® rodent model of liver fibrosis to pharmacologically<br />

interrogate these targets in NASH. AM063 and<br />

BMT-053011 are low nanomolar inhibitor of ATX and LPA1<br />

receptor respectively. The STAM ® mouse model of NASH is<br />

induced by a single subcutaneous injection of streptozotocin<br />

two days after birth and by feeding a high fat diet. Starting at<br />

6 weeks after disease initiation, BMT-053011 was administered<br />

twice daily, while AM063 and Telmisartan were administered<br />

once daily. Treatment was continued for the following 3<br />

weeks after initiation. AM063 showed significant decreases in<br />

plasma TG, fibrosis area and TNF-a expression levels, with a<br />

decreasing trend in blood glucose and macrophage activation<br />

compared to the Vehicle group. Treatment with BMT-053011<br />

showed significant decreases in liver weight, NAS, the fibrosis<br />

area and inflammation. The results, together with the decrease<br />

in ALT and inflammatory gene expression levels, demonstrate<br />

anti-inflammatory and hepatoprotective effects. Importantly, as<br />

NAS is one of the clinical endpoints for assessing the activity<br />

of NASH (Sanyal AJ. et al., Hepatology, 2011; 54:344), the<br />

observed changes in the treatment groups suggest potential<br />

translatability of BMT-053011 as an anti-NASH therapeutic<br />

for humans.<br />

Disclosures:<br />

Malavi Madireddi - Employment: Bristol Myers Squibb<br />

Gretchen Bain - Employment: PharmAkea Therapeutics<br />

Simeon Taylor - Consulting: Isis Pharmaceuticals, Aegerion, Yabao; Stock Shareholder:<br />

Bristol-Myers Squibb, Gilead, Celgene<br />

The following authors have nothing to disclose: James Swaney, Brian J. Murphy<br />

142<br />

Treatment with the FXR-TGR5 dual agonist INT-767<br />

arrests and reverses progression of NASH in mice fed a<br />

Western diet<br />

Xiaoxin Wang 1 , Yuhuan Luo 1 , David J. Orlicky 2 , Luciano Adorini 3 ,<br />

Moshe Levi 1 ; 1 Medicine, Univ Colorado, Aurora, CO; 2 Pathology,<br />

University of Colorado, Aurora, CO; 3 Intercept Pharmaceuticals,<br />

New York, NY<br />

Obesity and insulin resistance have become leading causes<br />

of nonalcoholic fatty liver disease (NAFLD) and nonalcoholic<br />

steatohepatitis (NASH). In the present study, we determined in<br />

mice with established NASH if treatment with INT-767, a dual<br />

agonist of the nuclear receptor farnesoid X receptor (FXR) and<br />

the G protein coupled receptor TGR5, can decrease disease<br />

progression. In these <strong>studies</strong>, C57BL/6Jmice were first fed a<br />

low fat (LF) or a Western diet (WD: high fat, high sucrose, and<br />

high cholesterol) for 3 months and then treated orally for an<br />

additional 2 months with vehicle or INT-767, 30 mg/kg body<br />

weight/d. Compared to mice fed a LF diet, feeding a WD<br />

for 5 months resulted in panlobular macrosteatosis with some<br />

microvesicular steatosis, increased peri-portal inflammation and<br />

lipogranuloma formation, and peri-portal and zone 1 sinusoidal<br />

fibrosis extending in places into zones 2 & 3. When mice<br />

on WD were treated with INT-767, there were marked and significant<br />

decreases in the percentage of mice showing hepatic<br />

macrosteatosis (83%), microsteatosis (100%), inflammation<br />

(100%) and fibrosis (47%) compared to non-treated WD fed<br />

mice. These histopathological changes were accompanied by<br />

significant decreases in liver cholesterol content (52%), as well<br />

as decreases in the lipogenic transcription factor SREBP-1c,<br />

pro-inflammatory mediator MCP-1, and profibrotic matrix protein<br />

Col1a1. INT-767 treatment also increased liver expression<br />

of proteins involved in mitochondrial biogenesis and energy<br />

expenditure including AMPK, Sirtuin 1, PGC-1α, Sirtuin 3, and<br />

ERRα. Using FXR knockout mice, the FXR selective agonist INT-<br />

747 (obeticholic acid), and the TGR5 specific agonist INT-777,<br />

we further determined that FXR plays a major role in mediating<br />

the effects of INT-767 in the liver, including the increases in<br />

PGC-1α and Sirtuin 3. The dual FXR-TGR5 agonist INT-767 is<br />

therefore able to markedly and significantly arrest and reverse<br />

progression of liver disease even when treatment is started in<br />

the presence of obesity, insulin resistance, and NASH.<br />

Disclosures:<br />

Luciano Adorini - Employment: Intercept Pharmaceuticals<br />

Moshe Levi - Grant/Research Support: Intercept, Genzyme-Sanofi, Daiichi Sankyo,<br />

Merck, Novartis<br />

The following authors have nothing to disclose: Xiaoxin Wang, Yuhuan Luo,<br />

David J. Orlicky<br />

143<br />

DRX-065, the stabilized R-enantiomer of pioglitazone is<br />

without PPARg agonist activity and exhibits the beneficial<br />

in vivo pharmacodynamic effects for the treatment<br />

of NASH<br />

Sheila H. DeWitt, Vincent Jacques, Lex H. Van der Ploeg; DeuteRx,<br />

Andover, MA<br />

Background: Pioglitazone is approved for the treatment of<br />

type 2 diabetes mellitus (T2DM) and is currently the only drug<br />

recommended off-label to treat patients with biopsy-proven<br />

nonalcoholic steatohepatitis (NASH). However, the use of<br />

pioglitazone for NASH has been limited due the side effect<br />

of weight gain. Pioglitazone is a mixture of two mirror-image<br />

compounds (the R- and S-enantiomers) that rapidly interconvert<br />

in vivo. The pharmacokinetic (PK) and pharmacodynamic<br />

(PD) properties of the individual enantiomers has never been<br />

reported in humans. Methods: We replaced the exchangeable<br />

hydrogen at the chiral center of pioglitazone by deuterium,<br />

a stable, naturally occurring isotope of hydrogen. Deuterium<br />

at the chiral center stabilizes the enantiomers and reduces<br />

their rate of interconversion. Stability <strong>studies</strong> were completed<br />

in aqueous buffer and in mouse and human plasma. In vitro<br />

PPARg agonist activity was determined in a TRAP220 coactivator<br />

recruitment assay. In vitro mitochondrial respiration activity<br />

was assessed in intact C2C12 cells. PK and weight gain <strong>studies</strong><br />

of the enantiomers were completed in male C57BL/6 mice.<br />

Efficacy was evaluated in a db/db mouse model of T2DM,<br />

where effects on metabolic and inflammatory markers were<br />

assessed. Results: Preparation of the deuterium-stabilized enantiomers<br />

of pioglitazone, d-R-pioglitazone (DRX-065) and d-S-pioglitazone<br />

(d-S-pio) enabled the evaluation of the properties of<br />

the individual enantiomers. DRX-065 is not a PPARg agonist<br />

at concentrations up to 100 mM while d-S-pio accounts for<br />

all of the PPARg agonist activity observed with racemic pioglitazone.<br />

DRX-065 is exclusively responsible for the effects of


282A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

pioglitazone on mitochondrial respiration. In the db/db mouse<br />

model of T2DM, DRX-065 significantly decreases blood glucose,<br />

serum triglycerides, and fatty acids, as well as markers of<br />

inflammation (cytokines, serum amyloid A). At a therapeutically<br />

effective dose in mice, DRX-065 also does not induce weight<br />

gain, unlike pioglitazone or d-S-pio. The effects of d-S-pio on<br />

weight gain are expected based on its strong PPARg agonist<br />

activity. Conclusion: We demonstrated in preclinical experiments<br />

that the stabilized deuterated R-enantiomer of pioglitazone,<br />

DRX-065, has pharmacological properties desirable<br />

for the treatment of NASH (mitochondrial function modulation,<br />

non-steroidal anti-inflammatory effects, and glucose lowering<br />

effects) without the undesired PPARg-related weight gain side<br />

effects. DRX-065 represents a potentially significant therapeutic<br />

improvement over pioglitazone, a drug already recommended<br />

off-label for the treatment of NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sheila H. DeWitt, Vincent<br />

Jacques, Lex H. Van der Ploeg<br />

144<br />

Peretinoin, an acyclic retinoid, suppresses steatohepatitis<br />

and development of tumorigenesis by activated<br />

Atg16L1-dependent autophagy in mice fed an atherogenic<br />

high-fat diet.<br />

Hikari Okada 1 , Masao Honda 1 , Kai Takegoshi 1 , Taro Yamashita 1 ,<br />

Naoto Nagata 2 , Toshinari Takamura 2 , Takuji Tanaka 3 , Shuichi<br />

Kaneko 1 ; 1 Gastroenterology, Kanazawa University Graduate<br />

School of Medical Science, Kanazawa, Japan; 2 Department of<br />

Disease Control and Homeostasis, Kanazawa University Graduate<br />

School of Medical Sciences, Kanazawa, Japan; 3 Cancer Research<br />

and Prevention, The Tohkai Cytopathology Institute, Gifu, Japan<br />

Objective Peretinoin, acyclic retinoid, is under phase III clinical<br />

trials in Japan to prevent the recurrence and development of<br />

HCC after surgical removal of the primary tumors. The mechanism<br />

of peretinoin for preventing HCC remains unclear. Mice<br />

fed an atherogenic high-fat diet (Ath HFD) developed steatohepatitis<br />

followed by hepatic fibrosis and HCC progression<br />

(Hepatology 2007). Here we investigated the suppressive<br />

effects of peretinoin on steatohepatitis and tumorigenesis in<br />

Ath HFD mice. Materials and Methods Three groups of 8-weekold<br />

mice (n=15-20/group) were fed a control diet or Ath HFD<br />

containing 0.01% or 0.03% peretinoin for 12, 30, and 60<br />

weeks. Then, 0.01% peretinoin was added to the Ath HFD<br />

at 40 weeks to examine the reversible effect of peretinoin on<br />

established fibrosis and steatosis in the liver. The degree of liver<br />

steatosis, hepatic fibrosis, tumor incidence, and liver weight<br />

was calculated. Expression of IL6, IL1β TNFα, CEBPα, collagen<br />

I/IV, pSTAT3, pNFkβ, ATG5, ATG7, ATG16L1, LC3b,<br />

and Lamp2 was evaluated by immunohistochemical staining,<br />

real-time PCR, and western blotting. The promoter analysis of<br />

ATG16L1 was performed by reporter gene assay and ChIP<br />

assay. Primary hepatocytes were isolated from male C57BL/6<br />

mice treated Ath HFD for 12 weeks and the effect of peretinoin<br />

on primary hepatocytes was evaluated in vitro. Results Mice<br />

fed an Ath HFD developed liver steatosis and liver fibrosis after<br />

12 and 30 weeks, and developed liver tumors after 60 weeks<br />

at 70% frequency. Peretinoin markedly reduced steatosis and<br />

fibrosis at 12 and 30 weeks and reduced tumor incidence to<br />

approximately 30%. Expression of IL6, IL1β, TNFα, collagen<br />

I/IV, pSTAT3, and pNFkβ was suppressed to approximately<br />

60% in mice fed an Ath HFD containing peretinoin compared<br />

with mice fed only an Ath HFD. We found peretinoin induced<br />

autophagy that was shown by the increased autophagosome<br />

formation by electron microscopy. We found increased LC3-II<br />

and co-localization of autophagosome (LC3-II) and lysosome<br />

(Lamp2) by immune fluorescent staining. Among various autophagy<br />

related factors, peretinoin increased the expression of<br />

Atg16L1 and promoted Atg16L1-dependent autophagy. In primary<br />

hepatocyte, palmitic acid and IL6 suppressed the expression<br />

of Atg16L1, while peretinoin rescued the Atg16L1 and<br />

suppressed pSTAT3 and pNFkβ. We found peretinoin induced<br />

the expression of CEBPα and CEBPα activated the promoter<br />

activity of Atg16L1. Conclusion Peretinoin suppresses the development<br />

of liver steatosis, inflammation, and tumorigenesis by<br />

activating Atg16L1-dependent autophagy. These results support<br />

the clinical efficacy of peretinoin for preventing HCC.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Masao Honda, Kai Takegoshi,<br />

Taro Yamashita, Naoto Nagata, Toshinari Takamura, Takuji Tanaka<br />

145<br />

Effects of norfloxacin therapy on survival in patients<br />

with Child-Pugh class C cirrhosis: Results of a randomized,<br />

double-blind, placebo-controlled, multicenter trial<br />

Richard Moreau 1,2 , Laure Elkrief 2 , Christophe Bureau 3 , Jean-<br />

Marc Perarnau 4 , Thierry Thevenot 5 , Faouzi Saliba 6 , Isabelle Ollivier-Hourmand<br />

15 , Alexandre Louvet 7 , Pierre Nahon 8 , Frédéric<br />

Oberti 9 , Rodolphe Anty 10 , Sophie Hillaire 11 , Blandine Pasquet 12 ,<br />

Violaine Ozenne 13 , Marika Rudler 14 , Marie Angele Robic 3 , Louis<br />

d’Alteroche 4 , Vincent Di Martino 5 , Pierre-Emmanuel Rautou 2 , Nathalie<br />

Gault 16 , Didier Lebrec 1,2 ; 1 UMR S1149, Center for Research<br />

on Inflammation (CRI), Inserm and Paris Diderot University, Clichy,<br />

France; 2 DHU UNITY, Service d’hépatologie, Hôpital Beaujon,<br />

APHP, Clichy, France; 3 Service d’hépato-gastroentérologie,, Hôpital<br />

Purpan CHU Toulouse et Université Paul Sabatier, Toulouse,<br />

France; 4 Unité d’Hépatologie, Hépatogastroentérologie, CHRU<br />

Tours, Tours, France; 5 Service d’Hépatologie et de Soins Intensifs<br />

Digestifs, Hôpital Jean Minjoz, Besançon, France; 6 Centre<br />

Hépato-Biliaire, Hôpital Paul Brousse, APHP, Villejuif, France;<br />

7 Service des maladies de l’appareil digestif, Hôpital Huriez, Lille,<br />

France; 8 Service d’Hépatologie, Hôpital Jean Verdier, APHP,<br />

Bondy, France; 9 Service d’hépato-gastroentérologie, CHU d’Angers,<br />

Angers, France; 10 Pôle Digestif and Inserm 1065, CHU<br />

de Nice, Nice, France; 11 Service de Médecine Interne, Hôpital<br />

Foch, Suresnes, France; 12 Unité de recherche clinique Paris Nord,<br />

Hôpital Bichat, APHP, Paris, France; 13 Hépato-gastroentérologie,<br />

Hôpital Lariboisière, APHP, Paris, France; 14 Hépatologie, GH<br />

Pitié-Salpêtrière, APHP, Paris, France; 15 Service d’hépato-gastro-entérologie<br />

et nutrition and Unité 1930, CHU Côte de Nacre,<br />

Caen, France; 16 Unité de recherche clinique Paris Nord, Hôpital<br />

Beaujon, APHP, Clichy, France<br />

Background & Aims: Whether long-term oral antibiotic therapy<br />

using a fluoroquinolone should be used to improve prognosis in<br />

patients with advanced cirrhosis is debated. Thus we addressed<br />

this question by investigating the effects of the administration<br />

of norfloxacin (a fluoroquinolone) in patients with advanced<br />

cirrhosis. Methods: We performed a multicenter, double-blind<br />

trial in which we randomly assigned patients with Child-Pugh<br />

class C cirrhosis to once-daily 400 mg oral norfloxacin (144<br />

patients) or placebo (147 patients) for 6 months. Surviving<br />

patients were followed-up for further 6 months after treatment<br />

cessation. The primary end point was overall survival at 6<br />

months; secondary end points included overall survival at 3<br />

and 12 months, and liver-related complications (bacterial infec-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 283A<br />

tion, kidney failure, encephalopathy, variceal hemorrhage).<br />

Patients were censored at the time of liver transplantation or<br />

spontaneous bacterial peritonitis. Results: 291 patients were<br />

enrolled rather than the projected 392 because of slow recruitment<br />

and termination of funding. Excessive alcohol consumption<br />

was the cause of cirrhosis in 115 patients (79.9%) of the<br />

norfloxacin group and 108 patients (73.5%) of the placebo<br />

group. By 6 months, 19 deaths (13.2%) occurred in the norfloxacin<br />

group as compared with 27 deaths (18.4%) in the<br />

placebo group, with a hazard ratio (HR) for death at 6 months<br />

of 0.69 (95% confidence interval [CI], 0.38 to 1.24; P=0.21).<br />

The HR for death at 3 months and 12 months were 0.53 (95%<br />

CI, 0.28 to 1.02; P=0.05) and 0.76 (0.48 to 1.22; P=0.26),<br />

respectively. There were no significant differences in terms of<br />

6-month rates of infections (20.8% vs. 22.4%), kidney failure<br />

(6.9% vs. 7.5%), encephalopathy (16.7% vs. 19.0%), variceal<br />

hemorrhage (3.5% vs. 4.1%). Post-hoc analysis showed that at<br />

enrollment, there were 56 patients (39.2%) in the norfloxacin<br />

group and 63 (42.9%) in the placebo group who had histological<br />

features of alcoholic hepatitis (P=0.52). In this subgroup<br />

of patients, the HR for death at 3 months, 6 months and 12<br />

months were 0.32 (95% CI, 0.12 to 0.86; P=0.01), 0.49<br />

(95% CI, 0.21 to 1.15; P=0.09) and 0.47 (95% CI, 0.23 to<br />

0.97; P=0.03), respectively. Conclusions: In this study, 6-month<br />

oral norfloxacin therapy did not improve overall survival and<br />

did not affect the development of liver-related complications<br />

in patients with Child-Pugh class C cirrhosis. In the subgroup<br />

of patients with alcoholic hepatitis, norfloxacin therapy may<br />

increase 3-month overall survival suggesting early beneficial<br />

effects of the antibiotic in this subgroup of patients.<br />

Disclosures:<br />

Christophe Bureau - Speaking and Teaching: Gore<br />

Jean-Marc Perarnau - Consulting: Gore and Associates<br />

Faouzi Saliba - Advisory Committees or Review Panels: Novartis, Roche, Genzyme,<br />

Vital therapies; Grant/Research Support: Astellas; Speaking and Teaching:<br />

Gambro, MSD, Gilead<br />

Isabelle Ollivier-Hourmand - Speaking and Teaching: gilead, jansen, bayer<br />

Vincent Di Martino - Advisory Committees or Review Panels: Gilead, France,<br />

Abbvie, BMS France; Board Membership: MSD France; Consulting: Gilead,<br />

France; Speaking and Teaching: Janssen, BMS France, Gilead France<br />

Pierre-Emmanuel Rautou - Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Richard Moreau, Laure Elkrief,<br />

Thierry Thevenot, Alexandre Louvet, Pierre Nahon, Frédéric Oberti, Rodolphe<br />

Anty, Sophie Hillaire, Blandine Pasquet, Violaine Ozenne, Marika Rudler, Marie<br />

Angele Robic, Louis d’Alteroche, Nathalie Gault, Didier Lebrec<br />

146<br />

Ischemic hepatitis following acute variceal bleed is<br />

ameliorated by N- acetyl cysteine (NAC) in cirrhotics: A<br />

prospective randomized controlled trial<br />

Avinash Kumar, Ankur Jindal, Rakhi Maiwall, Lovkesh Anand, Amrish<br />

Sahney, Shiv K. Sarin; Institute Of Liver and Biliary Sciences,<br />

New Delhi, India<br />

Background & aims. Ischemic hepatitis (IH) following acute<br />

variceal bleed (AVB) in cirrhotics carries ominous prognosis.<br />

Consequent hypotension & hepatic ischemia may result in centrilobular<br />

necrosis . N-acetylcysteine (NAC), a potent anti-oxidant<br />

may prevent cell injury & improve tissue oxygen delivery.<br />

We investigated the efficacy & safety of NAC in the prevention<br />

of IH & survival. Methods: Of 530 consecutive patients<br />

admitted with UGI bleed (Aug 2013-Apr 2015), cirrhotics with<br />

acute variceal bleed (n=214) were randomized in a prospective<br />

open label study (NCT02015403) to receive either standard<br />

of care (SOC; iv fluids, terlipressin, endoscopic variceal<br />

ligation, blood/blood products & antibiotics) plus NAC (150<br />

mg/kg/hr for 1 h followed by 12.5mg/kg/hr for 4 h, followed<br />

by 6.25 mg/kg for 67 hrs) [Group A(GrA), n=107) or<br />

SOC alone [Group B(GrB),n=107]. Patients were monitored<br />

for presence & severity of IH (increase in AST of ≥5 X ULN &<br />

bilirubin ≥1.5 X ULN in


284A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

patients, 50(4.9%) were found to have definite HP; 4 were<br />

excluded (2 on L-dopa and 2 on bromocriptine).The patients<br />

in Gr A and B were comparable in age (54.8± 6.7 vs., 54.3<br />

±7.4 yr, p=0.7) gender(Males 86%vs. 83%, p= 0.5),etiology<br />

of cirrhosis, MELD (17.9 ± 3.8vs. 18.2 ± 4.2,p=0.3),<br />

baseline arterial ammonia(131 ± 7.4vs. 132 ± 7.9mmol/l,<br />

p=.4), number of patients with recurrent HE: 13/22(59%)<br />

vs. 14/24(58%), p=0.7,persistent HE: 8/22 (36%) vs. 9/24<br />

(37.5%), p=0.6, and porto-systemic shunts: 17/21(80) vs.<br />

19/24(79), p= 0.6. Complete &partial response was seen in<br />

(none vs. 7 (29%), p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 285A<br />

on the underlying mechanisms of action of taurine and further<br />

long-term evaluation in cirrhosis and portal hypertension are<br />

warranted.<br />

Disclosures:<br />

Mattias Mandorfer - Consulting: Janssen; Speaking and Teaching: AbbVie, Gilead,<br />

Janssen, Boehringer Ingelheim, Bristol-Myers Squibb, Roche<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

The following authors have nothing to disclose: Remy Schwarzer, Rafael Paternostro,<br />

Birgit B. Heinisch, Philipp Schwabl, Arnulf Ferlitsch<br />

150<br />

Anticoagulation in patients with cirrhosis and portal<br />

vein thrombosis is associated with increased portal vein<br />

recanalization and better prognosis.<br />

Carlos Noronha Ferreira 1 , Teresa Rodrigues 2 , Ana Júlia Pedro 3 ,<br />

Paula Ferreira 1 , Margarida Sobral Dias 1 , Afonso Gonçalves 4 ,<br />

Leonor Xavier Brito 1 , Fatima Serejo 1 , Rui T. Marinho 1 , Cilénia<br />

B. Costa 1 , Narcisa Fatela 1 , Helena Cortez-Pinto 1 , Fernando<br />

Ramalho 1 , Paula Alexandrino 1 , José F. Velosa 1 ; 1 Serviço de Gastrenterologia<br />

e Hepatologia, Hospital de Santa Maria, Centro<br />

Hospitalar Lisboa Norte, Lisboa, Portugal; 2 Laboratório de Biomatemática,<br />

Faculdade de Medicina de Lisboa, Lisboa, Portugal;<br />

3 Serviço de Medicina 2, Hospital de Santa Maria - Centro Hospitalar<br />

Lisboa Norte, Lisboa, Portugal; 4 Serviço de Imagiologia,<br />

Hospital de Santa Maria, Centro Hospitalar Lisboa Norte, Lisboa,<br />

Portugal<br />

Introduction: Cirrhosis is recognized as a prothrombotic state.<br />

A recent study showed that prophylactic anticoagulation prevented<br />

portal vein thrombosis (PVT) and decreased episodes<br />

of decompensation of cirrhosis. Aims: To analyze the effect<br />

of anticoagulation on recanalization of non-tumoral PVT in<br />

patients with cirrhosis and its effect on prognosis. Methods: 69<br />

consecutive patients with cirrhosis diagnosed with non-tumoral<br />

PVT were studied. The clinical features at diagnosis of PVT and<br />

factors associated with anticoagulation use were studied. Decision<br />

to start anticoagulation was taken at the discretion of the<br />

clinician managing the patient. The effect of anticoagulation<br />

on PVT recanalization and mortality was analyzed. Results:<br />

The average age was 58.6±11.8 years and 44(64%) were<br />

males. Severity of cirrhosis: Median(Range) Child–Pugh(CP)<br />

score: 8(5-15), MELD score:13(6-35). CP class: A-15(22%),<br />

B-32(46%), C-22(32%). At diagnosis of PVT, 55(80%) were<br />

symptomatic. Variceal bleeding(VB) in 30(46%) and abdominal<br />

pain in 19(29%) were the main clinical presentations.<br />

Anticoagulation (LMWH–9, warfarin–16) was administered in<br />

25(36%) patients one of whom with cavernoma. Patients with<br />

VB were less likely to be given anticoagulation (p=0.037).<br />

There were no differences in age, gender, etiology, severity of<br />

cirrhosis and extent of PVT in patients receiving, or not, anticoagulation.<br />

Recanalization of PVT was assessed by at least one<br />

imaging study in 60 patients and recanalization (Total–13, partial<br />

– 9) of the portal vein was documented in 22(37%) patients.<br />

Median (Range) follow–up was 21(0-376) months. At the end<br />

of follow-up, 29(42%) patients died, of which sixteen deaths<br />

were related to infectious complications with no deaths due to<br />

anticoagulation related bleeding. By Cox regression analysis,<br />

factors associated with mortality at the end of follow-up were:<br />

Age (HR 1.040, 95% C.I. 1.002–1.078, p=0.037), CP score<br />

(HR 1.35, 95% C.I. 1.18–1.55, p


286A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Additional resources from government for HCV could aid in the<br />

elimination of HCV from the United States.<br />

Cost of HCV in the era of DAAs<br />

($1.5 billion USD), shielding x-ray components ($13.2 billion<br />

USD), and reducing greenhouse gas emissions ($413 billion<br />

USD). Conclusions: The scope of benefit conferred by new<br />

HCV therapies is large due to the high mortality risk of HCV,<br />

the curative effect of therapy and the size of the population<br />

infected. However, at current prices the incremental cost of<br />

therapy represents a 3478-fold increase in current nationwide<br />

spending on hepatitis C, more than 11% of all national health<br />

expenditures, and 11 times the annual NIH budget. Concurrent<br />

efforts to increase access to treatment as well as reduce drug<br />

costs should be prioritized.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mai T. Pho, Andrew I. Aronsohn,<br />

Benjamin P. Linas, Daniel C. Johnson, Jake R. Morgan, William V. Padula, David<br />

O. Meltzer<br />

Disclosures:<br />

Jagpreet Chhatwal - Consulting: Merck & Co., Inc., Gilead, Complete HEOR<br />

Solutions; Grant/Research Support: NIH/National Center for Advancing Translational<br />

Sciences<br />

The following authors have nothing to disclose: Xiaojie Wang, Mark S. Roberts,<br />

Mina Kabiri, Turgay Ayer, Julie M. Donohue, Fasiha Kanwal<br />

152<br />

Treating Hepatitis C in the US: measuring impact and<br />

value in the context of other major health interventions<br />

Mai T. Pho 1 , Andrew I. Aronsohn 1 , Benjamin P. Linas 2 , Daniel<br />

C. Johnson 1 , Jake R. Morgan 2 , William V. Padula 1 , David O.<br />

Meltzer 1 ; 1 Medicine, University of Chicago, Chicago, IL; 2 Boston<br />

Medical Center, Boston, MA<br />

Background: Hepatitis C (HCV) results in significant morbidity<br />

and mortality in the US. All-oral therapies, while curative, are<br />

costly. We estimate the economic value of providing treatment<br />

for individuals with HCV relative to other major interventions<br />

such as decreasing greenhouse gas emissions. Methods: We<br />

calculate the net monetary benefit (NMB) of therapy provided<br />

for all individuals estimated to be infected with HCV in the US<br />

using a value of statistical life (VSL) and value of statistical life<br />

year (VSLY) framework. A VSL of 5 million US dollars (USD) is<br />

used, and we calculate the VSLY as the quotient of the VSL and<br />

the discounted expected remaining life-years for the general<br />

population. We estimate the total population of HCV based<br />

on prevalence, demographic and genotype data from the<br />

National Health and Nutrition Evaluation Survey from 2003-<br />

2010. The NMB of therapy is the product of VSLY and the<br />

remaining discounted life-years associated with treatment and<br />

no treatment of the total population, subtracted by the incremental<br />

cost. Regimens examined included 12 weeks of sofosbuvir<br />

(SOF) and ledipasvir (LDV) for genotype (GT)1, 12-16<br />

weeks of SOF and ribavirin (RBV) for GT2, and 24 weeks of<br />

SOF and RBV for GT3 infections. Remaining life expectancy<br />

and costs are projected using the Hepatitis C Virus Cost-Effectiveness<br />

(HEP-CE) simulation model. Prices for HCV drugs are<br />

varied from 100% to 55% to reflect offsets due to rebates and<br />

other discounts. Benefits and costs are reported in the context<br />

of federally regulated health interventions. Results: Compared<br />

to no treatment, treating 2.7 million individuals with HCV will<br />

result in 9.5 million life-years gained at an incremental cost of<br />

$347 billion USD. The NMB of HCV therapy is $2.2 trillion<br />

USD. When the price of medications is discounted by 45%,<br />

the NMB of therapy is $2.4 trillion USD, with an incremental<br />

cost of therapy of $187 billion USD. The NMB of HCV<br />

treatment exceeds influenza vaccinations at long-term facilities<br />

153<br />

Hepatitis C (HCV) Infection in Baby Boomers Are Independently<br />

Associated with Mortality and Resource Utilization<br />

Mehmet Sayiner 2 , Mark Wymer 2 , Pegah Golabi 2 , Joel S. Ford 3 ,<br />

Indie Srishord 2 , Zobair M. Younossi 2,1 ; 1 Center For Liver Disease,<br />

Department of Medicine, Inova Fairfax Hospital, Falls Church,<br />

VA; 2 Betty and Guy Beatty Center for Integrated Research, Inova<br />

Health System, Falls Church, VA; 3 Department of Medicine, Inova<br />

Health System, Falls Church, VA<br />

Background and aim: HCV is more common among Baby<br />

boomers (BB) (born 1946-1965). As this cohort age, they will<br />

increasingly become Medicare eligible. Our aim was to evaluate<br />

resource utilization and short term mortality of BB-Medicare<br />

recipients with HCV. Method: We used inpatient and outpatient<br />

Medicare databases (2005-2010) with clinical data,<br />

admission and discharge information, ICD-9 codes, procedure<br />

codes, and billing information. Medicare patients who were<br />

considered BB were identified using all HCV-related ICD-9<br />

codes. Charlson Index was calculated as a measure of comorbidity.<br />

The outcome measures included resource utilizations<br />

[payment per case and inpatient length of stay (LOS)] and short<br />

term mortality (90 days). Results: We included 1,153,862<br />

BB who received Medicare services from 2005 to 2010. Of<br />

these, 3.2% (N=37,365) were HCV (+). During this period,<br />

Medicare-BB in the inpatient setting increased from 39,793 to<br />

55,235, and their claims increased from 78,924 to 106,232.<br />

During the study period, both overall mortality [8.94% to<br />

10.25% (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 287A<br />

category of 45-49 (4.21%, [3.14-5.28]), ESRD (966.31%,<br />

[954.86-977.88]), disabled status (43.22%, [41.67-44.80]),<br />

Charlson score (46.78%, [46.31-47.26]), and year study<br />

(2.72%, (2.58-2.85]) were all independently associated with<br />

an increase in total payments. Conclusions: Prevalence of HCV<br />

is higher in Baby boomers Medicare recipients. For the entire<br />

BB cohort, HCV positivity is independently associated with<br />

higher mortality and resource utilization.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Mehmet Sayiner, Mark Wymer,<br />

Pegah Golabi, Joel S. Ford, Indie Srishord<br />

154<br />

Bariatric surgery for the treatment of nonalcoholic steatohepatitis:<br />

Results of a Markov Model<br />

Matthew Klebanoff 1,2 , Kathleen E. Corey 2,3 , Lee M. Kaplan 2,3 ,<br />

Raymond T. Chung 2,3 , Chin Hur 2,3 ; 1 Institute for Technology<br />

Assessment, Massachusetts General Hospital, Boston, MA; 2 Gastrointestinal<br />

Unit, Massachusetts General Hospital, Boston, MA;<br />

3 Harvard Medical School, Boston, MA<br />

Obese individuals are at heightened risk of developing nonalcoholic<br />

steatohepatitis (NASH), which can give rise to cirrhosis<br />

and hepatocellular carcinoma. Current treatments for NASH<br />

are poorly tolerated or have little long-term data to support their<br />

use. A growing body of evidence suggests that bariatric surgery<br />

not only leads to weight loss but may also improve NASH.<br />

Using a Markov model, we determined the gain in life years<br />

(LYs) and quality-adjusted life years (QALYs) following laparoscopic<br />

Roux-en-Y gastric bypass (L-RYGB) for patients in various<br />

weight classes with varying degrees of NASH fibrosis (F0-F3).<br />

After surgery, a percentage of patients who benefit (69.7% in<br />

base-case, from Mathurin et al., 2009) enter a ‘NASH remission’<br />

state and stop progressing toward cirrhosis, although<br />

they may later relapse (Fig. 1). The primary analysis does not<br />

model surgery in F4 patients due to a lack of data in this group.<br />

The model incorporated life year and quality-of-life decrements<br />

associated with L-RYGB and complications. Extensive sensitivity<br />

analyses examined the impact of model input uncertainty on<br />

results. For all classes of obesity (mild, moderate and severe),<br />

surgery led to a gain in LYs and QALYs. When we excluded<br />

benefits of weight loss to explore the potential impact of surgery<br />

on non-obese NASH patients, surgery reduced life expectancy<br />

by 0.02 LYs for F0 patients, but increased LYs for F1-F3<br />

patients. Surgery also decreased quality-adjusted survival for<br />

F0 and F1 patients by 0.28 and 0.22 QALYs, respectively,<br />

but increased QALYs for F2 and F3 patients. The number of<br />

F3 patients needed to treat (NNT) to prevent one liver-related<br />

death was seven; for F2 patients, the NNT was 17. Conclusions:<br />

Bariatric surgery is effective as a treatment for patients<br />

with NASH in all obesity classes. When we focused only on the<br />

benefit due to NASH remission by eliminating any weight loss<br />

benefit from the model, bariatric surgery improved life expectancy<br />

for F1-F3 patients, but increased quality-adjusted survival<br />

only for patients with more advanced fibrosis (F2 and F3), with<br />

relatively favorable NNTs.<br />

Disclosures:<br />

Kathleen E. Corey - Advisory Committees or Review Panels: Gilead; Speaking<br />

and Teaching: Synageva<br />

Lee M. Kaplan - Consulting: Johnson and Johnson, GI Dynamics, Novo Nordisk;<br />

Grant/Research Support: Ethicon<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Matthew Klebanoff, Chin Hur<br />

155<br />

A Trial-based Model of Liver Transplant and Liver-related<br />

Death in Patients with Primary Biliary Cirrhosis<br />

(PBC)<br />

Marco Carbone 1 , Richard Pencek 2 , Tracy J. Mayne 2 , Tonya Marmon<br />

2 , George F. Mells 3 , David Shapiro 2 ; 1 Division of Gastroenterology<br />

and Hepatology, Department of Medicine, Addenbrooke’s<br />

Hospital, Cambridge, United Kingdom; 2 Intercept Pharmaceuticals,<br />

Inc., San Diego, CA; 3 Division of Gastroenterology and<br />

Hepatology, Department of Medicine, University of Cambridge,<br />

Cambridge, United Kingdom<br />

Background: The UK-PBC research group derived and validated<br />

a predictive model for liver transplant or liver-related<br />

death in PBC patients based on Cox proportional hazards<br />

regression analysis of the UK-PBC cohort (n=4022). Predictors<br />

included ALP, bilirubin, ALT, albumin and platelet count. Obeticholic<br />

acid (OCA) is a selective FXR agonist in development for<br />

the treatment of PBC. In a recent Phase 3 trial (POISE), OCA<br />

treatment was associated with significant improvements in ALP,<br />

bilirubin and transaminases compared to placebo +UDCA.<br />

Objective: To apply the UK-PBC risk algorithm to the POISE<br />

trial data to predict effect of OCA on liver transplant and liver-related<br />

death. Methods: In POISE, 216 PBC patients with<br />

inadequate response or intolerance to UDCA were randomly<br />

assigned to: OCA 10 mg, OCA titration (starting at 5 mg and<br />

adjusted to 10 mg at 6 months based on clinical response) or<br />

placebo; pre-trial UDCA continued. Treatment effect on liver<br />

biochemistry was assessed at 12 months. The UK-PBC algorithm<br />

assessed risk for liver-transplant or liver-related death at<br />

5, 10 and 15 years based on 12-month change from baseline.<br />

Results: The demographics of the POISE cohort were: mean age<br />

56 years, 91% female, 94% white, 93% on UDCA. At baseline,<br />

the UK-PBC algorithm indicated that placebo patients were<br />

at slightly higher risk for events compared to both OCA arms.<br />

At 12 months, predicted risk had significantly improved with<br />

OCA, primarily due to improved ALP and bilirubin. Risk worsened<br />

in the placebo arm, primarily due to increased bilirubin.<br />

Predicted risk at 5, 10 and 15 years is shown below. Based on<br />

absolute risk reduction at 15 years, the number needed to treat<br />

with OCA to avoid liver transplant or liver-related death was<br />

13 (OCA 10 mg) and 11 (OCA titration). Conclusions: The<br />

UKPBC risk model is a validated prognostic indicator of liver<br />

transplant-free survival. When applied to patients in POISE,


288A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

the UK-PBC algorithm predicts a significant decrease in risk for<br />

liver transplant and liver-related death in patients treated with<br />

OCA+UDCA relative to placebo+UDCA. OCA has the potential<br />

to be an important new treatment option for PBC patients<br />

unable to achieve adequate response with UDCA.<br />

Disclosures:<br />

Richard Pencek - Employment: Intercept Pharmaceuticals; Stock Shareholder:<br />

Intercept Pharmaceuticals<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

Tonya Marmon - Employment: Intercept Pharmaceuticals, Inc; Stock Shareholder:<br />

Intercept Pharmaceuticals, Inc<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: Marco Carbone, George F. Mells<br />

156<br />

Costs per Diagnosed Liver Disease Stage Among Individuals<br />

with Chronic Hepatitis C Virus in the Veterans<br />

Health Administration<br />

Jennifer R. Kramer 1 , Peter Richardson 1 , Danielle Liffmann 2 , Hashem<br />

B. El-Serag 1 , David B. Rein 2 ; 1 Department of Medicine, Michael E.<br />

DeBakey VA Medical Center and Baylor College of Medicine,<br />

Houston, TX; 2 NORC at the University of Chicago, Chicago, IL<br />

Background: The prevalence of hepatitis C virus (HCV) infection<br />

in the Veteran Health Administration (VHA) is high, but<br />

previous research has not evaluated the costs of medical care<br />

for diagnosed patients. Methods: We used the VHA HCV<br />

Clinical Case Registry and Health Economics Resource Center<br />

average cost data of inpatient, outpatient, and long-term care<br />

to estimate the incremental annual cost of diagnosed HCV.<br />

We compared the total healthcare costs of HCV antibody and<br />

RNA + patients (cases) to HCV antibody + but HCV RNA -<br />

patients (controls) identified from FY 1999 to 2009. Cases<br />

were matched 1:1 to all available controls using a propensity<br />

score that included demographic and clinical characteristics.<br />

Patients were required to have an available AST/platelet ratio<br />

(APRI) value in the index year (year of the earliest RNA test)<br />

and one subsequent year. In each year, we classified patients<br />

using ICD-9 codes and lab data (highest APRI in year) into 8<br />

clinical stages: APRI 2, decompensated<br />

cirrhosis (DCC), hepatocellular carcinoma (HCC),<br />

transplant/post-transplant care, and death while infected with<br />

HCV. Patients with DCC or HCC prior to HCV index date<br />

were excluded. We estimated the incremental annual costs<br />

for the 6 years after index date of each stage of HCV using<br />

a hierarchical longitudinal two-part health expenditure model<br />

(logistic, generalized linear model assuming a gamma distribution<br />

and log link) and adjusted for clustering by incorporating<br />

patient level random effects in both the logistic and gamma<br />

models allowing these effects to be correlated. Results: There<br />

were 157,303 patients (140,169 cases/17,134 controls) in<br />

our source population; 52.8 years old (sd=8.0), 50.6% white,<br />

25.9% Black, 96.7% male. The analysis set included 17,134<br />

cases that were propensity matched to the 17,134 controls.<br />

In 2009, 52.1% had APRI 2, 5.6% DCC, 0.7% HCC, 0.1% transplant,<br />

and 3.9% died. Using 2015 dollars, on average, HCV RNA<br />

- patients with an APRI


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 289A<br />

with the new balloon score. Most showed a stepwise increase<br />

across the range of the score. Conclusion: The new balloon<br />

score shows excellent correlation with clinical disease features.<br />

Replacement of the old balloon score with the new balloon<br />

score in the NAS would give more weight to ballooning as well<br />

as increasing the dynamic range of the NAS.<br />

NASH CRN New Balloon Score<br />

Disclosures:<br />

Elizabeth M. Brunt - Consulting: Synageva; Independent Contractor: Rottapharm<br />

Brent A. Neuschwander-Tetri - Advisory Committees or Review Panels: Nimbus<br />

Therapeutics, Bristol Myers Squibb, Janssen, Mitsubishi Tanabe, Conatus,<br />

Scholar Rock<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: David E. Kleiner, Patricia H. Belt,<br />

Cynthia A. Behling, Ryan M. Gill, Cynthia D. Guy, Mark L. Van Natta<br />

158<br />

Overweight in late adolescence is associated with<br />

decompensated liver disease later in life after 39 years<br />

of follow-up<br />

Hannes Hagström 1 , Per Stål 1 , Rolf W. Hultcrantz 1 , Tomas Hemmingsson<br />

2,3 , Anna Andreasson 4,5 ; 1 Center for Digestive Diseases,<br />

Unit of Hepatology, Karolinska University Hospital. Department<br />

of Medicine, Huddinge, Karolinska Institutet, Stockholm, Sweden,<br />

Stockholm, Sweden; 2 Institute of Environmental Medicine, Karolinska<br />

Institutet, Stockholm, Sweden; 3 Centre for Social Research<br />

on Alcohol and Drugs, Stockholm University, Stockholm, Sweden;<br />

4 Stress Research Institute, Stockholm University, Stockholm, Sweden;<br />

5 Division of Family Medicine, Karolinska Institutet, Stockholm,<br />

Sweden<br />

Background: The prevalence of both obesity and liver diseases<br />

is increasing globally. Obesity is associated with a worse prognosis<br />

in different liver diseases. If overweight and obesity in<br />

late adolescence is an independent predictor of liver disease<br />

later in life is not studied. We investigated if overweight per se<br />

predicts development of liver disease and decompensated liver<br />

disease later in life with a up to 39 year follow-up. Materials<br />

and methods: Data from 49,321 men (18-21 years) conscribed<br />

to military service in Sweden between 1969 and 1970 was<br />

used. Conscription during this time was mandatory and >97%<br />

of the male population was available for this study. Weight and<br />

height measured at conscription were used to calculate body<br />

mass index (BMI). Data was collected from the national patient<br />

register to identify any diagnosis of liver disease at a Swedish<br />

hospital from time of conscription until the end of 2009.<br />

A multivariate logistic regression model was used to estimate<br />

odds ratios (OR) of decompensated liver disease (hepatorenal<br />

syndrome [HRS], hepatocellular carcinoma [HCC], ascites,<br />

varices or hepatic encephalopathy [HE]) for persons with BMI<br />

over as compared to under 25. The model was adjusted for<br />

alcohol use, smoking, use of narcotics, social status and blood<br />

pressure at the time of conscription. Results: Mean BMI at conscription<br />

was 20.97 kg/m 2 , where 6.6% had a BMI > 25 and<br />

0.8% had a BMI > 30. During a follow-up period of in mean<br />

37.9 (+/-4.8 years, range 0-39) years of follow-up, 525 persons<br />

were diagnosed with liver disease. Of these, 206 persons<br />

developed decompensated liver disease (18 cases of HRS, 31<br />

HCC, 79 ascites, 124 varices and 7 HE as first manifestation<br />

of decompensation, where 43 persons had multiple diagnoses<br />

at presentation). Mean time to decompensation was 29.8<br />

years (95% Confidence Interval [CI] 28.8-30.8 years). BMI ><br />

25 was significantly associated with development of decompensated<br />

liver disease both in the univariate (OR 2.09, 95%CI<br />

1.38-3.16, p


290A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

than children with Zone 3 steatosis (73% vs 58%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 291A<br />

olites predictive of fibrosis stage. RESULTS: Age, gender, BMI,<br />

current/past alcohol use/smoking, total HDL/LDL cholesterol<br />

or triglycerides did not differ with fibrosis stage. In univariate<br />

testing, plasma concentrations of alpha-glutamyltyrosine N-acetylcitrulline<br />

palmitoyl-palmitoyl-glycerophosphocholine, glycocholate<br />

glutarate, fucose and fumarate associated with fibrosis<br />

stage (p=0.0007). After controlling for age, gender and DM,<br />

N-acetylcitrulline, alpha glutamytyrosine, palmitoyl-palmitoyl-glycerophosphocholine,<br />

lignoceric acid, glycocholate, and<br />

1-arachidonoyl associated with fibrosis (p=0.0009). Multivariate<br />

analysis using logistic regression yielded an Area Under<br />

the Receiver Operating Characteristic (AUROC) of 0.8804<br />

for I-urobilinogen, malate, glutarate (pentanedioate), taurochenodeoxycholate,<br />

3-hydroxyoctanoate, fucose, isoleucine,<br />

palmitoyl-palmitoyl-glycerophosphocholine, N-acetylcitrulline.<br />

Sparse ordinal regression was used to predict fibrosis stage,<br />

resulting in a correlation coefficient of 0.3757 (p=5x10 -8 )<br />

after leave-one-out cross-validation. The model identified a<br />

69 metabolite signature predictive of fibrosis stage. CONCLU-<br />

SIONS: Analysis of serum from NAFLD patients revealed alterations<br />

in bile acids, steroids hormones, branched-chain amino<br />

acid catabolism, TCA cycle metabolism, and mitochondrial<br />

function in stage 3-4 fibrosis. The role of these metabolites as<br />

potential biomarkers for hepatic fibrosis requires validation.<br />

Pathways involved in bile acid and/or steroid hormone metabolism,<br />

branched-chain amino acids, and/or mitochondrial function<br />

may offer novel targets for anti-fibrotic therapies in NAFLD.<br />

Disclosures:<br />

Lauren N. Bell - Employment: Metabolon, Inc.<br />

Regis Perichon - Employment: Metabolon<br />

Manal F. Abdelmalek - Consulting: Islet Sciences; Grant/Research Support:<br />

Tobira, Gilead Sciences, NIH/NIDDK, Synageva, Genfit Pharmaceuticals,<br />

Immuron, Galmed, TaiwanJ Pharma, Intercept, NGM Pharmaceuticals<br />

The following authors have nothing to disclose: Ricardo Henao, Nigar Hasanaliyeva,<br />

Jacob Wulff, Cynthia A. Moylan, James T. Lu, Cynthia D. Guy, Anna Mae<br />

Diehl<br />

162<br />

Beneficial effects of the dual PPAR α-δ agonist, GFT505,<br />

on hepatic and cardiometabolic markers in adult NASH<br />

patients.<br />

Stephen A. Harrison 3 , Arun J. Sanyal 2 , Sven M. Francque 4 , Pierre<br />

Bedossa 5 , Lawrence Serfaty 6 , Manuel Romero-Gomez 7 , Paul<br />

Cales 8 , Manal F. Abdelmalek 9 , Stephen H. Caldwell 10 , Joost<br />

Drenth 11 , Quentin M. Anstee 12 , Dean W. Hum 13 , Rémy Hanf 13 ,<br />

Alice Roudot 13 , Sophie Megnien 13 , Bart Staels 14 , Vlad Ratziu 1 ;<br />

1 Hepatology, Hopital Pitie Salpetriere, Paris, France; 2 Internal<br />

Medicine/Division of Gastroenterology, Hepatology and Nutrition,<br />

Virginia Commonwealth University, Richmond, VA; 3 Gastroenterology,<br />

Brooke Army Medical Center, Fort Worth, TX; 4 Antwerp<br />

Univesrity Hospital, Gastroenterology Hepatology, Antwerp, Belgium;<br />

5 Pathology, Hopital Beaujon, Clichy, France; 6 Hepatology<br />

Department, Hospital Saint-Antoine, Paris, France; 7 Valme University<br />

Hospital, Digestive Diseases, Sevilla, Spain; 8 Hepatology<br />

Department, Centre Hospitalier Universitaire d’ Angers, Angers,<br />

France; 9 Medicine, Duke University, Durham, NC; 10 Hepatology<br />

Department, University of Virginia, Charlotesville, VA; 11 Hepatology<br />

Department, RUNMC, Nijmegen, Netherlands; 12 Institute<br />

of Cellular Medicine, Newcastle University, Newcastle, United<br />

Kingdom; 13 Genfit, Loos, France; 14 INSERM U1011, European<br />

Genomic Institute for Diabetes (EGID), Université Lille 2, Lille,<br />

France<br />

Introduction: NASH patients (pts) have a high prevalence of cardiometabolic<br />

risk factors, hence cardiovascular disease is the<br />

leading cause of death in this population. Having demonstrated<br />

the efficacy of GFT505 on the histological reversal of NASH,<br />

here we evaluated the biochemical and cardiometabolic effects<br />

of the drug in in this large, international, phase 2 randomized,<br />

controlled trial of adult pts with biopsy-proven NASH. Methods:<br />

Pts from the ITT population of the GOLDEN505 trial randomized<br />

to the GFT505 120 mg (GFT120) and placebo (PBO)<br />

arms were compared for treatment effects on plasma lipid,<br />

glycemic, insulin resistance, inflammatory and liver markers.<br />

Results are expressed as effect size vs. placebo (LS mean±SE).<br />

Analyses were conducted for different stages of disease severity<br />

based on the NAFLD Activity Score (NAS) at baseline (NAS 3:<br />

mild, NAS 4-5: moderate, NAS>5: severe), and fibrosis stage<br />

(F0-F1 vs. F2-F3 according to Kleiner et al). Results: Compared<br />

to PBO, GFT120 significantly improved GGT (-29.31±6.36<br />

U/L, p


292A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Vlad Ratziu - Advisory Committees or Review Panels: GalMed, Abbott, Genfit,<br />

Enterome, Gilead; Consulting: Tobira, Intercept, Exalenz, Sanofi-Synthelabo,<br />

Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Sven M. Francque, Pierre<br />

Bedossa, Joost Drenth<br />

163<br />

Osteopontin is involved in chronic HBV infection and<br />

enhances HBV replication and HBsAg secretion.<br />

Sandra Phillips, Jason D. Coombes, Sameer Mistry, Roger Williams,<br />

Wing-Kin Syn, Shilpa Chokshi; Institute of Hepatology,<br />

Foundation for Liver Research, London, United Kingdom<br />

Background/Aims: Hepatitis B virus (HBV) requires host cellular<br />

machinery such as cyclophilins to support its ongoing propagation<br />

(Phillips et al, Gastroenterology 2014). These host proteins<br />

represent ideal candidates for therapeutic interventions<br />

as they are generally expected to have a lower frequency of<br />

drug-resistance and antiviral efficacy across genotypes. We<br />

have also recently described a role for the host protein Osteopontin<br />

(OPN), a pro-fibrogenic downstream effector of the<br />

Hedgehog pathway, in enhancing HCV replication (Choi et al,<br />

Clinical Sciences 2014). Whilst increased levels of blood OPN<br />

have similarly been reported in chronic Hepatitis B (CHB), its<br />

role in viral replication remains unknown. This study aimed<br />

to evaluate the role of OPN in HBV replication, HBsAg secretion<br />

and HBV-driven liver injury. Material and Methods: Stably<br />

(HepG2215), transiently (HUH-7) transfected and infected<br />

(HepaRG) cell lines, producing full HBV virions and HBsAg<br />

particles were cultured over 72h in the presence/absence of<br />

several concentrations of recombinant OPN (recOPN). In addition,<br />

secreted OPN was neutralized using OPN-specific aptamers.<br />

Cells and supernatants were harvested at baseline, 24, 48<br />

and 72 hours. Intracellular OPN mRNA and HBV-DNA levels<br />

were quantitated by Real-Time qPCR. HBsAg levels were measured<br />

by ELISA. OPN levels were also measured by ELISA in<br />

cell culture supernatants and in sera of controls and HBeAg(+)<br />

CHB patients who were either treatment naive or treated with<br />

potent antiviral agents. In addition, expression of OPN was<br />

assessed in explanted livers from healthy and HBV-cirrhotic<br />

patients. Results: Serum levels and intrahepatic expression of<br />

OPN were significantly increased in CHB patients compared<br />

to controls (up to 7 fold; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 293A<br />

165<br />

The HBx-DLEU2 lncRNA complex regulates transcription<br />

from cellular genes and the HBV cccDNA<br />

Francesca Guerrieri 1,2 , Letizia Chiodo 1 , Debora Salerno 1,2 , Safaa<br />

Jeddari 2 , Giancarlo Ruocco 1 , Massimo Levrero 2,1 ; 1 CLNS@SAPI-<br />

ENZA, Istituto Italiano di Tecnologia (IIT), Rome, Italy; 2 Dept of<br />

Internal Medicine, La Sapienza University, Rome, Italy<br />

Background: HBx regulatory protein is required for HBV<br />

cccDNA transcription/viral replication and contributes to HBV<br />

oncogenicity. HBx affects the epigenetic control of HBV viral<br />

chromatin, by preventing HDACs and PMRT1 recruitment onto<br />

the cccDNA, as well as of cellular chromatin, by modulating<br />

the recruitment of chromatin modifying enzymes. We previously<br />

showed that HBx binds to the regulatory regions of several<br />

ncRNAs (75 miRNAs and 34 lncRNAs). DLEU2 lncRNA<br />

overlaps with the autophagy TRIM13 gene in the opposite<br />

orientation. Upregulation of specific DLEU2 splicing variants<br />

correlates with the development of solid and onco-hematolocic<br />

malignancies. Objectives: Aim of this study is to clarify the role<br />

of HBx in the regulation of the DLEU2/TRIM13 transcriptional<br />

unit and their functions. Methods: The anti-HBx ChIPSeq was<br />

performed in mock and HBV-replicating HepG2 cells using an<br />

Illumina NGS platform. Independent ChIPs were analyzed by<br />

TaqMan real-time PCR using lncRNA and gene specific primers.<br />

The nCounter Nanostring technology and real-time RT-PCR<br />

were used to evaluate lncRNAs and target gene expression,<br />

respectively. Specific LNA longRNA GapmeRs (Exiqon) were<br />

used for highly efficient inhibition of DLEU2 lncRNA function.<br />

Results: ChIPSeq analysis of HBx global chromatin recruitment<br />

revealed a specific binding to 34 lncRNAs regulatory<br />

regions. Using a custom Nanostring panel we found that all<br />

putative lncRNAs targets were modulated in HBV-replicating<br />

HepG2 cells. Focusing on DLEU2 lncRNA, we demonstrated<br />

that HBx is able to deregulate both its expression and the overlapping<br />

gene TRIM13. HBx binding to the DLEU2 promoter<br />

affects its epigenetic control, changes DLEU2 splicing profile<br />

and up-regulates the antisense gene TRIM13. Selective degradation<br />

of DLEU2 RNA resulted in a reduced H4 acetylation<br />

on the TRIM13 promoter and a ~50% reduction of TRIM13<br />

expression in HBV replicating HepG2 cells. In silico analysis of<br />

a model HBx protein tertiary structure and DLEU2 tertiary structure<br />

suggested a direct interaction. Using a RIP (RNA Immune<br />

Precipitation) approach we confirmed HBx-DLEU2 interaction<br />

in vivo. Finally, we found that DLEU2 inactivation inhibits the<br />

expression of the HBx targets SREBP2 and miR138 as well as<br />

HBV cccDNA transcription, with a sharp decrease of pgRNA<br />

levels, thus suggesting a functional relevance of the DLEU2-<br />

HBx interaction of HBV replication. Conclusion: HBx binds to<br />

the DLEU2 promoter region to modify its epigenetic regulation<br />

change the DLEU2 splicing profile. HBx forms a functional complex<br />

with DLEU2 and regulates HBV replication.<br />

Disclosures:<br />

Massimo Levrero - Advisory Committees or Review Panels: BMS, Jansen, Gilead,<br />

Tekmira, Galapagos, Medimmune; Speaking and Teaching: MSD, Roche<br />

The following authors have nothing to disclose: Francesca Guerrieri, Letizia Chiodo,<br />

Debora Salerno, Safaa Jeddari, Giancarlo Ruocco<br />

166<br />

Viral Expression and Molecular Profiling of Liver Tissue<br />

and Microdissected Hepatocytes in Hepatitis D Virus<br />

(HDV) - Associated Hepatocellular Carcinoma (HCC)<br />

Patrizia Farci 1 , Ashley B. Tice 1 , Ronald E. Engle 1 , Fausto Zamboni<br />

2 , Marta Melis 1 , Zhifeng Long 3 , Jaime Rodriguez-Canales 4 ,<br />

Jeffrey Hanson 4 , Michael R. Emmert-Buck 4 , David E. Kleiner 4 ,<br />

Giacomo Diaz 5 ; 1 Laboratory of Infectious Diseases, National Inst<br />

of Health, Bethesda, MD; 2 Liver Transplantation Center, Brotzu<br />

Hospital, Cagliari, Italy; 3 Personal Diagnostix, Gaithersburg, MD;<br />

4 Laboratory of Pathology, National Cancer Institute, Bethesda,<br />

MD; 5 Department of Biomedical Sciences, University of Cagliari,<br />

Cagliari, Italy<br />

Although HCC develops in a high proportion of patients<br />

with chronic hepatitis D, there are no data on the molecular<br />

mechanisms of HDV-induced hepatocarcinogenesis. Because<br />

of the vital dependence of HDV on HBV, the role of HDV in<br />

promoting HCC is unknown. We used genomic techniques<br />

to dissect the role of HDV and HBV. Microarray (Affymetrix<br />

Human U133 Plus2) was performed on 33 specimens of<br />

whole liver tissue (WLT, tumor vs. non-tumor) and selected<br />

laser capture-microdissected hepatocytes (LCM, malignant vs.<br />

non-malignant hepatocytes) obtained at liver transplant (LT)<br />

from 5 patients with HDV-HCC; 29 WLT specimens from 7<br />

patients with non-HCC HDV cirrhosis served as controls. A<br />

parallel analysis was performed on WLT and LCM from 11<br />

patients with HBV-HCC without HDV. Microarray data were<br />

analyzed using BRB-Array Tools, with multivariate permutation<br />

F-tests (FDR


294A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

167<br />

Characterisation of the immune profile in Chronic Hepatitis<br />

B, with CyTOF, to identify biomarkers of immune<br />

control following NUC therapy discontinuation<br />

Upkar S. Gill 1 , Laura Rivino 2 , Nina Le Bert 3 , Kamini Kunasegaran<br />

2 , Damien Tan 3 , Sarene Koh 3 , Machteld Van Den Berg 2 , Yang<br />

Cheng 4 , Navjyot K. Hansi 1 , Graham R. Foster 1 , Evan W. Newell<br />

4 , Antonio Bertoletti 2,3 , Patrick T. Kennedy 1 ; 1 Hepatology Unit,<br />

Centre for Immunobiology, Blizard Institute, Barts and The London,<br />

School of Medicine & Dentistry, QMUL, London, United Kingdom;<br />

2 Program Emerging Viral Diseases, Duke-NUS Graduate Medical<br />

School, Singapore, Singapore; 3 Infection & Immunity Program,<br />

Singapore Institute for Clinical Sciences, Agency for Science, Technology<br />

& Research (A*STAR), Singapore, Singapore; 4 Singapore<br />

Immunology Network, Singapore Agency for Science, Technology<br />

& Research (A*STAR), Singapore, Singapore<br />

INTRODUCTION: The absence/functional exhaustion of<br />

HBV-specific T cells is the hallmark of chronic HBV infection<br />

(CHB); conversely a robust functional response of these cells<br />

is associated with viral control. Current therapies are limited<br />

in their ability to restore the functional HBV-specific repertoire.<br />

Consequently, Nucleot(s)ide analogue (NUC) therapy remains<br />

indefinite in the majority. Clinical parameters alone cannot<br />

distinguish in which patients NUC therapy can be safely discontinued,<br />

with durable immune control from those who will<br />

relapse and develop hepatic flares (HF). We studied a cohort<br />

of virally suppressed patients on potent NUC therapy prior to<br />

and after treatment discontinuation to characterise the immune<br />

profile associated with viral control. PATIENTS & METHODS:<br />

PBMC were analysed at 4-weekly intervals prior to and after<br />

NUC discontinuation. The frequency of HBV-specific T cells<br />

was assessed by IFNγ ELISPOT after 10 day expansion in the<br />

presence of HBV peptides spanning the entire viral proteome.<br />

Phenotypic and functional characteristics of T & NK cells were<br />

studied with an in-depth longitudinal analysis of the expression<br />

of >35 markers involved in cell activation, differentiation and<br />

exhaustion, by cytometry time of flight (CyTOF), a novel mass<br />

cytometry technology. Serum cytokine and chemokine levels<br />

(IL-1β, IL-6, TNF-α, IL-10, CXCL-8, CXCL-10) were also measured<br />

using Luminex. RESULTS: Prior to NUC discontinuation,<br />

patients could be divided into 2 groups based on the presence/absence<br />

of HBV-specific T cells. Upon NUC discontinuation,<br />

those patients with detectable frequencies of circulating<br />

HBV-specific T cells were able to control HBV replication and<br />

did not demonstrate HF’s. In these patients HBV-specific T cells<br />

preferentially target HBV polymerase followed by core proteins.<br />

In patients without viral control who developed HF, we<br />

noted a significant increase in serum CXCL10 and IL10 levels,<br />

correlating with ALT. Patients without HF’s were characterised<br />

by increased frequencies of a CD8+ T cell subset expressing<br />

CD56, CD107, CCR5, CD127, CD28, 2B4, KLRG1, IFNγ,<br />

TNFα, MIP1β & IL2. NK cell profiles did not differ depending<br />

on viral rebound/HF upon NUC discontinuation, but all CHB<br />

patients expressed significantly higher levels of the exhaustion<br />

marker, KLRG1 on NK cells, compared with healthy controls.<br />

CONCLUSIONS: Our data suggest that HBV-polymerase<br />

specific T cells represent a potential biomarker that can be<br />

used to predict those patients who will control HBV after NUC<br />

discontinuation. These results further support the concept that<br />

HBV-specific T cells are associated with viral control rather than<br />

contributing to immunopathology.<br />

Disclosures:<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

Patrick T. Kennedy - Grant/Research Support: Roche, Gilead; Speaking and<br />

Teaching: BMS, Roche, Gilead<br />

The following authors have nothing to disclose: Upkar S. Gill, Laura Rivino, Nina<br />

Le Bert, Kamini Kunasegaran, Damien Tan, Sarene Koh, Machteld Van Den Berg,<br />

Yang Cheng, Navjyot K. Hansi, Evan W. Newell, Antonio Bertoletti<br />

168<br />

Hepatitis B virus X protein stimulates HBV replication by<br />

regulating transcription factors associated with DNA or<br />

histone methylation<br />

Naoki Oishi 1,2 , Xuyang Wang 2 , Kazunori Kawaguchi 1,2 , Masao<br />

Honda 1,2 , Seishi Murakami 2 , Shuichi Kaneko 1,2 ; 1 Department of<br />

Gastroenterology, Kanazawa University Hospital, Kanazawa,<br />

Japan; 2 Department of Disease Control and Homeostasis,<br />

Kanazawa University Hospital, Kanazawa, Japan<br />

Background: Hepatitis B virus (HBV), a small enveloped DNA<br />

virus, chronically infects more than 350 million people worldwide<br />

and causes liver diseases, from hepatitis to cirrhosis and<br />

liver cancer. HBx is a multifunctional protein encoded by the<br />

HBV genome that stimulates HBV replication. Previously, we<br />

found that HBx has an important role in stimulating HBV transcription<br />

and replication and that the transcriptional transactivation<br />

function of HBx may be critical for its augmentation<br />

effect on HBV replication. However, the molecular mechanism<br />

of HBx in transcriptional coactivation remains unclear. In this<br />

study, we provide a new insight into the coactivator mechanism<br />

and the possibility of a new treatment target for HBV replication<br />

and HBV-related hepatocarcinogenesis. Methods: We<br />

used a retroviral vector to introduce wild-type HBx (HBxwt) or<br />

empty vector (EV) into HepG2 cells. Gene expression profiling<br />

was performed using Affymetrix GeneChip Human U133A2.0<br />

ver.2.0 arrays according to the manufacturer’s protocol. Unsupervised<br />

hierarchical clustering analysis and class comparison<br />

analysis were performed with BRB-Array Tools software version<br />

4.2.2. Transcription factor analysis was performed using<br />

Ingenuity Pathway Analysis (IPA; Ingenuity Systems) to identify<br />

potential upstream transcription factors (TFs). We used array<br />

data from 244 chronic hepatitis B patients for analyzing clinical<br />

features. Results: Unsupervised hierarchical clustering analysis<br />

of HepG2-HBxwt or HepG2-EV cells revealed that the gene<br />

expression profiles of these cell lines were significantly different.<br />

Class comparison analysis between both cell lines identified<br />

803 HBx-related genes. Six HBx-related activated TFs that<br />

affected HBV replication were identified by IPA, small interfering<br />

RNA, and inhibitor analyses. Three of these six TFs control<br />

HBV replication by modifying DNA or histone methylation. By<br />

analysis of 244 hepatitis B patients, these three TFs were found<br />

to be expressed at a significantly highly level in HBeAg(+)<br />

HBeAb(-) cases than in HBeAg(-)HBeAb(+) cases. Moreover,<br />

these TFs up-regulated the expression of stemness markers<br />

(EpCAM, AFP, and SOX9) and epithelial-mesenchymal transition<br />

markers (ZEB1, ZEB2, and VIM) and were associated with<br />

a poor prognosis of HCC. Conclusions: HBx upregulated three<br />

TFs associated with DNA or histone methylation. Moreover,<br />

HBx stimulates HBV replication and hepatocarcinogenesis by<br />

modifying DNA or histone methylation. Our study suggests that<br />

TFs activated by HBx will be an important therapeutic target<br />

against both virus replication and hepatocarcinogenesis aimed<br />

at the prevention of HCC.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 295A<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Naoki Oishi, Xuyang Wang,<br />

Kazunori Kawaguchi, Masao Honda, Seishi Murakami<br />

169<br />

DNA Methylation Markers for Detection of Extrahepatic<br />

Cholangiocarcinoma: Discovery, Tissue validation, and<br />

Pilot Testing in Biliary Brush Samples<br />

Hassan M. Ghoz 1 , Mohammed M. Aboelsoud 5 , Tracy C. Yab 1 ,<br />

Calise K. Berger 1 , William R. Taylor 1 , Xiaoming Cao 1 , Patrick<br />

H. Foote 1 , Nasra H. Giama 1 , Catherine D. Moser 1 , Douglas W.<br />

Mahoney 4 , Emily G. Barr Fritcher 2 , Benjamin R. Kipp 3 , Thomas C.<br />

Smyrk 3 , Lewis R. Roberts 1 , David Ahlquist 1 , John B. Kisiel 1 ; 1 Division<br />

of Gastroenterology and Hepatology, Mayo Clinic, Rochester,<br />

MN; 2 Molecular Anatomic Pathology, Mayo Clinic, Rochester,<br />

MN; 3 Anatomic Pathology, Mayo Clinic, Rochester, MN; 4 Biomedical<br />

Statistics and Infomatics, Mayo Clinic, Rochester, MN;<br />

5 Memorial Hospital, Pawtucket, RI<br />

Background & Aims: Cholangiocarcinoma (CCA) has poor<br />

prognosis due to late stage presentation. Molecular markers<br />

may offer improved accuracy for early detection. We have<br />

identified discriminant DNA methylation markers for intrahepatic<br />

CCA by next-generation sequencing. We now explore<br />

this discovery approach further with extrahepatic CCA (eCCA)<br />

and pilot best markers on brush cytology specimens. Methods:<br />

Reduced-representation bisulfite sequencing (RRBS) was performed<br />

on DNA extracted from 18 frozen eCCA tissue samples<br />

and matched, adjacent benign biliary epithelia or liver parenchyma.<br />

Differentially methylated regions (DMRs) with at least<br />

3 CpGs were ranked by area under the receiver operating<br />

characteristics curve (AUC) & by tumor:normal ratio and then<br />

technically validated by methylation specific PCR (MSP) on<br />

DNA from same samples. Best DMRs were selected for biological<br />

validation on DNA from independent tissues comprising<br />

15 eCCA cases and 60 controls (6 adjacent bile duct, 18<br />

adjacent liver, 18 white blood cell samples, 18 normal colon<br />

epithelia) using MSP. Biologically valid DMRs were then blindly<br />

assayed on DNA extracted from independent archival biliary<br />

brushing specimens including 14 perihilar (pCCA) & 4 distal<br />

(dCCA) cases and 18 matched cytology-negative controls<br />

(CTRL), 4 of which had primary sclerosing cholangitis (CTRL-<br />

PSC). Results: From 5.5 million CpGs, 3674 significant DMRs<br />

were mapped; 46 were selected for technical validation from<br />

which 18 DMRs had an AUC of 0.75–1.0. In biological validation,<br />

8 of these showed an AUC >0.75 in eCCA tissues. In<br />

brushings, methylated EMX1, HOXA1, VSTM2B.764, KC01,<br />

BMP3, SALL1, PTGDR, and RYR2, showed sensitivities of 100%<br />

89%, 83%, 78%, 72%, 72%, 72%, and 72%, respectively,<br />

at 90% specificity. Accuracy of EMX1 on brushings is shown<br />

(figure). Conclusion: Whole-methylome discovery by next-generation<br />

DNA sequencing yielded novel, highly-discriminant<br />

methylation markers for eCCA. Results were validated in independent<br />

tissues as well as cytology brushings. Further clinical<br />

evaluation is clearly warranted.<br />

Disclosures:<br />

Tracy C. Yab - Patent Held/Filed: Exact Sciences; Stock Shareholder: Exact<br />

Sciences<br />

William R. Taylor - Patent Held/Filed: Exact Sciences; Stock Shareholder: Exact<br />

Sciences<br />

Douglas W. Mahoney - Patent Held/Filed: Exact Sciences<br />

Benjamin R. Kipp - Grant/Research Support: Abbott Molecular Inc.<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

David Ahlquist - Advisory Committees or Review Panels: exact sciences; Consulting:<br />

exact sciences; Grant/Research Support: exact sciences; Stock Shareholder:<br />

exact sciences<br />

John B. Kisiel - Grant/Research Support: Exact Sciences Corporation<br />

The following authors have nothing to disclose: Hassan M. Ghoz, Mohammed M.<br />

Aboelsoud, Calise K. Berger, Xiaoming Cao, Patrick H. Foote, Nasra H. Giama,<br />

Catherine D. Moser, Emily G. Barr Fritcher, Thomas C. Smyrk<br />

170<br />

Circulating Osteopontin and Prediction of Hepatocellular<br />

Carcinoma Development in a Large European Population<br />

Talita Duarte-Salles 2 , Sandeep Misra 1 , Magdalena Stepien 2 ,<br />

Mazda Jenab 2 , Pierre Hainaut 2 , Laura Beretta 1 ; 1 Molecular and<br />

Cellular Oncology, MD Anderson Cancer Center, Houston, TX;<br />

2 International Agency for Research on Cancer (IARC-WHO), Lyon,<br />

France<br />

Background: Prevention and early detection strategies are<br />

urgently needed to reduce the burden and mortality of hepatocellular<br />

carcinoma (HCC), the second leading cause of cancer<br />

death worldwide. We previously identified osteopontin<br />

(OPN) as a promising marker for the early detection of HCC.<br />

In this study, we investigated the association between pre-diagnostic<br />

circulating OPN levels and HCC incidence and OPN<br />

performance for early detection of HCC in a large population-based<br />

cohort. Methods: A nested-case control study was<br />

conducted within the European Prospective Investigation into<br />

Cancer and Nutrition (EPIC) cohort study, a large longitudinal<br />

cohort of >520,000 participants from 10 Western European<br />

countries. During a mean follow-up of 4.8 years, 100 HCC<br />

cases were identified. Each case was matched to two controls<br />

by incidence density sampling. OPN levels were measured in<br />

baseline plasma samples. Hepatitis infection status and other<br />

relevant biomarkers such as alpha-fetoprotein (AFP) were also<br />

measured. Conditional logistic regression models were used to<br />

calculate multivariable odds ratio (OR) and 95% confidence<br />

intervals (95%CI) for tertiles and continuous OPN levels in relation<br />

to HCC. Receiver operating characteristics curves (ROC)<br />

were constructed to determine the discriminatory accuracy of<br />

OPN alone or in combination with other liver function biomarkers<br />

in the prediction of HCC. Behavior of OPN in relation to<br />

time of HCC diagnosis was also evaluated. Results: Circulating<br />

OPN level was significantly associated with increased HCC<br />

risk (per 10% increment in OPN level, OR multivariable<br />

=1.30;


296A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

95%CI: 1.14-1.48), also among hepatitis-free participants<br />

(OR multivariable<br />

=1.19; 95%CI: 1.05-1.34). Adding liver function<br />

tests to OPN improved the discrimination of those subjects who<br />

developed HCC (AUC=0.86) while AFP didn’t improve the<br />

model. In contrast, within two years of HCC diagnosis, the<br />

combination of OPN and AFP was best able to predict HCC<br />

risk (AUC=0.88) while further adjustment for liver function tests<br />

or chronic hepatitis B/C infection did not improve the model.<br />

Conclusions: In this study, pre-diagnostic serum OPN concentration<br />

was associated with higher risk of first incident HCC<br />

and the association was stronger among cases diagnosed<br />

during the first two years of follow-up. OPN in combination<br />

with liver function tests improved the detection of HCC in this<br />

low-risk population. However, OPN in combination with AFP<br />

was the best predictive model for detection of HCC within<br />

two years of follow-up and did not change substantially after<br />

excluding hepatitis positive participants, or after adjustment for<br />

biomarkers of liver function.<br />

Disclosures:<br />

The following authors have nothing to disclose: Talita Duarte-Salles, Sandeep<br />

Misra, Magdalena Stepien, Mazda Jenab, Pierre Hainaut, Laura Beretta<br />

171<br />

Risk factors of complications in liver adenomatosis.<br />

Barbara Willandt 1 , Ragna Vanslembrouck 2 , Raymond Aerts 3 ,<br />

Baki Topal 3 , Wim Laleman 1 , David Cassiman 1 , Schalk Van Der<br />

Merwe 1 , Werner Van Steenbergen 1 , Chris Verslype 1 , Frederik<br />

Nevens 1 ; 1 Hepatology, UZ Leuven, Leuven, Belgium; 2 Radiology,<br />

UZ Leuven, Leuven, Belgium; 3 Adominal Surgery, UZ Leuven, Leuven,<br />

Belgium<br />

Introduction & aims: Hepatic adenomatosis is a disorder characterized<br />

by multiple adenomas in an otherwise normal liver,<br />

originally defined as a distinct entity from a solitary hepatocellular<br />

adenoma. Due to the rarity of the disorder there is<br />

lack of consensus about treatment strategies. We explored the<br />

risk factors of complications and whether the genotype-phenotype<br />

classification used for solitary hepatic adenomas has the<br />

same prognostic significance in this disorder. Methods: We<br />

re-analyzed 35 patients with liver adenomatosis : > 10-20<br />

nodules (n=23pts) and > 20 (n=12pts); 6 patients with 5-10<br />

nodules were evaluated separately; 2 pts excluded because<br />

of glycogen storage disease. Biopsies, taken in the largest<br />

nodules, were available in 30 patients. Results: The median<br />

follow-up period was 6 years (range 1-24 years). All patients<br />

(median age 38 year, range 18-54) were female except one<br />

and 91% of them had a history of long term use of oral contraceptives<br />

(median 17 years, range 5-34). BMI was > 25 kg/m 2<br />

in 61% of patients and steatosis of the surrounding liver tissue<br />

was present in 60%. Histological subtype analysis showed: 17<br />

patients (63%) with inflammatory adenomas - in one patient<br />

associated with β-catenin activated lesions, 9 patients (33%)<br />

with HNF1-α mutation positive lesions of which 2 patients also<br />

had a β-catenin activated lesion and in 1 patient only β-catenin<br />

activated adenomata were identified. Inflammatory adenomas<br />

were clearly associated with high BMI (> 25 kg/m 2 ) and steatosis<br />

(13/16 (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 297A<br />

The following authors have nothing to disclose: Barry Schlansky, Willscott E.<br />

Naugler<br />

173<br />

Vitamin K dosing during sorafenib treatment for hepatocellular<br />

carcinoma markedly prolonged overall survival<br />

as well as progression free survival probably by<br />

augmented ischemic tumor cell damage<br />

Yoshimichi Haruna, Atsuo Inoue; Department of Gastroenterology<br />

and Hepatology, Osaka General Medical Center, Osaka, Japan<br />

Backgrounds and Aims. Sorafenib is the only oral anticancer<br />

agent for advanced hepatocellular carcinoma (HCC). However,<br />

its anticancer effects are not satisfying. It was reported<br />

that combination of vitamin K and sorafenib drastically inhibited<br />

HCC growth in vitro and in animal experiments. In this<br />

study, we examined advantage of vitamin K dosing during<br />

sorafenib treatment for HCC, focusing on different manners<br />

of changes in serum des-γ-carboxy prothrombin (DCP) levels<br />

during vitamin K combination and sorafenib alone groups.<br />

Patients and Methods. Fifty-five patients (male/female 42/13,<br />

age (mean ± SD) 73.9 ± 8.3, the Barcelona Clinic Liver Cancer<br />

(BCLC) stage B/C 21/34) were retrospectively studied.<br />

Twenty out of them were orally given vitamin K2 (45mg daily)<br />

during sorafenib treatment. There were no relevant differences<br />

between the vitamin K combination and sorafenib alone groups<br />

in characteristics at baseline. We compared the time to radiologic<br />

progression and overall survival (OS) in the two groups.<br />

The radiologic assessment was performed according to modified<br />

RECIST. Serum DCP levels was tested before and 8 weeks<br />

after the start of treatment. Results. The disease control rate was<br />

78.9% vs. 22.9% in sorafenib + vitamin K group and sorafenib<br />

alone one, respectively (P < 0.001). Median progression free<br />

survival (PFS) was 8.0 months vs. 2.5 months in sorafenib +<br />

vitamin K group and sorafenib alone one, respectively (P <<br />

0.001). Furthermore, vitamin K dosing markedly prolonged OS<br />

(median survival: 23.0 months vs. 10.0 months, P = 0.002).<br />

Interestingly, serum DCP levels were increased in partial<br />

response (PR) + stable disease (SD) cases of sorafenib alone<br />

group (2.42±0.92 Ç 2.78±0.95 Log mAU/mL, P = 0.051)<br />

despite suppressed tumor growth. In contrast, those of vitamin<br />

K-dosing group showed remarkable decline (2.35±0.56<br />

Ç 1.38±0.20 Log mAU/mL, P = 0.001). Discussion. DCP,<br />

a tumor marker, is also known as an indicator of ischemic<br />

status of HCC. The DCP elevation observed in sorafenib alone<br />

group seems to show ischemic status of tumor cells in deteriorated<br />

angiogenesis caused by sorafenib. On the other hand,<br />

there are many reports suggesting that the DCP is autologous<br />

growth factor augmenting HCC growth and a paracrine factor<br />

enhancing tumor angiogenesis. Thus, DCP elevation resulting<br />

from sorafenib treatment could incomplete the antitumor effect.<br />

In contrast, the suppressed DCP production under pharmacological<br />

vitamin K dosing could complete antitumor action of<br />

sorafenib. Conclusion. The nontoxic agent, vitamin K, dosing<br />

during sorafenib treatment for HCC markedly prolonged OS<br />

as well as PFS probably by augmented ischemic damage of<br />

tumor cells.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yoshimichi Haruna, Atsuo Inoue<br />

174<br />

Outcomes following Radiofrequency Ablation of small<br />

HCC: Impact of etiology<br />

Suraj Sharma 2 , Katherine Pratte 1 , Matthew Kowgier 2 , Morris Sherman<br />

2 ; 1 General Internal Medicine, University of Toronto, Toronto,<br />

ON, Canada; 2 Toronto Liver Clinic, Toronto Western Hospital,<br />

Toronto, ON, Canada<br />

Background: Among patients who have undergone surgical<br />

resection, etiology of HCC may be an important predictor of<br />

transplant-free survival (TFS), with some <strong>studies</strong> showing worse<br />

survival in HCV, compared to HBV. However, there is a paucity<br />

of data on the impact of etiology on outcomes in patients<br />

undergoing RFA. This study aims to characterize a cohort of<br />

patients undergoing RFA for small HCCs, and determine the<br />

impact of etiology of HCC on TFS. Methods: Chart review was<br />

conducted on all patients undergoing RFA at the University<br />

Health Network between January 1, 2008 and December 31,<br />

2011. Patients were followed for outcomes until January 1,<br />

2014. Data collected included etiology of chronic liver disease,<br />

severity of liver disease and size and location of HCC.<br />

Complete ablation was defined as the absence of residual disease<br />

at 3 months post RFA. Recurrence of HCC was defined as<br />

tumor foci developing at a location of prior RFA treatment – this<br />

may be early (1 year). Other treatment<br />

modalities offered to patients were noted. Outcome measures<br />

included transplant-free survival and tumor recurrence. Results:<br />

256 patients underwent RFA during the study period. 87% of<br />

the patients had documented evidence of cirrhosis. 81 patients<br />

with HBV (79% on treatment), 112 patients with HCV (54%<br />

treated, 20% SVR), and 63 patients with nonviral etiologies<br />

were included. 24% of the cohort was female. The mean age<br />

at diagnosis of HCC was 62.3 (±10.1) years. The mean size<br />

of the largest lesion at diagnosis was 2.8cm (±1.6cm). All<br />

patients underwent at least one session of RFA (mean 1.9 treatments).<br />

The average size of RFA-treated lesions was 2.4cm<br />

(±0.96cm). 79% of the patients had complete ablation. Recurrence<br />

occurred in 74 patients (29%), with mean recurrence-free<br />

interval of 498 days (±383d). 109 (42%) patients died or<br />

underwent liver transplant during the follow-up period. The<br />

primary outcome (transplant or death) was reached in 27%<br />

of HBV patients (median TFS 1.5 years), 53% of HCV patients<br />

(median TFS 1.7 years) (p=0.0003 HBV vs. HCV), and 40%<br />

of nonviral patients (median TFS 1.5 years). In a multivariable<br />

analysis including age, etiology and MELD score, patients<br />

with HCV had significantly worse survival than patients with<br />

HBV (HR 2.6, p


298A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

175<br />

Transcriptomic Analysis Of Human Hepatocytes In<br />

3-Dimensional Cultures With Maintained Perfusion And<br />

Transport Offers Insights Into The Pharmacotoxicology<br />

Of Clinically Relevant Concentrations Of Obeticholic<br />

Acid.<br />

Arun J. Sanyal 2 , Robert Figler 1 , Brett R. Blackman 1 , Svetlana<br />

Marukian 1 , Sol Collado 1 , Mark Lawson 1 , Aaron J. Mackey 1 ,<br />

David Manka 1 , Brian R. Wamhoff 1 , Ajit Dash 1 ; 1 HemoShear Therapeutics,<br />

Charlottesville, VA; 2 Virginia Commonwealth University,<br />

Richmond, VA<br />

NASH has emerged as the most common cause of chronic liver<br />

disease in the Western world. Recently, obeticholic acid (OCA)<br />

a semi-synthetic bile acid derivative and farnesoid-X-receptor<br />

(FXR) agonist has been shown to improve insulin sensitivity in<br />

humans with type 2 diabetes and liver histology in subjects<br />

with NASH. The precise mechanisms by which these effects<br />

are mediated are not fully known. AIMS: To use an unbiased<br />

transcriptomic analysis of hepatocytes exposed to clinically relevant<br />

doses of OCA to evaluate its effects on pathophysiologically<br />

relevant pathways. METHODS: We previously described<br />

a dynamic 3-dimensional system using liver-derived blood flow<br />

and transport parameters to restore primary human hepatocyte<br />

biology, allowing for culture at close to physiological insulin/<br />

glucose conditions and eliciting drug responses at clinically-relevant<br />

concentrations. This system represents an advance over<br />

existing in vitro systems, which require supra-physiological insulin<br />

concentrations and are associated with loss of polarity and<br />

transport functions. We applied this system to gain insights into<br />

the pharmacotoxicological effects of OCA. Primary hepatocytes<br />

from 5 human donors were exposed to OCA for 48 hours<br />

at concentrations approximating clinical therapeutic (0.5 μM)<br />

and supratherapeutic (10 μM) levels. Whole genome transcriptomics<br />

was performed by RNAseq, and the data analyzed<br />

using both unbiased algorithms and biased interrogation of<br />

specific pathways. RESULTS: A dose-dependent suppression<br />

of Cyp7a and Cyp27a along with upregulation of bile salt<br />

canalicular export transporters ABCB4 and ABCB11 and basolateral<br />

transporters OSTA and OSTB were noted as expected.<br />

Interestingly, FGF19, not known to be highly expressed in<br />

hepatocytes, was strongly upregulated in our system by OCA<br />

suggesting an autocrine suppression of FXR expression. A concomitant<br />

suppression of LXR and HNF4α signaling associated<br />

with NROB2 activation was also seen. Pleiotropic effects of<br />

OCA included suppression of TGFβ as well as TNFα signaling<br />

pathways, and could explain its beneficial effect of reducing<br />

inflammatory and fibrotic changes in NASH. Consistent with<br />

the hypercholesterolemia reported with OCA treatment was the<br />

upregulation of ApoB, the major lipoprotein of LDL. CONCLU-<br />

SIONS: OCA, in clinically relevant conditions, induces metabolic,<br />

anti-inflammatory and anti-fibrotic/oncogenic signaling<br />

in primary hepatocytes. These findings provide a mechanistic<br />

basis for OCA effects in NASH.<br />

Disclosures:<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Robert Figler - Employment: HemoShear Therapeutics, LLC<br />

Brett R. Blackman - Board Membership: HemoShear LLC; Management Position:<br />

HemoShear LLC; Patent Held/Filed: HemoShear LLC; Stock Shareholder:<br />

HemoShear LLC<br />

Svetlana Marukian - Employment: HemoShear LLC<br />

Mark Lawson - Employment: Hemoshear<br />

Aaron J. Mackey - Employment: HemoShear, LLC<br />

David Manka - Employment: HemoShear Therapeutics; Stock Shareholder:<br />

HemoShear Therapeutics<br />

Brian R. Wamhoff - Stock Shareholder: HemoShear, LLC<br />

Ajit Dash - Employment: HemoShear LLC<br />

The following authors have nothing to disclose: Sol Collado<br />

176<br />

Inhibition of NF-kB by deoxycholic acid induces miR-<br />

21/PDCD4-dependent hepatocelular apoptosis<br />

Pedro M. Rodrigues 2 , Marta B. Afonso 2 , André L. Simão 2 , Pedro<br />

M. Borralho 1 , Cecília M. Rodrigues 1 , Rui E. Castro 1 ; 1 iMed.<br />

ULisboa & Dep. of Biochemistry and Human Biology, Faculty of<br />

Pharmacy, University of Lisbon, Lisboa, Portugal; 2 iMed.ULisboa,<br />

Faculty of Pharmacy, University of Lisbon, Lisbon, Portugal<br />

MicroRNAs (miRNAs/miRs) play a key regulatory role in<br />

metabolic liver function. In particular, miR-21 deregulated<br />

expression has been associated with a wide spectrum of liver<br />

disorders, contributing to disease development and progression.<br />

We have demonstrated that deoxycholic acid (DCA), a<br />

cytotoxic bile acid implicated in the pathogenesis of non-alcoholic<br />

fatty liver disease, inhibits miR-21 expression in hepatocytes.<br />

Still, the regulatory mechanisms behind miR-21 inhibition<br />

and its contribution for cell demise are lacking. We aimed to<br />

unveil the mechanisms underlying modulation of miR-21 by<br />

DCA and to evaluate their exact contribution to DCA-induced<br />

cell death. Primary rat hepatocytes were treated with 25-200<br />

mM DCA for 24 h. Cell death, viability and caspase-3 activity<br />

were measured by the ApoTox-Glo Triplex Assay. miR-21<br />

expression was measured by qRT-PCR. Programmed cell death<br />

4 (PDCD4), a miR-21 pro-apoptotic target, was evaluated by<br />

immunoblotting and after transfecting cells with a reporter luciferase<br />

plasmid. NF-kB, IkB and active caspase-2 levels were<br />

also measured by immunoblotting, while NF-kB activation was<br />

evaluated using a specific luciferase assay and by analyzing<br />

NF-kB subcellular localization. Reactive oxygen species (ROS)<br />

levels were analysed using the fluorescent probe 2’,7’-dichlorodihydrofluorescein<br />

diacetate. Finally, for functional <strong>studies</strong>,<br />

miR-21, PDCD4, caspase-2 and NF-kB were modulated using<br />

both genetic and pharmacologic approaches, and a well-established<br />

antioxidant, N-acetyl-L-cysteine (NAC). Our results<br />

show that DCA inhibits miR-21 expression in a dose-dependent<br />

manner, while increasing PDCD4 protein levels, with a<br />

concomitant decrease in cell viability and an increase in cell<br />

death, caspase-2/-3 activation and ROS production. Either<br />

miR-21 overexpression or PDCD4 silencing hampered DCA-induced<br />

cell death. Furthermore, NF-kB activity was decreased in<br />

a similar pattern to miR-21 expression in DCA-treated hepatocytes.<br />

In fact, NF-kB inhibition, using a selective chemical inhibitor<br />

(BAY 11-7085), further decreased miR-21 and increased<br />

PDCD4 expression levels and apoptosis by DCA. In agreement,<br />

opposite results were observed in cells overexpressing<br />

NF-kB or incubated with NAC. In conclusion, NF-kB represents<br />

a major target of DCA in regulating the miR-21/PDCD4 pathway<br />

whereas, in turn, oxidative stress and caspase-2 activation<br />

are two key upstream mechanisms leading to inhibition of<br />

NF-kB transcriptional activity by DCA. As such, these signaling<br />

circuits constitute appealing targets for bile acid- and/or<br />

miR-21-associated liver pathologies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pedro M. Rodrigues, Marta B.<br />

Afonso, André L. Simão, Pedro M. Borralho, Cecília M. Rodrigues, Rui E. Castro


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 299A<br />

177<br />

Critical role of sirtuin 1-mitofusin 2 axis in ischemia/<br />

reperfusion injury in aged livers<br />

Sooyeon Lee, Kristina L. Go, Rebecca Y. U, Joseph A. Flores-Toro,<br />

Ivan Zendejas, Kevin E. Behrns, Jae-Sung Kim; Surgery, University<br />

of Florida, Gainesville, FL<br />

INTRODUCTION: Ischemia/reperfusion (I/R) injury is a fundamental<br />

obstacle in liver resection and transplantation surgery.<br />

Aged livers have markedly less reparative capacity after I/R<br />

than young livers. Impaired autophagy and consequent onset<br />

of mitochondrial dysfunction contributes to this age-dependent<br />

increase in I/R injury. Autophagy, a primary catabolic pathway<br />

in the liver, is strongly modulated by sirtuin 1 (SIRT1),<br />

a deacetylase boosting cell survival and longevity. The AIM<br />

of this study is to investigate the role of SIRT1 in I/R injury in<br />

aged livers. METHODS: Hepatocytes from male C57/BL6 mice<br />

at 3 (young) and 26 months (aged) of age were subjected to<br />

simulated I/R. Biochemical, genetic and imaging analysis were<br />

performed to assess cell death, autophagy flux, and mitochondrial<br />

function. RESULTS: While 2 h of ischemia has no effects<br />

on young hepatocytes, this short-term I/R impaired autophagy,<br />

and rapidly induced the mitochondrial permeability transition<br />

and necrotic cell death in aged hepatocytes. Immunoblotting<br />

analysis indicated a near-complete loss of SIRT1 in aged cells<br />

after I/R, which did not occur in young cells. Interestingly,<br />

adenoviral overexpression SIRT1 in aged cells failed to mitigate<br />

I/R injury, suggesting that additional factor(s) is required<br />

to support cell survival. Immunoprecipitation approaches<br />

revealed an interaction of SIRT1 with mitofusin 2 (MFN2), a<br />

mitochondrial outer membrane protein. MFN2 levels in aged<br />

cells became undetectable after a short-term I/R, while the<br />

depletion of MFN2 was not observed in young cells. Similar<br />

to SIRT1, MFN2 overexpression alone did not confer cytoprotection.<br />

However, co-overexpression of both MFN2 and SIRT1<br />

protected aged cells against I/R, implying a critical cross-talk<br />

between SIRT1 and MFN2. To further investigate the importance<br />

of SIRT1-MFN2 interaction, deletion mutants of MFN2<br />

were constructed and expressed into HEK293T cells. Consistent<br />

with the findings in hepatocytes, SIRT1 overexpression in<br />

wild type cells not only deacetylated MFN2 but also increased<br />

autophagic flux. However, autophagy stimulated by SIRT1 was<br />

subdued in the N-terminal deletion mutants (Δ2-92 and Δ262-<br />

392), suggesting a pivotal role of the N-terminal domain of<br />

MFN2 in SIRT1-mediated autophagy induction. CONCLUSION:<br />

Our findings show that the depletion of both SIRT1 and MFN2<br />

attributes to defective autophagy, mitochondrial dysfunction<br />

and cell death in aged hepatocytes after I/R. The SIRT1-MFN2<br />

axis may be an integral factor that governs survivability of<br />

aged livers after reperfusion.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sooyeon Lee, Kristina L. Go,<br />

Rebecca Y. U, Joseph A. Flores-Toro, Ivan Zendejas, Kevin E. Behrns, Jae-Sung<br />

Kim<br />

178<br />

HMGB1-driven Feedforward Hepatocyte Necroptosis<br />

Circuit in Lethal Acetaminophen-induced liver injury.<br />

Charlotte Minsart 1 , Claire Liefferinckx 1 , Sandrine Rorive 2,3 , Arnaud<br />

Lemmers 4,1 , Eric Quertinmont 1 , Eric Trépo 1,4 , Isabelle Salmon 2,3 ,<br />

Jacques Devière 4,1 , Christophe Moreno 4,1 , Isabelle A. Leclercq 5 ,<br />

Richard Moreau 6,7 , Thierry Gustot 4,1 ; 1 Laboratory of Experimental<br />

Gastroenterology, Université Libre de Bruxelles, Brussels, Belgium;<br />

2 Pathology, Erasme Hospital, Brussels, Belgium; 3 DIAPATH- center<br />

for microscopy and molecular imaging (CMMI), Gosselies,<br />

Belgium; 4 Gastroenterology and Hepato-Pancreatology, Erasme<br />

Hospital, Brussels, Belgium; 5 Laboratory of Hepato-Gastroenterology,<br />

Institut de Recherche Expérimentale et Clinique, Université<br />

Catholique de Louvain, Brussels, Belgium; 6 INSERM Unité 1149,<br />

Centre de Recherche sur l’inflammation (CRI), Paris, France; 7 UMR<br />

S_1149, Université Paris Diderot, Paris, France<br />

Background & Aims: Release of damage-associated molecular<br />

patterns, in particular High-mobility group box (HMGB)<br />

1, contributes to acetaminophen (APAP)-induced liver injury<br />

but the mechanisms involved are currently incompletely understood.<br />

The aim of the study is to investigate the contribution of<br />

HMGB1 in vivo and in vitro at early time points of the APAP-induced<br />

liver injury and its role in the propagation of necrosis<br />

process. Methods: APAP hepatotoxicity was induced in vivo<br />

by intraperitoneal injection in C57Bl/6 mice and in vitro on<br />

cultured HepaRG cells. HMGB1 was quantified by ELISA or<br />

immuno-staining. Cell death was determined by MTT, ALT, LDH<br />

and caspase-3 assays. Glycyrrhizin (GL) and ethyly pyruvate<br />

(EP) was used to inhibit HMGB1. Liposomal clodronate was<br />

administrated to mice to deplete Kupffer cells (KC). Expression<br />

of HMGB1 receptors was assessed by RT-PCR and flow<br />

cytometry. Dabrafenib and necrostatin-1was used to inhibit<br />

receptor-interacting protein (RIP)3 and RIP1 respectively.<br />

Results: In APAP-challenged mice, GL inhibited the HMGB1<br />

release (decrease of serum levels and increase in hepatocellular<br />

retention in centrolobular area) with improved survival.<br />

Depletion of KC in mice exacerbated APAP-induced hepatocyte<br />

necrosis and HMGB1 release suggesting that HMGB1 did<br />

not act through KC activation. Addition of APAP on cultured<br />

HepaRG induced cell necrosis characterized by LDH release<br />

without caspase-3 activation, and HMGB1 release. Moreover,<br />

HepaRG were exposed to APAP for 6 hours and the so-conditioned<br />

medium induced cell death of unexposed HepaRG.<br />

Inhibition of HMGB1 by GL or EP reduced APAP- and conditioned<br />

medium-induced HepaRG necrosis and further HMGB1<br />

release. Exposure of HepaRG and primary human hepatocytes<br />

to rhHMGB1 resulted in their death, underlining that HMGB1<br />

acts directly on hepatocytes. HepaRG expressed previously<br />

described HMGB1 receptors (TLR2, TLR4 and TLR9) at mRNA<br />

and protein level. Pre-treatment of HepaRG by dabrafenib, a<br />

specific RIP3 inhibitor, and not by necrostatin-1prevented this<br />

HMGB1-induced cell death. Conclusion: HMGB1 contributes<br />

to APAP-induced liver injury through a RIP3-dependent hepatocyte<br />

necroptosis using a feedforward mechanism. Inhibition<br />

of HMGB1 at the early phase of APAP-induced liver injury<br />

improved animal survival by reducing the propagation of this<br />

regulated hepatocyte necrosis.<br />

Disclosures:<br />

Christophe Moreno - Consulting: Abbvie, Janssen, Gilead, BMS; Grant/Research<br />

Support: Janssen, Gilead, Roche, Astellas<br />

Isabelle A. Leclercq - Independent Contractor: Genfit<br />

The following authors have nothing to disclose: Charlotte Minsart, Claire Liefferinckx,<br />

Sandrine Rorive, Arnaud Lemmers, Eric Quertinmont, Eric Trépo, Isabelle<br />

Salmon, Jacques Devière, Richard Moreau, Thierry Gustot


300A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

179<br />

Oxaloacetic Acid Protects the Liver From Warm Ischemia/Reperfusion<br />

Injury<br />

Grégory Merlen, Benoit Lacoste, Benoît Dupont, Valérie-Ann Raymond,<br />

Marc Bilodeau; CRCHUM, Montreal, QC, Canada<br />

Background: Liver ischemia/reperfusion (I/R) is an important<br />

cause of liver damage early after liver transplantation, post<br />

liver resection or during hemorrhagic shock. One of the mechanisms<br />

of cell death occurring in that setting is the disruption of<br />

mitochondrial activity that lead to alterations in cellular energy<br />

metabolism. We hypothesized that oxaloacetic acid (OAA),<br />

which is at the crossroads of gluconeogenesis, amino acid<br />

metabolism and the citric acid cycle, could refuel the energy<br />

metabolism during I/R and therefore protect the liver against<br />

injury. In vitro, we have already demonstrated that OAA was<br />

the most potent citric acid intermediate in its capacity to protect<br />

rat hepatocytes from hypoxia. We have also shown that the<br />

administration of OAA considerably reduces the extent of liver<br />

injury in the left portal vein ligation model of warm liver ischemia<br />

in the rat. We here analyze the potential protective effect<br />

of this compound in a model of hepatic I/R. Methods: Animals<br />

were subjected to 1 hour of left portal pedicle ligation (common<br />

bile duct, left hepatic artery and left portal vein) followed<br />

by reperfusion by unclamping. Animals treated with OAA<br />

[100mg/kg] received a bolus injection through the ileocolic<br />

vein 30 minutes before surgery. The extent of liver I/R injury<br />

was assessed by serum transaminase levels measurements, histological<br />

signs of tissue damage and liver weight (edema).<br />

Results: Treatment with OAA before reperfusion significantly<br />

decreased the AST levels released in blood. After 2 hours of<br />

reperfusion the AST and ALT levels were respectively decreased<br />

by 43%±14% (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 301A<br />

Here, we investigated the underlying mechanisms of immune<br />

tolerance by the gut-liver axis, focusing on interactions between<br />

gut microbiota and hepatic immune competent cells. METHODS:<br />

Male C57BL/6 mice were administered an initial and subsequent<br />

sub-lethal dose of ConA to induce immunological tolerance.<br />

Liver mononuclear cells were separated 12 h after the<br />

final ConA injection and analyzed by flow cytometry. Immune<br />

cell subset cytokine production stimulated with TLR ligands was<br />

measured and the composition of intestinal bacterial flora was<br />

evaluated by T-RFLP and comprehensive metagenomics. Intestinal<br />

permeability was measured by FITC-dextran following<br />

ConA injection. Mice were treated with antibiotics (ampicillin,<br />

neomycin, metronidazole, and vancomycin) or fecal microbiota<br />

transplantation to clarify the role of the gut-liver axis in liver<br />

tolerance. RESULTS: A single ConA injection induced severe<br />

liver inflammation but ConA administration 7 days after initial<br />

injection increased immunosuppresive CD11c + DCs (D7-cDC)<br />

and regulatory T cells in the liver and promoted immunological<br />

tolerance. D7-cDC had regulatory characteristics and produced<br />

IL-10 and TGF-β upon TLR9 ligand stimulation especially<br />

from damaged hepatocytes. As an initiation of liver tolerance,<br />

intestinal permeability was significantly increased at 4 h following<br />

a single ConA injection. Importantly, the composition of<br />

intestinal bacterial flora changed and the ratio of Clostridium<br />

subcluster XIV to Bacteroides increased thereafter. Transplantation<br />

of fecal microbiota derived from mice post-ConA administration,<br />

but not from untreated mice, to gut sterilized mice<br />

induced immunosuppressive CD11c + cDCs and regulatory T<br />

cells in the liver and reduced liver injury by ConA. A similar<br />

result was observed in germ-free and gnotobiotic mice inoculated<br />

with fecal microbiota derived from ConA-injected SPF<br />

mice. CONCLUSIONS: Dysbiosis with increased intestinal permeability<br />

following ConA administration promotes the migration<br />

of immunosuppressive cells in the liver in preparation for<br />

further liver injury. This study identifies a novel homeostasis<br />

pathway that regulates immune activation and tolerance in the<br />

liver through the gut-liver axis.<br />

Disclosures:<br />

Takanori Kanai - Grant/Research Support: Mitsubishi Tanabe Pharma Corporation<br />

The following authors have nothing to disclose: Nobuhiro Nakamoto, Hirotoshi<br />

Ebinuma, Po-sung Chu, Nobuhito Taniki, Takeru Amiya, Akihiro Yamaguchi,<br />

Shunsuke Shiba, Hidetsugu Saito<br />

182<br />

Chronic ethanol feeding suppresses Con A induced T cell<br />

response and hepatitis in ALDH2-deficient mice<br />

Yanhang Gao 1,2 , Yong He 1 , Dechun Feng 1 , Zhou Zhou 1 , Bin<br />

Gao 1 ; 1 NIAAA, National Institutes of Health, Rockville, MD;<br />

2 Hepatology, The first hospital of Jilin University, Changchun,<br />

China<br />

Background and Aims: Aldehyde dehydrogenase 2 (ALDH2)<br />

is well known about its role on detoxifying aldehydes in ethanol<br />

metabolism. An ALDH2 inactivating mutation is the most<br />

common single point mutation in humans, mostly found in East<br />

Asians, which can cause acetaldehyde accumulation after<br />

alcohol consumption. However, how acetaldehyde accumulation<br />

affects T cell response and hepatitis in ALDH2-deficent<br />

individuals remains unknown. Methods: Wide-type and ALDH2<br />

knockout mice were subjected to ethanol feeding for 6 weeks,<br />

followed by injection of a single dose of Concanavalin A (Con<br />

A). Liver injury and serum cytokine levels were evaluated.<br />

Results: Con A injection rapidly induced T cell hepatitis, which<br />

recapitulates the histological and pathological sequelae of T<br />

cell-mediated hepatitis in viral hepatitis patients. Compared<br />

with ethanol-fed wild-type mice, ethanol-fed ALDH2 -/- mice had<br />

lower degree of liver damage post Con A injection, as demonstrated<br />

by the lower level of serum alanine aminotransferase<br />

(ALT), the less infiltration of neutrophils, the fewer number of<br />

activated macrophages, and the smaller area of necrosis in<br />

the liver. Furthermore, serum levels of several pro-inflammatory<br />

cytokines including IFN-γ, TNF-α, MCP-1, IL-4, IL-6, IL-10,<br />

IL12p70, were lower in ethanol-fed ALDH2 knockout mice than<br />

in wide-type mice post Con A injection. In agree with serum<br />

cytokine levels, hepatic activation of the IFN-, IL-6, IL-4 downstream<br />

signaling molecule signal transducer and activator of<br />

transcription (STAT1, 3, 6) were lower in ALDH2 knockout mice<br />

than in wide-type mice. Furthermore, ethanol-fed ALDH2 knockout<br />

mice had much higher levels of plasma corticosterone than<br />

ethanol-fed wild-type mice. Inhibition of endogenous glucocorticoid<br />

activity by pretreatment with the glucocorticoid receptor<br />

antagonist RU-486 restored Con A-mediated liver injury in<br />

ALDH2-deficeint mice. Conclusions: Chronic ethanol feeding<br />

results in greater elevation of plasma corticosterone levels in<br />

ALDH2 knockout mice than in wild-type mice. These elevated<br />

corticosterone levels inhibit Con A-induced T cell response and<br />

hepatitis in ALDH2 knockout mice. ALDH2-deficient individuals<br />

who are excessive drinkers may have reduced T cell response<br />

and are more susceptible to hepatitis viral infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yanhang Gao, Yong He, Dechun<br />

Feng, Zhou Zhou, Bin Gao<br />

183<br />

The antifibrotic effect of IL-4Rα signaling depends on<br />

macrophage subsets prevalent during liver fibrosis progression<br />

and reversal<br />

Shih-Yen Weng 1 , Santosh Vijayan 1 , Xiaoyu Wang 1 , Yilang Tang 4 ,<br />

Kornelius Padberg 1 , Yong Ook Kim 1 , Brombacher Frank 5 , Jeff R.<br />

Crosby 3 , Michael L. McCaleb 3 , Ari Waisman 4 , Ernesto Bockamp 1 ,<br />

Detlef Schuppan 1,2 ; 1 Institute of Translational Immunology, University<br />

Medicine, Johannes Gutenberg University, Mainz, Germany,<br />

Mainz, Germany; 2 Beth Israel Deaconess Medical Center, Boston,<br />

MA; 3 Isis Pharmaceuticals, Carlsbad, CA; 4 Institute for Molecular<br />

Medicine, Mainz, Germany; 5 Institute of Infectious Disease and<br />

Molecular Medicine, Cape town, South Africa<br />

Background and aims: In response to various stimuli, macrophages<br />

can be functionally divided into M1 (inflammatory) and<br />

M2 (anti-inflammatory) subsets. Alteration of the M1-M2 ratio<br />

likely impacts liver fibrosis progression and reversal. IL-4Rα,<br />

which is activated by IL-4 and IL-13 has been linked to M2<br />

polarization. However, the functional role of IL-4Rα in liver<br />

fibrosis progression and reversal remained unclear. Methods:<br />

IL-4Rα -/- and wild type mice were treated with CCL 4<br />

via oral<br />

gavage for 6 weeks. Liver fibrosis reversal was investigated<br />

after 2 weeks of CCL 4<br />

withdrawal. Antisense oligonucleotides<br />

(ASO) were applied intraperitoneally (40mg/kg, 3 times per<br />

week) during the 6 week s of CCL 4<br />

treatment or the during 2<br />

weeks off CCL4. Results: At 6 weeks of CCL 4<br />

treatment, IL-4Rα -/-<br />

mice had 25% less fibrosis than their wild type littermates. Flow<br />

cytometry analysis of the liver of IL-4Rα -/- mice showed less<br />

inflammation indicated by a reduction of B cells, and CD4<br />

and CD8 T cells. The proportion of resident M1 macrophages<br />

was significantly increased and that of proinflammatory monocytic<br />

Ly6c hi macrophages largely reduced in IL-4Rα -/- livers.<br />

To further validate IL-4Rα function, mice with cell specific deletion<br />

of IL-4Rα were generated and subjected to CCL 4<br />

treatment.<br />

No overt change of liver fibrosis induction was found<br />

in T cell-specific knockout mice (IL-4Rα ∆CD4 ). However, after 6<br />

weeks of CCL4-treatment, fibrosis was significantly attenuated<br />

in mice with myeloid cell-specific IL-4Rα deletion (IL-4Rα ∆LysM ).


302A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

In contrast, IL-4Rα ∆LysM mice displayed retarded fibrosis resolution<br />

as evidenced by slower clearance of hydroxyproline<br />

and Sirius-Red stained collagen at 2 weeks off CCL 4<br />

. This was<br />

accompanied by a reduction of Ly-6c lo restorative and M2<br />

macrophages and an increase of M1 macrophages. In a therapeutic<br />

approach, we applied an ASO targeting IL-4Rα (IL-4Rα<br />

ASO) to CCL 4<br />

-treated mice. IL-4Rα ASO strongly attenuated<br />

fibrosis progression but mitigated fibrosis resolution after CCL 4<br />

withdrawal. Conclusion: Our <strong>studies</strong> demonstrate that IL-4Rα<br />

modulates fibrosis progression and reversal in discordant ways<br />

through macrophages. During progression, IL-4Rα signaling<br />

increases inflammation and fibrosis by activating Ly6c hi macrophages;<br />

during reversal, IL-4Rα potentiates restorative (M2,<br />

Ly6c lo ) macrophage signaling. Accordingly, IL-4Rα ASO treatment<br />

attenuates fibrosis progression, but retards reversal.<br />

Disclosures:<br />

Jeff R. Crosby - Employment: ISIS Pharmaceuticals<br />

Michael L. McCaleb - Employment: Isis Pharmaceuticals; Stock Shareholder: Isis<br />

Pharmaceuticals<br />

The following authors have nothing to disclose: Shih-Yen Weng, Santosh Vijayan,<br />

Xiaoyu Wang, Yilang Tang, Kornelius Padberg, Yong Ook Kim, Brombacher<br />

Frank, Ari Waisman, Ernesto Bockamp, Detlef Schuppan<br />

184<br />

Liver sinusoidal endothelial cells induce neutrophil<br />

extracellular traps in liver sterile inflammation via<br />

IL-33/ST2 axis<br />

Hai Huang, Hamza Yazdani, Patricia Loughran, Heth R. Turnquist,<br />

Allan Tsung; Department of Suegery, University of Pittsburgh Medical<br />

Center, Pittsburgh, PA<br />

Both liver sinusoidal endothelial cells (LSECs) and neutrophils<br />

are involved and interact in liver ischemia/reperfusion (I/R)<br />

injury. Damaged LSECs, as the major sources of IL-33, play<br />

important role in neutrophil infiltration during liver I/R. We<br />

recently found that in response to damage associated molecular<br />

patterns (DAMPs), infiltrated neutrophils forming neutrophil<br />

extracellular traps (NETs), exacerbate sterile inflammatory<br />

injury during liver I/R. We here sought to determine the role<br />

of IL-33 released from LSECs in NET formation during liver I/R.<br />

WT or IL-33 KO mice were subjected to a non-lethal warm liver<br />

I/R model. Recombinant IL-33 (rIL-33) or soluble ST2 (sST2,<br />

decoy receptor of IL-33) was administered in select groups.<br />

IL-33 KO mice or sST2-treated WT mice were significantly protected<br />

from I/R injury, having significantly less serum sALT and<br />

hepatic necrosis, lower levels of systemic cytokines, abrogation<br />

of proinflammatory signaling pathways compared to WT mice.<br />

rIL-33 administered during I/R exacerbated hepatotoxicity and<br />

systemic inflammation. Although significant neutrophils infiltration<br />

was found in both IL-33 KO and WT mice, NETs formation<br />

decreased in IL-33 KO mice during liver I/R compared<br />

with WT mice, associated with significant less serum level of<br />

myeloperoxidase (MPO)-DNA complexes and tissue level of<br />

citrullinated-histone H3 (NET markers). In addition, treatment<br />

of sST2 reduced NET formation whereas significant increased<br />

NETs were found in rIL-33 treated mice after liver I/R. In vitro,<br />

IL-33 released from LSECs that were subjected to hypoxia (1%<br />

O 2<br />

) promotes NET formation. Directly using rIL-33 stimulates<br />

neutrophils confirmed IL-33/ST2-MyD88 signaling pathway in<br />

NET formation. Using either IL-33 KO LSECs, or co-stimulated<br />

neutrophil with IL-33 neutralizing antibody or sST2 blocked<br />

NET formation. Our study demonstrates IL-33 from LSECs initiates<br />

a feed-forward mechanism to neutrophils inducing NETs<br />

formation in excessive inflammation and hepatotoxicity during<br />

liver I/R.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hai Huang, Hamza Yazdani,<br />

Patricia Loughran, Heth R. Turnquist, Allan Tsung<br />

185<br />

15-PGDH prevents LPS/GalN-induced acute liver injury<br />

through inhibiting Kupffer cell activation<br />

Lu Yao, Chang Han, Tong Wu; pathology and laboratory medicine,<br />

tulane university, New Orleans, USA Minor Outlying Islands<br />

The NAD+-dependent 15-hydroxyprostaglandin dehydrogenase<br />

(15-PGDH) catalyzes the oxidation of the 15(S)-hydroxyl<br />

group of Prostaglandin E 2<br />

(PGE 2<br />

), converting the pro-inflammatory<br />

PGE 2<br />

to the anti-inflammatory 15-keto-PGE 2<br />

(an<br />

endogenous ligand for peroxisome proliferator-activated receptor-gamma<br />

[PPAR-γ]). PPAR-γ is a ligand-dependent transcriptional<br />

factor, which plays an important role in regulation of<br />

inflammatory cell activation. To evaluate the significance of<br />

15-PGDH/PPAR-γ cascade in liver inflammation, we generated<br />

transgenic mice with targeted expression of 15-PGDH in the<br />

liver (15-PGDH Tg) and the animals were subjected to lipopolysaccharide<br />

(LPS)/Galactosamine (GalN) induced acute liver<br />

inflammation and tissue damage. Compared to the wild type<br />

mice, the 15-PGDH Tg mice showed lower levels of alanine<br />

aminotransferase (ALT) and aspartate aminotransferase (AST),<br />

less liver tissue damage, less hepatic apoptosis/necrosis, less<br />

macrophage activation, and lower inflammatory cytokine production.<br />

In cultured Kupffer cells, treatment with 15-keto-PGE 2<br />

or the conditioned medium (CM) from 15-PGDH Tg hepatocyes<br />

inhibited LPS-induced cytokine production. Both 15-keto-PGE 2<br />

and the CM from15-PGDH Tg hepatocyes also significantly<br />

up-regulated the expression of PPAR-γ down-stream genes in<br />

Kupffer cells. On the other hand, 15-PGDH overexpression<br />

or 15-keto-PGE 2<br />

treatment of hepatocytes did not influence<br />

TNF-α-induced hepatocyte apoptosis. These results suggest<br />

that the resistance of 15-PGDH Tg mice to LPS/GalN-induced<br />

liver injury is mediated through 15-keto-PGE 2<br />

, which activates<br />

PPAR-γ in Kupffer cells and inhibited inflammatory cytokine<br />

production. Consistent with this assertion, we observed that the<br />

PPAR-γ antagonist, GW9662, reversed the effect of 15-keto-<br />

PGE 2<br />

in Kupffer cell (in vitro) and restored the susceptibility<br />

of 15-PGDH Tg mice to LPS/GalN-induced acute liver injury<br />

(in vivo). Taken together, our findings provide novel evidence<br />

that hepatic overexpression of 15-PGDH protects against<br />

LPS/GalN-induced acute liver injury by inhibiting Kupffer cell<br />

inflammatory response. Thus, 15-keto-PGE 2<br />

is an endogenous<br />

PPAR-γ ligand that may be utilized as a pharmacological agent<br />

to ameliorate liver inflammation and tissue damage.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 303A<br />

Disclosures:<br />

The following authors have nothing to disclose: Lu Yao, Chang Han, Tong Wu<br />

186<br />

IL-17 signaling in hepatocellular carcinoma promoted<br />

by ethanol.<br />

Hsiao-Yen Ma 2,1 , Jun Xu 1,2 , Mengxi Sun 1,2 , David A. Brenner 2 ,<br />

Tatiana Kisseleva 1 ; 1 Department of Surgery, UC San Diego, La<br />

Jolla, CA; 2 Department of Medicine, UC San Diego, La Jolla, CA<br />

Hepatocellular carcinoma (HCC) is a malignant tumor made of<br />

cells dedifferentiated from mature hepatocytes, which usually<br />

arises in patients with end stage cirrhosis. Progression of HCC<br />

is associated with upregulation of inflammation and constitutive<br />

activation of STAT3. Meanwhile, prolonged alcohol consumption<br />

has strong immunosuppressive and hepatotoxic effects,<br />

implying chronic alcohol consumption would promote HCC<br />

development. Furthermore, ALD patients showed significant<br />

increase in plasma and hepatic IL-17A expression. Blockage<br />

of IL-17 signaling significantly reduced alcohol induced steatohepatitis<br />

and fibrosis in preclinical mouse models. However,<br />

the role of IL-17 signaling in alcohol promoted HCC remains<br />

unclear. AIM: To determine the role of IL-17A signaling pathway<br />

during alcohol promoted HCC development and evaluate<br />

IL-17A as a therapeutic target for HCC. METHODS: Two different<br />

models were used to assess the IL-17 signaling in alcohol<br />

promoted HCC. First, hepatic carcinogen (DEN) induced<br />

HCC in which WT and IL-17RA null mice (KO) were injected<br />

with single dose of DEN at 14 days old. Second, steatohepatitis<br />

associated HCC in which MUP-uPA transgenic mice were<br />

crossed with WT and IL-17RA KO to generate MUP-uPA|WT<br />

(MWT) and MUP-uPA|IL-17RA KO (MKO) mice. By 12 weeks<br />

old, each genotype was assigned to pair-fed group and EtOH<br />

group to receive Liber-Decarli HFD or alcohol diet for 18 weeks<br />

respctively. The HCC development in both WT and KO mice<br />

were compared and Stat3 signal pathway was evaluated. To<br />

investigate how IL-17A signal regulating HCC associated fibroblasts<br />

and macrophage, MC-38-GFP cancer cells were injected<br />

into the spleen of WT and KO recipients for 10 days and the<br />

tumor size and numbers were compared. RESULTS: Chronic<br />

administration of the Liber-DeCarli alcohol diet accelerated<br />

tumorogenesis in both WT/MWT and KO/MKO mice. Blockade<br />

of IL-17 signaling significantly suppressed tumor size and<br />

tumor number in both pair-fed and EtOH group. In MUP-uPA<br />

mice, liver fibrosis was increased after EtOH feeding, which<br />

was decreased by blockade of IL-17 signaling. Blockade of<br />

IL-17 signaling showed less inflammation (reduced TNFa, IL1b,<br />

TGFb1), mediators of metastasis (MMP7, MMP9) and ROS<br />

production (Nox1, Nox2). As expected, CYP2E1 was elevated<br />

in EtOH group, but suppressed in KO mice. CONCLUSION:<br />

Chronic alcohol consumption accelerated tumor development<br />

which was mediated by IL-17 signaling pathway. The role of<br />

IL-17 signaling during alcohol promoted HCC may be via regulating<br />

STAT3 signaling pathway and upregulation of CYP2E1.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hsiao-Yen Ma, Jun Xu, Mengxi<br />

Sun, David A. Brenner, Tatiana Kisseleva<br />

187<br />

Deletion of fibrocytes in mice attenuates experimental<br />

liver fibrosis.<br />

Jun Xu 1,2 , Min Cong 1,2 , Tae Jun Park 1,2 , David A. Brenner 1 , Tatiana<br />

Kisseleva 2 ; 1 Department of Medicine, UCSD, San Diego, CA;<br />

2 Department of Surgery, UC San Diego, La Jolla, CA<br />

BACKGROUND:Bone marrow (BM) fibrocytes, designated as<br />

CD45 + and Collagen type I + (Col1a1) cells, are recruited to<br />

the injured liver, however, their role in liver fibrosis remains<br />

unclear. AIM: To determine the role of fibrocytes in pathogenesis<br />

of liver fibrosis. METHODS: 1The contribution of fibrocytes to<br />

liver fibrosis was studied in BM chimeric mice devoid of fibrocytes<br />

(DFibrocyte mice), in which fibrocyte death was induced<br />

by expression of Diphtheria toxin a (DTA) in CD45 + Col1a1 +<br />

fibrocytes. Specifically, tamoxifen-inducible Col1a1 ER-Cre<br />

mice were crossed with Rosa26 flox-Stop-flox-YFP reporter mice ±<br />

Rosa26 flox-Stop-flox-DTA mice, and used as donors for BM transplantation<br />

into lethally irradiated wt mice to generate wt and<br />

DFibrocyte mice. 2The role of Col1a1 in regulation of fibrocyte<br />

function was studied in BM chimeric Col1a1 5’SL-/- -into-wt<br />

mice, in which mutation of 5’stemloop (5’SL -/- mice) prevented<br />

proper Col1a1 translation in fibrocytes. 3)All mice were subjected<br />

to CCl 4<br />

for 6 w. The therapeutic potential of Serum<br />

Amyloid P (SAP), a natural inhibitor of fibrocytes, was tested.<br />

4To translate our findings to humans, the role of fibrocytes<br />

as prognostic biomarker was evaluated in patients with liver<br />

cirrhosis. RESULTS: Cell fate mapping revealed that 20% of<br />

hepatic fibrocytes become a-SMA myofibroblasts, while 80%<br />

become myeloid cells, which serve as a significant source of<br />

TGFβ1, IL-6, and IL-1b1 in CCl 4<br />

treated wt mice. Deletion of<br />

fibrocytes/progeny in DFibrocyte mice resulted in inhibition<br />

of CCl 4<br />

induced liver fibrosis by 50%, as shown by reduced<br />

Col1a1, a-SMA, and TGFb1 mRNA expression. Interestingly,<br />

fibrocyte ablation also resulted in suppression of hepatic pro-inflammatory<br />

macrophages (M1) and promote proliferation of<br />

anti-inflammatory macrophages (M2), suggesting that fibrocytes<br />

regulate M1/M2 macrophage polarization. Next, we<br />

tested if Col1a1 expression affects fibrocyte function. Indeed,<br />

liver fibrosis was reduced in Col1a1 5’SL-/- into-wt mice by >30%,<br />

and was associated with impaired proliferation of BM fibrocytes.<br />

Surprisingly, activation of non-fibrocyte-derived myeloid<br />

lineages was also reduced, indicating that fibrocytes have an<br />

important immunoregulatory function. In support, the number<br />

of circulating fibrocytes in patients correlated with the stage<br />

of liver fibrosis, and with low levels of serum SAP. Administration<br />

of SAP significantly ameliorated liver fibrosis in CCl 4<br />

-wt<br />

mice. CONCLUSION: Fibrocytes contribute to liver fibrosis indirectly<br />

by increasing liver inflammation and secreting fibrogenic<br />

agonists. Targeting fibrocytes with SAP may become a novel<br />

therapy of liver fibrosis in patients with reduced serum levels<br />

of SAP.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jun Xu, Min Cong, Tae Jun Park,<br />

David A. Brenner, Tatiana Kisseleva


304A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

188<br />

Splenectomy attenuates advanced liver fibrosis stimulating<br />

hepatic accumulation of anti-fibrogenic monocytes/<br />

macrophages<br />

Yuji Iimuro 1,2 , Akito Yada 1 , Toshihiro Okada 1 , Naoki Uyama 1 ,<br />

Jiro Fujimoto 1 ; 1 Department of Surgery, Hyogo College of Medicine,<br />

Hyogo, Japan; 2 Surgery, Nirasaki Municipal Hospital, Nirasaki,<br />

Japan<br />

Splenectomy (SP), which is performed in cirrhotic patients with<br />

hypersplenism, has been reported to improve liver function;<br />

however the underlying mechanism remains obscure. The AIM<br />

of the present study was to investigate the mechanism using<br />

a murine model and human liver tissues. Methods: C57BL/6<br />

male mice were allowed to drink water including thioacetamide<br />

(TAA: 300 mg/l) ad libitum for 32 weeks. After SP at<br />

32 weeks, mice were sacrificed on days 1, 7, and 28, respectively,<br />

while TAA-administration was continued. Perioperative<br />

changes in peripheral blood and liver tissues were analyzed.<br />

In cirrhotic patients with hepatocellular carcinoma (HCC), who<br />

underwent laparoscopic SP in advance, non-tumorous liver<br />

tissues obtained from resected specimen were histologically<br />

analyzed. Results: TAA treatment of mice for 32 weeks reproducibly<br />

achieved advanced liver fibrosis with splenomegaly,<br />

thrombocytopenia, and leukocytopenia, well representing<br />

the compensated liver cirrhosis in humans. After SP, thrombocytopenia<br />

and leukocytopenia were quickly improved, and<br />

liver fibrosis was obviously attenuated. Among the leukocytes,<br />

monocytes/macrophages were selectively increased in peripheral<br />

blood, as well as in the liver. Meanwhile, progenitor-like<br />

cells expressing CK-19, EpCAM, or CD-133 appeared along<br />

the fibrous scar in the liver after TAA treatment, and gradually<br />

disappeared after SP. Monocytes/macrophages accumulated<br />

in the liver, most of which were negative for Ly-6C, existed<br />

adjacent to the hepatic progenitor-like cells. qRT-PCR indicated<br />

increased canonical Wnt and decreased Notch signals in<br />

the liver, and Wnt2 protein expression in monocytes/macrophages<br />

was confirmed. A significant amount of β-catenin accumulated<br />

in the adjacent progenitor-like cells, and Ki67-positive<br />

relatively small hepatic cells were significantly increased. Protein<br />

expression of MMP-9, to which Ly-6G-positive neutrophils<br />

contributed, was also increased in the liver after SP. In cirrhotic<br />

patients with HCC, thrombocytopenia and lekocytopenia were<br />

quickly attenuated after SP, and the percentage of monocytes<br />

in peripheral blood was selectively increased. In the their liver<br />

tissues, remarkable accumulation of round- or oval-shaped macrophages/monocytes<br />

positive for CD163 was detected after<br />

SP, and they existed closely to the CK19-positive cells, which<br />

spread out from the ductular reactions, suggesting interaction<br />

between these cells. Conclusions: The hepatic accumulation of<br />

monocytes/macrophages, the reduction of fibrosis, and the<br />

gradual disappearance or expansion of hepatic progenitor-like<br />

cells, possibly play significant roles in the tissue remodeling<br />

process in cirrhotic livers after SP.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yuji Iimuro, Akito Yada, Toshihiro<br />

Okada, Naoki Uyama, Jiro Fujimoto<br />

189<br />

The role of Mesothelin in the pathogenesis of cholestatic<br />

liver fibrosis<br />

Yukinori Koyama 2,1 , Ping Wang 2,1 , David A. Brenner 1 , Tatiana<br />

Kisseleva 2 ; 1 Medicine, University of Californiam, San Diego, La<br />

Jolla, CA; 2 Surgery, University of California, San Diego, CA<br />

Background: There are no curative treatments for patients<br />

with the cholestatic liver diseases, primary sclerosing cholangitis<br />

(PSC), and primary biliary cirrhosis (PBC). Although the<br />

etiology of PSC and PBC remains unclear, activated portal<br />

fibroblasts (aPFS) have been implicated in pathogenesis of<br />

cholestatic clinical and experimental liver fibrosis. We have<br />

identified that mesothelin (Msln) is a specific marker for portal<br />

fibroblasts. Aim: To examine the role of Msln in cholestatic liver<br />

fibrosis in mice. Methods: Using Msln as a specific marker of<br />

aPFs, we investigated the development of BDL-induced liver<br />

fibrosis in mice devoid of aPFs (Msln DTA mice) that were generated<br />

by expression of Diphtheria toxin-a (DTA) in Msln + cells<br />

(by crossing the Msln ER-Cre and Rosa26 flox-Stop-flox-DTA mice).<br />

Next, the role of Msln in cholestatic liver fibrosis was examined<br />

in Msln knockout mice (Msln -/- mice). To explore the role Msln<br />

in TGF-β signaling, we generated transgenic Msln DSmad4 mice<br />

in which Smad4 was specifically deleted in Msln + aPFs. Wt<br />

littermates served as controls. All mice were subjected to bile<br />

duct ligation (BDL), and livers were analysed for development<br />

of liver fibrosis. aPFs from wt and Msln -/- mice were isolated<br />

and analysed by RNA-Seq. Finally, to examine if Msln can be<br />

a new therapeutic target for cholestatic liver fibrosis, BDL wt<br />

mice were treated with anti-Msln Ab (5 mg, 10 mg / mouse).<br />

Results: aPF-ablated Msln DTA mice developed 50% less fibrosis<br />

in response to BDL than wt mice, demonstrating the requirement<br />

of aPFs for cholestatic liver fibrosis. Msln-/- mice had<br />

less fibrosis in response to BDL (5 and 17 days), but this effect<br />

was not observed in carbontetrachloride (CCl 4<br />

)-treated mice,<br />

demonstrating a role for Msln in cholestatic but not hepatotoxic<br />

induced liver fibrosis. Msln -/- aPFs exhibit a defect in migration<br />

and proliferation by modulating TGF-β1 signaling, as demonstrated<br />

by reduced number of cells populating the scratch area,<br />

and decreased expression of Ki-67 by Msln -/- aPFs. Furthermore,<br />

we found that expression of TGF-β1 target genes, such<br />

as TGF-β2, PAI-1 and Activin1, were strongly downregulated<br />

in Msln -/- aPFs compared with wt aPFs, and were associated<br />

with reduced phosphorylation of Smad2 when stimulated with<br />

TGF-β1. The Msln DSmad4 mice developed less fibrosis. Administration<br />

of anti-Msln antibody to BDL-injured mice injury attenuated<br />

liver fibrosis in a dose dependent manner (compared to<br />

IgG control). Conclusion: aPFs and in particular the expression<br />

of Msln in aPFs play important roles in the pathogenesis of<br />

cholestatic liver fibrosis. Msln might become a novel target for<br />

treatment of cholestatic liver fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yukinori Koyama, Ping Wang,<br />

David A. Brenner, Tatiana Kisseleva


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 305A<br />

190<br />

Knockout of the substance P receptor, neurokinin-1,<br />

reduces liver fibrosis in cholestatic bile duct ligated (BDL)<br />

mice<br />

Ying Wan 2 , Nan Wu 3 , Julie Venter 3 , Yuyan Han 3 , Holly A.<br />

Standeford 4 , Haibo Bai 2 , Shanika Avila 3 , Heather L. Francis 1 ,<br />

Lindsey Kennedy 4 , Micheleine Guerrier 3 , Tina Kyritsi 3 , Tami Annable<br />

4 , Shannon S. Glaser 1 , Gianfranco Alpini 1 , Fanyin Meng 1 ;<br />

1 Research, S&W DDRC and Medicine, Central Texas Veterans<br />

Health Care System and Scott & White, Temple, TX; 2 Operational<br />

Funds, BaylorScott&White, Temple, TX; 3 Medicine, Texas A&M<br />

University HSC, Temple, TX; 4 Research, Central Texas Veterans<br />

Health Care System, Temple, TX<br />

During the progression of cholestatic liver diseases, proliferating<br />

cholangiocytes acquire neuroendocrine features and<br />

secrete sensory neurotransmitters such as α-calcitonin gene<br />

related peptide (α-CGRP) and substance P (SP). We have previously<br />

shown that: (i) SP, a proteolytic product of tachykinin<br />

(Tac1) that is deactivated by membrane metalloendopeptidase<br />

(MME), stimulates cholangiocarcinoma growth by interacting<br />

with its receptor, NK1R; and (ii) knockdown of NK1R decreases<br />

biliary hyperplasia in cholestatic bile duct ligated (BDL) mice.<br />

Thus, we aim to determine the autocrine/paracrine role of the<br />

SP/NK1R axis on liver fibrosis in normal and cholestatic mice.<br />

Methods: In normal and BDL WT and NK1RKO mice, we measured<br />

serum SP levels and the mRNA expression of Tac1 and<br />

MME in total liver samples and purified small and large cholangiocytes.<br />

We measured: (i) liver fibrosis by Sirius Red staining<br />

in liver sections; and (ii) the mRNA expression of the fibrosis<br />

genes, α-SMA and fibronectin, in total liver tissues from: (i)<br />

normal wild-type (WT) mice treated for 1 wk with saline or<br />

SP (2.5 mmol/Kg BW by IP implanted minipumps); and (ii)<br />

normal and BDL WT and NK1R KO mice. In vitro, murine<br />

cholangiocyte lines were treated with BSA (basal) or SP (100<br />

nM) in the absence/presence of L-733, 060 (25 μM, NK1R<br />

antagonist) for 6-72 hr before evaluating the expression of the<br />

aforementioned fibrotic genes. Results: There was an increase<br />

in SP serum levels in normal WT mice treated with SP and in<br />

BDL WT compared to normal WT mice. The expression of Tac1<br />

increased in BDL WT mice, whereas MME mRNA expression<br />

decreased in BDL WT compared to normal mice; an opposite<br />

pattern was observed in BDL NK1RKO mice compared to<br />

BDL WT mice. There was enhanced fibrosis and fibrosis gene<br />

expression in normal WT mice treated with SP and in BDL WT<br />

mice compared to normal WT mice. A significant decrease<br />

in liver fibrosis and the expression of α-SMA and fibronectin<br />

was observed in BDL NK1RKO mice compared to BDL WT<br />

mice. Small cholangiocytes isolated from WT BDL mice and<br />

NK1RKO mouse livers displayed a less fibrotic phenotype<br />

when compared to large cholangiocytes. In addition, large<br />

cholangiocytes derived from NK1RKO mice showed a significant<br />

reduction in fibrotic markers relative to BDL WT mice.<br />

Treatment of large murine cholangiocytes with SP increased the<br />

expression of fibrosis genes, whereas L-733, 060 reduced the<br />

SP-induced expression of α-SMA and fibronectin. Conclusion:<br />

We propose that modulation of the synthesis of neurotransmitters<br />

modulated by sensory innervation may be important in the<br />

modulation of liver fibrosis during the progression of cholestatic<br />

disorders.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ying Wan, Nan Wu, Julie Venter,<br />

Yuyan Han, Holly A. Standeford, Haibo Bai, Shanika Avila, Heather L. Francis,<br />

Lindsey Kennedy, Micheleine Guerrier, Tina Kyritsi, Tami Annable, Shannon S.<br />

Glaser, Gianfranco Alpini, Fanyin Meng<br />

191<br />

Loss of Cytoglobin Exacerbates Liver Fibrosis and Cancer<br />

Development in Steatohepatitis by Activating the<br />

Oxidative Stress Pathway<br />

Thuy T. Le 2 , Yoshinari Matsumoto 2 , Tuong Thi Van Thuy 2 , Hoang<br />

Hai 2 , Katsutoshi Yoshizato 1 , Norifumi Kawada 2 ; 1 Academic Advisor’s<br />

Office, PhoenixBio Co.Ltd., Hiroshima, Japan; 2 Department<br />

of Hepatology, Graduate School of Medicine, Osaka City University,<br />

Osaka, Japan<br />

Objective: To clarify the role of cytoglobin (CYGB) in the<br />

development of liver fibrosis and cancer in humans and in a<br />

mouse model of non-alcoholic steatohepatitis (NASH). Methods:<br />

CYGB expression was assessed in patients with NASH or<br />

hepatocellular carcinoma (HCC). Cygb-deficient mice (Cygb / )<br />

and primary hepatic stellate cells (HSCs) isolated from Cygb /<br />

or wild-type (WT) mice were characterized. A mouse NASH<br />

model was generated in Cygb / and WT mice by feeding a<br />

CDAA define diet for 8-32 weeks. A subset of mice was treated<br />

with the antioxidant N-acetylcysteine (NAC). Histopathology<br />

and gene expression were analyzed. Results: CYGB was<br />

expressed in HSCs of intact human livers, and its expression<br />

was reduced in NASH and HCC patients. In Cygb / mice, 67%<br />

of those aged 1 to 2 years spontaneously exhibited abnormalities<br />

and cancer development in multiple organs, including<br />

the liver, lung, lymph nodes and heart, compared with 4.7%<br />

in WT mice (p


306A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

192<br />

Circulating exosomes from healthy mice attenuate<br />

hepatic stellate cell activation and are anti-fibrotic in<br />

vivo<br />

Li Chen 1 , Ruju Chen 1 , Sherri Kemper 1 , David Brigstock 1,2 ; 1 Clinical<br />

& Translational Research, Nationwide Childrens Hospital, Columbus,<br />

OH; 2 Surgery, The Ohio State University, Columbus, OH<br />

Exosomes are 50-150 nm membranous vesicles that shuttle a<br />

complex molecular cargo (miRs, mRNAs, proteins) between<br />

producer and recipient cells resulting in epigenetic regulation<br />

of cell function. After their release from producer cells, exosomes<br />

traverse intercellular spaces and may either be taken<br />

up by neighboring cells or enter body fluids and be dispersed<br />

systemically. The goal of our work was to determine if hepatic<br />

stellate cells (HSC) are regulated by circulating exosomes.<br />

Exosomes were purified from the serum of healthy or fibrotic<br />

Swiss Webster male mice and tested for their effect on (i)<br />

activation or fibrogenesis in primary cultures of mouse HSC,<br />

or (ii) CCl 4<br />

-induced liver injury in mice. Exosomes (10μg/ml)<br />

from the serum of healthy mice resulted in decreased CTGF,<br />

αSMA or collagen α1(I) mRNA after treatment of D9 HSC for<br />

24hrs (p < 0.01) whereas transcript expression was increased<br />

or unaffected when the cells were exposed to serum exosomes<br />

from fibrotic mice. In a 10-day CCl 4<br />

injury model in<br />

male BAC reporter transgenic (TG) CTGF-EGFP FX156GSat/<br />

Mmucd Swiss Webster mice expressing GFP under the control<br />

of the CTGF promoter, hepatic GFP or αSMA expression was<br />

attenuated by i.p. administration every other day (40μg/g) of<br />

serum exosomes from healthy mice, but not from fibrotic mice<br />

(p < 0.01). In 5-wk CCl 4<br />

fibrosis models, i.p. administration of<br />

serum exosomes (40 μg/g) from healthy mice during the last 2<br />

wks of CCl 4<br />

treatment caused a dramatic and dose-dependent<br />

decrease in hepatic GFP-CTGF production. This was associated<br />

with decreased expression of CTGF, αSMA or collagen<br />

mRNA or protein (p < 0.01). The same effects occurred in<br />

wild type mice. To identify the principal miRs in serum exosomes<br />

that are differentially expressed between normal versus<br />

fibrotic mice, miR profiling was performed on serum exosomes<br />

using a miRnome miR PCR Array. Non-biased ranking of the<br />

data identified miR-23b, -34c, -106a, -151, -200b, -455, -483<br />

-532, -653, and -687 as the principal differentially expressed<br />

miRs and as candidate anti-fibrotics with the ability to attenuate<br />

expression of CTGF, αSMA or collagen in cultured HSC. These<br />

<strong>studies</strong> show that circulating exosomes from healthy individuals<br />

are instrinsically anti-fibrotic and that this property is attributable,<br />

at least in part, to specific miR constituents. We propose<br />

that circulating exosomes offer a novel platform for liver fibrosis<br />

therapy and that continued interrogation of their complex<br />

molecular payload will result in the identification of additional<br />

therapeutic candidates.<br />

Disclosures:<br />

David Brigstock - Stock Shareholder: FibroGen<br />

The following authors have nothing to disclose: Li Chen, Ruju Chen, Sherri Kemper<br />

193<br />

Combining high-throughput genome-wide RNAi and<br />

computational pharmacological screening identifies<br />

glyburide as a potent therapeutic candidate for liver<br />

disease due to alpha-1-antitrypsin deficiency<br />

Yan Wang, Murat C. Cobanoglu, Jie LI, Pamela Hale, Tunda Hidvegi,<br />

Micheal Ewing, Andrew Chu, Gary A. Silverman, Stephen C.<br />

Pak, Ivet Bahar, David H. Perlmutter; Departments of Pediatrics and<br />

Computational Biology, University of Pittsburgh School of Medicine,<br />

Pittsburgh, PA<br />

Alpha-1-antitrypsin deficiency (ATD) causes severe liver disease<br />

by the pathologic accumulation of the misfolded alpha-1-antitrypsin<br />

Z (ATZ) in hepatocytes. Liver transplantation is the only<br />

therapy currently available for this liver disease. In this study<br />

we investigated the possibility of discovering novel therapeutic<br />

drug candidates by interrogating a C. elegans model of ATD<br />

with genome-wide RNAi screening and computational (druggene<br />

interaction) pharmacological strategies. High-content<br />

genome-wide RNAi screening of our C. elegans model of ATD<br />

was utilized to identify genes that could modify the proteotoxicity<br />

of ATZ. Out of more than 16,000 RNAi, the initial screen<br />

identified 104 genes that are potential positive modifiers of<br />

ATZ proteotoxicity. These genes were then analyzed as potential<br />

targets of known drugs using the DrugBank 3.0 computational<br />

database. One of the drugs identified by this approach<br />

is glyburide (glibenclamide, GLB), a sulfonylurea drug that has<br />

been used broadly in clinical medicine as an oral hypoglycemic<br />

agent. In this study, we show that GLB mediates a marked<br />

decrease in steady-state levels of misfolded ATZ in a mammalian<br />

cell line model of ATD and that this effect is dose- and<br />

time-dependent with an optimal dose of 40 mM. Pulse-chase<br />

labeling experiments indicate that GLB specifically and selectively<br />

mediates an increase in the rate of intracellular degradation<br />

of ATZ. Mechanistic <strong>studies</strong> suggest that GLB increases<br />

autophagic flux but does not affect the phosphorylation of<br />

mTOR substrates and so its effect on the autophagy pathway is<br />

TOR-independent. As GLB is known to induce cellular calcium<br />

influx via action on the ATP-sensitive potassium channel, we<br />

investigated whether GLB could affect the intracellular accumulation<br />

of ATZ by acting on intracellular calcium signaling<br />

and found that calcium chelator BAPTA-AM partially blocks the<br />

GLB-mediated reduction of ATZ levels. Most importantly, GLB<br />

reduces hepatic ATZ load and hepatic fibrosis when administered<br />

systemically to the PiZ mouse model of ATD. Preliminary<br />

<strong>studies</strong> of GLB analogues suggest that its effects on degradation<br />

of ATZ can be dissociated from its effects on insulin secretion.<br />

In conclusion, these results indicate: 1) GLB reduces the hepatic<br />

proteotoxicity of ATZ accumulation by a calcium-dependent,<br />

mTOR-independent effect on intracellular degradation; 2) GLB<br />

(analogues) is a very appealing therapeutic drug candidate for<br />

ATD because of its wide margin of safety; 3) interrogation of<br />

a genetically tractable disease model by combining computational<br />

with high-content automated RNAi screening strategies is<br />

a valid and powerful drug discovery platform.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yan Wang, Murat C. Cobanoglu,<br />

Jie LI, Pamela Hale, Tunda Hidvegi, Micheal Ewing, Andrew Chu, Gary A.<br />

Silverman, Stephen C. Pak, Ivet Bahar, David H. Perlmutter


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 307A<br />

194<br />

Enteral Obeticholic Acid Prevents Hepatic Cholestasis in<br />

Total Parenteral Nutrition-Fed Neonatal Pigs<br />

Douglas Burrin 1 , Yanjun Jiang 1 , Zhengfeng Fang 2 , Barbara Stoll 1 ,<br />

Gregory J. Guthrie 1 , Hongtao Wang 3 , Ignacio R. Ipharraguerre 4,5 ,<br />

Jose J. Pastor 4 ; 1 USDA Children’s Nutrition Research Center,<br />

Department Pediatrics, Baylor College of Medicine, Houston,<br />

TX; 2 Animal Nutrition Institute, Sichuan Agricultural University,<br />

Chengdu, China; 3 Pediatric Gastroenterology, Hepatology, Nutrition,<br />

Department Pediatrics, Baylor College of Medicine, Houston,<br />

TX; 4 Lucta SA, Barcelona, Spain; 5 Institute of Human Nutrition and<br />

Food Science, University of Kiel, Kiel, Germany<br />

Total parenteral nutrition (TPN) is a vital support for neonatal<br />

infants with congenital or acquired gastrointestinal (GI) disorders<br />

and requiring small bowel resection. An adverse outcome<br />

associated with prolonged TPN use is parenteral nutrition<br />

associated cholestasis (PNAC). We previously showed that<br />

enteral chenodeoxycholic acid (CDCA) treatment reduced<br />

PNAC. We hypothesized that the protective effects of CDCA<br />

were mediated by modulation of FXR-target genes involved<br />

in bile acid homeostasis. The aim of the current study was to<br />

compare the physiological effects of a selective FXR agonist,<br />

obeticholic acid (OCA) vs CDCA on hepatic bile acid homeostasis<br />

in TPN-fed piglets. Term, newborn pigs were assigned<br />

to receive complete TPN (PN), TPN + enteral CDCA (30 mg/<br />

kg), or TPN+enteral OCA (0.5, 5, 15 mg/kg) daily for 19 d.<br />

The daily parenteral lipid was Intralipid given at 10 g/kg. Endpoints<br />

of PNALD and bile acid homeostasis were measured.<br />

We found that, compared to PN pigs, treatment with high dose<br />

of OCA (OCA5 and OCA15), but not CDCA and low dose<br />

of OCA (OCA0.5), reduced serum PNAC markers including<br />

bilirubin, gamma-glutamyl transferase (GGT), total bile acid, triglyceride,<br />

and very-low-density lipoprotein. Compared to PN,<br />

OCA 5 and 15, but not OCA 0.5 or CDCA, reduced the total<br />

plasma bile acid concentration by increasing the proportional<br />

transfer of hepatic bile acid into the gallbladder, suggesting<br />

increased bile flow. TPN-induced ductopenia as measured by<br />

percentage of intact bile ducts per portal tract was prevented<br />

by OCA suggesting preservation of bile ducts. The major bile<br />

acids in plasma were glyco-conjugated forms of CDCA, hyocholic<br />

acid and hyodeoxycholic acid. OCA5 and OCA15, but<br />

not CDCA, suppressed hepatic expression of CYP7A1, while<br />

CYP27A1 and CYP8B1 mRNA remained unchanged. The bile<br />

acid detoxification enzyme CYP3A29 mRNA was inhibited by<br />

both CDCA and OCA treatments. OCA, but not CDCA upregulated<br />

hepatic mRNA involved in hepatobiliary bile acid and bilirubin<br />

transport into bile including bile salt export pump (BSEP),<br />

multidrug resistance protein 1(MDR1), and multidrug resistance<br />

protein 4 (MRP4). OCA5 and OCA15 induced hepatic and<br />

ileal FGF19 expression more than CDCA in pigs. We further<br />

found that OCA5 and OCA15, but not CDCA, inhibited the<br />

hepatic expression of inflammatory marker interleukin-8. Contrary<br />

to our previous study, we found that CDCA did not prevent<br />

PNAC. We suspect that this was due to the higher lipid<br />

load infused. We conclude that enteral OCA is more effective<br />

than CDCA in prevention of PNAC via upregulation of FXR-target<br />

genes involved in preservation of hepatobiliary transporters<br />

and bile duct function.<br />

Disclosures:<br />

The following authors have nothing to disclose: Douglas Burrin, Yanjun Jiang,<br />

Zhengfeng Fang, Barbara Stoll, Gregory J. Guthrie, Hongtao Wang, Ignacio R.<br />

Ipharraguerre, Jose J. Pastor<br />

195<br />

Biliatresone causes GSH reduction in cholangiocytes,<br />

leading to cholangiocyte damage, increased monolayer<br />

permeability, and periductular myofibroblasts: insights<br />

into the mechanism of biliary atresia<br />

Orith Waisbourd-Zinman 1 , Hong Koh 2,3 , Pierre-Marrie Lavrut 2,4 ,<br />

Shannon Tsai 2 , John R. Porter 5 , Michael Pack 2 , Rebecca G.<br />

Wells 2 ; 1 Gastroenterology, Hepatology and Nutrition, Children’s<br />

Hospital of Philadelpahia, Philadelpahia, PA; 2 Gastroenterology,<br />

Perelman School of Medicine, University of Pennsylvania, Philadelphia,<br />

PA; 3 Department of Pediatrcis, Division of Gastroenterology,<br />

Hepatology and Nutrition, Yonsei University College of Medicine,<br />

Severance Children’s Hospital, Seoul, Korea (the Republic of);<br />

4 Pathology, Academic Hospital of Lyon, Lyon, France; 5 Biological<br />

Sciences, University of the Sciences, Philadelphia, PA<br />

Background: Biliary atresia (BA) is an acquired fibro-obliterative<br />

disease of unknown etiology affecting the extrahepatic<br />

bile ducts of the newborn. We previously reported that the<br />

novel toxin biliatresone causes selective atresia of the extrahepatic<br />

biliary tree in zebrafish and microtubule instability and<br />

lumen obstruction in a cholangiocyte 3D culture model. The<br />

toxin strongly and spontaneously binds to glutathione (GSH).<br />

We hypothesized that GSH reduction is responsible for the<br />

effects of the toxin. Methods: We used several mouse cholangiocyte<br />

culture systems: 2D culture, 3D spheroid culture,<br />

and ex vivo culture of neonatal extra-hepatic bile ducts. We<br />

treated these with biliatresone and compounds that regulate<br />

GSH, then stained and imaged the cells. We measured<br />

GSH levels in cholangiocytes treated with the compounds at<br />

different time points. In order to assess lumen permeability,<br />

we used a rhodamine efflux assay with live cell imaging of<br />

cholangiocytes in 3D cultures. Results: Cholangiocytes in 2D<br />

culture demonstrate a 70% decrease in GSH 1h following<br />

toxin treatment, with recovery to baseline by 6h. In 3D culture,<br />

lowering GSH with buthionine sulfoximine (BSO) mimicked<br />

the effects of biliatresone, with microtubule instability, loss of<br />

polarity, and lumen obstruction. N-Acetyl-L-cysteine (L-NAC),<br />

which increases GSH, protected 3D cultures from the effects<br />

of biliatresone. Neonatal bile ducts cultured ex vivo showed<br />

patchy disruption of the cholangiocyte monolayer and lumen<br />

obstruction after treatment with biliatresone or BSO, with<br />

protection by L-NAC but not the stereoisomer D-NAC. A time<br />

course study of cholangiocytes in 3D culture treated with biliatresone<br />

showed that microtubule injury occurred first, within 3<br />

hours, followed by abnormalities in F-actin, polarity, and tight<br />

junctions, and finally lumen obstruction. Rhodamine remained<br />

in the lumens of vehicle-treated cholangiocyte spheroids for up<br />

to 12 h, but effluxed within 3 h after biliatresone treatment,<br />

suggesting increased epithelial permeability. In the ex vivo bile<br />

duct model, staining for a-smooth muscle actin was increased<br />

in the stroma surrounding the cholangiocyte layer after treatment<br />

with biliatresone, BSO, or D-NAC plus biliatresone, but<br />

remained at baseline levels after treatment with L-NAC plus biliatresone.<br />

Conclusions: Biliatresone rapidly decreases GSH in<br />

mouse cholangiocyte and bile duct cultures, initially disrupting<br />

microtubules then compromising cholangiocyte polarity and<br />

increasing lumen permeability. In vivo, this is a potential cause<br />

of obstruction and periductular fibrosis and may be important<br />

to the pathophysiology of BA.<br />

Disclosures:<br />

The following authors have nothing to disclose: Orith Waisbourd-Zinman, Hong<br />

Koh, Pierre-Marrie Lavrut, Shannon Tsai, John R. Porter, Michael Pack, Rebecca<br />

G. Wells


308A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

196<br />

A functional classification of ABCB4 missense variations<br />

identified in progressive familial intrahepatic cholestasis<br />

type 3<br />

Jean-Louis Delaunay 1 , Anne-Marie Durand-Schneider 1 , Claire Dossier<br />

1 , Thomas Falguières 1 , Julien Gautherot 1 , Tounsia Aït-Slimane 1 ,<br />

Chantal Housset 1 , Emmanuel Jacquemin 2 , Michèle Maurice 1 ;<br />

1 UMR_S 938, UPMC & Inserm, Paris Cedex 12, France; 2 Hépatologie<br />

Pédiatrique, APHP, Hôpital Bicêtre, Kremlin-Bicêtre, France<br />

Progressive familial intrahepatic cholestasis type 3 (PFIC3) is<br />

caused by bi-allelic ABCB4 variations. More than 70 % of<br />

disease-causing ABCB4 variations are missense variations.<br />

The aims of the study were i) to propose a functional classification<br />

of ABCB4 missense variations; ii) to test the rescue of<br />

trafficking-defective mutants by cyclosporins. Methods: Twelve<br />

single-nucleotide missense variations that were identified in 9<br />

PFIC3 patients (6 homozygous, 3 composite heterozygous)<br />

and located in or close to transmembrane domains (F357L,<br />

S346I, P726L, T775M, Q855L, G954S), first intracellular loop<br />

(T175A), first nucleotide-binding domain (T424A, N510S,<br />

I541F, L556R), and linker region (R652G), were reproduced in<br />

the ABCB4 cDNA. The mutants thus obtained were expressed<br />

in HepG2 (transiently) and HEK 293 cells (stably), and compared<br />

with wild type ABCB4, to determine their impact on the<br />

ABCB4 expression or phosphatidylcholine transport activity,<br />

measured by a fluorimetric assay. Results: Four mutants were<br />

either fully (I541F and L556R) or partially (S346I and Q855L)<br />

retained in the endoplasmic reticulum, in an immature endoglycosidase<br />

H-sensitive form, of lower molecular weight than<br />

the mature N-glycosidase F-sensitive form. Phosphatidylcholine<br />

secretion by these mutants was less than 10 % that of wild<br />

type, except for Q855L (30 %). Rescue of these trafficking<br />

defects, i.e. exit from the endoplasmic reticulum and increase<br />

of the mature form at the bile canaliculi, was obtained by cell<br />

treatments with cyclosporins A or C, and to a lesser extent<br />

B, D or H. Four mutations with little or no effect on ABCB4<br />

expression at the bile canaliculi, caused a significant reduction<br />

(F357L, T775M and G954S) or absence (P726L) of phosphatidylcholine<br />

secretion. Two mutants (T424A and N510S) were<br />

normally processed and expressed at the bile canaliculi but<br />

their stability, assessed by protein decay, was reduced. We<br />

found no defect of the T175A mutant, nor of R652G, previously<br />

described as a polymorphism, which herein co-existed<br />

with another variant (G954S) on the same allele, in two siblings.<br />

In patients, the most severe phenotypes appreciated by<br />

the duration of transplant-free survival, were caused by ABCB4<br />

variants that were markedly retained in the endoplasmic reticulum,<br />

and expressed in a homozygous status. In conclusion,<br />

ABCB4 variations can be classified as follows: nonsense variations<br />

(I), and on the basis of current findings, missense variations<br />

that primarily affect the traffic (II), activity (III) or stability<br />

(IV) of the protein. Among missense variants, those in class II<br />

cause the most severe clinical phenotypes and can be rescued<br />

by pharmacological therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jean-Louis Delaunay, Anne-Marie<br />

Durand-Schneider, Claire Dossier, Thomas Falguières, Julien Gautherot, Tounsia<br />

Aït-Slimane, Chantal Housset, Emmanuel Jacquemin, Michèle Maurice<br />

197<br />

Liver-specific deletion of the insulin receptor profoundly<br />

reduces accumulation and proteotoxicity of misfolded<br />

α 1<br />

-antitrypsin Z (ATZ) in a mouse model<br />

Andrew Chu 1 , Tunda Hidvegi 1 , Souvik Chakraborty 1 , Micheal<br />

Ewing 1 , Amitava Mukherjee 1 , Patrick Araya 2 , Pamela Hale 1 ,<br />

Christine Dippold 1 , Yan Wang 1 , Jie LI 1 , Evelyn Akpadock 1 , Hou<br />

Ming Cai 1 , Stephen C. Pak 1 , Donna B. Stolz 2 , Gary A. Silverman<br />

1 , David H. Perlmutter 1 ; 1 Pediatrics - Gastroenterology, Children’s<br />

Hospital of Pittsburgh of UPMC; University of Pittsburgh,<br />

Pittsburgh, PA; 2 University of Pittsburgh, Pittsburgh, PA<br />

Alpha-1-antitrypsin deficiency (ATD)-associated liver disease<br />

is an important cause of cirrhosis and hepatocellular carcinoma,<br />

with liver transplantation currently the only treatment<br />

for advanced disease. Insulin signaling is a key regulator of<br />

proteostasis mechanisms, with inhibition of insulin signaling<br />

linked to increased longevity in animals. The purpose of our<br />

study was to determine whether abrogation of insulin signaling<br />

reduces ATZ accumulation and proteotoxicity in a mouse<br />

model of ATD-associated liver disease. To investigate the effect<br />

of loss of insulin signaling, we generated double transgenic<br />

mice (IRZZ) by crossing liver-specific insulin receptor knockout<br />

mice (LIRKO) to the PiZ mouse model of ATD-associated<br />

liver disease. Histological analysis of livers from male mice<br />

at 6 months and 12-14 months demonstrated a 60 percent<br />

reduction of intracellular ATZ globules by PAS-D staining and<br />

a 50 percent reduction in hepatic fibrosis by Sirius Red staining<br />

in IRZZ mice compared with PiZ mice. There was also a<br />

complete elimination of nodular regeneration in the IRZZ liver.<br />

This is consistent with results from a C. elegans model of ATD<br />

in which RNAi corresponding to several intermediates in the<br />

insulin signaling pathway reduce intracellular ATZ accumulation.<br />

To determine the mechanism of reduced hepatic ATZ load<br />

in the IRZZ mice, we carried out pulse-chase radiolabelling on<br />

isolated hepatocytes and found a significant increase in intracellular<br />

degradation of ATZ in the IRZZ compared to the PiZ<br />

hepatocytes. To determine whether specific gene expression<br />

signatures or signaling pathways were involved in the effect<br />

of insulin receptor deletion, we used transcriptomic analysis to<br />

compare hepatic RNA levels in IRZZ to PiZ and control (LIRKO<br />

and wild-type) mice. Compared to PiZ, the gene expression<br />

pattern in IRZZ mice showed downregulation of collagen V<br />

consistent with the marked decrease in Sirius Red staining,<br />

upregulation of genes associated with hepatocellular proliferation/regeneration<br />

(prominin 1 and 2), and upregulation of<br />

genes involved in lysosomal function (lysosomal ATPase, H +<br />

transporting, V0 subunit D isoform, Atp6v0d2). Furthermore,<br />

using immunofluorescence for the lysosomal membrane protein<br />

LAMP2 and transmission electron microscopy, we found a<br />

marked increase in number of lysosomes in the IRZZ as compared<br />

to the PiZ liver, suggesting activation of the autophagolysosomal<br />

system in the IRZZ model. These results indicate that<br />

the insulin signaling pathway suppresses a key mechanism by<br />

which the liver attempts to protect itself from misfolded ATZ and<br />

suggest that activation of the autophagolysosomal system is the<br />

target of this mechanism.<br />

Disclosures:<br />

The following authors have nothing to disclose: Andrew Chu, Tunda Hidvegi,<br />

Souvik Chakraborty, Micheal Ewing, Amitava Mukherjee, Patrick Araya, Pamela<br />

Hale, Christine Dippold, Yan Wang, Jie LI, Evelyn Akpadock, Hou Ming Cai,<br />

Stephen C. Pak, Donna B. Stolz, Gary A. Silverman, David H. Perlmutter


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 309A<br />

198<br />

Anti-inflammatory and Antifibrotic Activity of NGM282,<br />

A Novel Variant of FGF19, in an Mdr2-Deficient Mouse<br />

Model of Primary Sclerosing Cholangitis<br />

Lei Ling, Mei Zhou, Hong Yang, Marc Learned, Stephen Rossi,<br />

Alex M. DePaoli, Hui Tian; NGM Biopharmaceuticals, Inc., South<br />

San Francisco, CA<br />

Background and Aims: Primary sclerosing cholangitis (PSC) is<br />

a chronic inflammatory biliary disease characterized by periductal<br />

fibrosis and stricture formation. Recent data showed<br />

decreased expression and intraductular production of FGF19,<br />

which may worsen pre-existing inflammation and fibrosis.<br />

NGM282 is an engineered variant of FGF19 that inhibits<br />

Cyp7a1 but lacks the tumorigenic activity seen with FGF19.<br />

Mice deficient in phospholipid flippase multidrug resistant<br />

protein 2 (Mdr2-/-) are a preclinical model for PSC as they<br />

develop liver histopathology similar to human PSC. We studied<br />

FGF19 and NGM282 in Mdr2-/- mice to characterize the<br />

potential biologic activity in PSC. Methods: 12wk old Mdr2-/-<br />

mice with established hepatobiliary disease received 1 dose of<br />

adeno-associated virus carrying NGM282, FGF19 or control.<br />

Serum liver enzymes were analyzed at 4 and 24wks post-dose.<br />

Liver tissue was collected at 24wks and stained with Hematoxylin<br />

& Eosin and Sirius Red. Results: Significant decreases in<br />

alkaline phosphatase (ALP) were seen in male and female mice<br />

at 4wks with FGF19 (males=349+26 U/L to 107+16 U/L;<br />

females=565+49 U/L to 98+7 U/L; p


310A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

The following authors have nothing to disclose: Ryan B. Perumpail, Andy Liu,<br />

Channa R. Jayasekera, Robert J. Wong<br />

the 300 mg/kg arm compared to 8 (29%) reinfections in the<br />

control. Conclusions: The majority of LT recipients have CP B or<br />

C cirrhosis pre-LT and frequent renal dysfunction early post-LT.<br />

Despite these medical complexities, preliminary results suggest<br />

that Civacir is well-tolerated in the peri-transplant setting and<br />

effective (at the 300 mg/kg dose) in preventing HCV reinfection.<br />

200<br />

Prevention of Recurrent Hepatitis C in Liver Transplant<br />

(LT) Recipients with Civacir ® : Preliminary Results on<br />

Safety and Efficacy, Including Patients with Renal/Liver<br />

Dysfunction<br />

Norah Terrault 1 , Roshan Shrestha 2 , Sanjaya K. Satapathy 3 , Sher<br />

Linda 4 , Jens Rosenau 5 , Jacqueline G. O’Leary 6 , Thomas D. Schiano<br />

7 , Lewis W. Teperman 8 , James Spivey 9 , Jeffrey Campsen 10 ,<br />

John M. Vierling 11 , Elizabeth C. Verna 12 , David W. Victor 13 ,<br />

George Therapondos 14 , Kalyan R. Bhamidimarri 15 , Seraphin<br />

Kuate 16 , Gregory Osgood 16 , Judith Blacklidge 16 , Nicole<br />

Daelken 16 , Shailesh Chavan 16 ; 1 University of California San Francisco,<br />

San Francisco, CA; 2 Piedmont Transplant Institute, Atlanta,<br />

GA; 3 University of Tennessee Health Sciences Center, Memphis,<br />

TN; 4 University of Southern California, Los Angeles, CA; 5 University<br />

of Kentucky, Lexington, KY; 6 Baylor University Medical Center,<br />

Dallas, TX; 7 Mount Sinai Medical Center, New York, NY;<br />

8 New York University, New York, NY; 9 Emory University, Atlanta,<br />

GA; 10 University of Utah, Salt Lake City, UT; 11 Baylor College of<br />

Medicine, Houston, TX; 12 Columbia University, New York, NY;<br />

13 Houston Methodist Hospital, Houston, TX; 14 Ochsner Clinic<br />

Foundations, New Orleans, LA; 15 University of Miami, Miami, FL;<br />

16 Biotest Pharmaceuticals Corporation, Boca Raton, FL<br />

Background and Aims: Prevention of hepatitis C virus (HCV)<br />

recurrence post-LT is important as it eliminates the risk of severe<br />

or progressive disease and reduces the complexity of post-LT<br />

management. Treatment in the peri-transplant period is especially<br />

difficult in patients with severe renal impairment, which<br />

often restricts antiviral therapy (AVT). Safe use of hepatitis B<br />

immune globulin for the prevention of hepatitis B recurrence<br />

in post-LT patients is well established with >30 years of experience.<br />

Civacir ® a hepatitis C immune globulin contains a broad<br />

variety of antibodies against HCV and may offer a safe prophylactic<br />

approach in the peri-transplant period. Methods: Study<br />

988 is an ongoing open-label, randomized phase III clinical<br />

trial evaluating Civacir 200 or 300 mg/kg vs. standard of care<br />

(no AVT treatment post-LT) in HCV-infected patients (any genotype)<br />

awaiting LT. Wait-listed patients receiving pre-LT AVT for<br />

≤24 weeks and with HCV RNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 311A<br />

patients with end-stage liver disease secondary to chronic HCV<br />

infection will reach age 60 and require LT. While previous<br />

<strong>studies</strong> have evaluated trends in HCV-related LT, none have<br />

specifically investigated HCV-related LT among the BB cohort<br />

age > 60. Methods: We utilized the United Network for Organ<br />

Sharing registry to evaluate the impact of chronic HCV infection<br />

on long term survival following LT among BB patients age > 60<br />

from 2003-2012 (MELD era). Overall patient survival following<br />

LT was analyzed with Kaplan Meier methods and multivariate<br />

Cox proportional hazards adjusted for age, sex, hepatocellular<br />

carcinoma, race/ethnicity, year of LT, diabetes mellitus,<br />

MELD score, albumin, ascites, and hepatic encephalopathy.<br />

Results: Overall, among patients age > 60, representing 15%<br />

of patients, survival following LT was significantly lower in the<br />

HCV BB cohort compared to the non-HCV BB cohort (5-year<br />

survival: 64.5% vs 77.4%, p


312A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

203<br />

Long-term complete prophylaxis withdrawal in liver-transplanted<br />

patients for HBV-related cirrhosis is safe<br />

and frequently associated with spontaneous ANTI-Hbs<br />

seroconversion<br />

Ilaria Lenci 1,3 , Daniele Di Paolo 1 , Martina Milana 1 , Francesco<br />

Santopaolo 1 , Leonardo Baiocchi 1 , Valentina Svicher 2 , Laura Tariciotti<br />

3 , Giuseppe Tisone 3 , Carlo Federico Perno 2 , Mario Angelico 1 ;<br />

1 Liver Unit, Tor Vergata University, Rome, Italy; 2 Virology Unit, Tor<br />

Vergata University, Rome, Italy; 3 Liver Transplant Unit, Tor Vergata<br />

University, Rome, Italy<br />

Background and aims. Long-term use of hepatitis-B-virus (HBV)<br />

specific immunoglobulins (HBIg) and/or nucleoside analogs<br />

(NA) is recommended to prevent HBV reinfection in HBsAg<br />

positive liver transplant (LT) recipients. Searching for a more<br />

rational and tailored approach, we showed (J. Hepatol. 2011;<br />

55:587-593) that weaning of HBV prophylaxis was feasible<br />

and safe in a selected group of LT recipients considered at<br />

“low risk” of HBV recurrence (undetectable HBV-DNA at LT<br />

and undetectable intra-hepatic total HBV-DNA and ccc-DNA<br />

5 years after LT), with small percentage of HBV recurrence, in<br />

the absence of clinically relevant events during the first 2 years<br />

after HBIg and NA withdrawal. We report here the extended<br />

follow-up of this selected group of LT recipients, up to 6 years<br />

(median months 78) after complete prophylaxis withdrawal.<br />

Methods. The study included 30 LT recipients, all HBeAg negative,<br />

with HBV genotype D, including 6 with HCV and 7 with<br />

HDV coinfections, who initially received HBIg and NA (mostly<br />

lamivudine) for at least 5 years following LT. Sequential HBIg<br />

and NA withdrawal was performed in two six-month periods<br />

under strict monitoring of HBV virology, based on monthly<br />

blood tests and liver biopsies every 6 months for the first two<br />

years. All patients were then followed with clinical and virological<br />

test at 3-6 months intervals. Results. Only two patients<br />

(6.7%) underwent sustained HBV recurrence, with HBsAg and<br />

HBV-DNA reappearance after 45 and 78 months of prophylaxis<br />

withdrawal. Both received tenofovir therapy and promptly<br />

returned HBV-DNA negative and did not experience any clinical<br />

event. Four patients showed a transient HBsAg positivity (2,<br />

4, 6 and 45 months after HBIg and NA withdrawal, respectively),<br />

in one case with transient HBV-DNA detectable, followed<br />

by later anti-HBs seroconversion. Additional 14 patients<br />

mounted a measurable anti-HBs titer during the long-term follow<br />

up, without showing transient HBsAg/HBV-DNA detectability.<br />

Thus, 28 (93.3%) patients were HBsAg and HBV-DNA negative<br />

and 18 (60%) were anti-HBs positive at completion of the<br />

extended follow up. The average lastly detected anti-HBs titer<br />

was 276 IU/L), with 2 patients showing titers >1000 IU/L.<br />

Anti-HBs seroconversion occurred after a median of 70 months<br />

(range: 32-70) after complete prophylaxis withdrawal and was<br />

unrelated to HCV or HDV coinfections. Conclusion. Complete<br />

prophylaxis withdrawal is safe in patients liver-transplanted<br />

for HBV-related disease at low risk of recurrence and in the<br />

majority of cases is followed by spontaneous anti-HBs seroconversion.<br />

Disclosures:<br />

Mario Angelico - Advisory Committees or Review Panels: Gilead, Janssen;<br />

Grant/Research Support: Roche; Speaking and Teaching: GSK, Roche, Gilead,<br />

Novartis, BMS, Bayer<br />

The following authors have nothing to disclose: Ilaria Lenci, Daniele Di Paolo,<br />

Martina Milana, Francesco Santopaolo, Leonardo Baiocchi, Valentina Svicher,<br />

Laura Tariciotti, Giuseppe Tisone, Carlo Federico Perno<br />

204<br />

Impact of Sofosbuvir-Based Regimens on Renal Function<br />

in Liver Transplant Recipients: Results of a Multicenter<br />

Study<br />

Nabiha Faisal 1 , Eberhard L. Renner 1 , Marc Bilodeau 2 , Bandar<br />

Aljudaibi 3 , Geri Hirsch 4 , Eric M. Yoshida 5 , Tarana Hussaini 5 ,<br />

Peter Ghali 6 , Stephen E. Congly 7 , Mang M. Ma 8 , Jennifer Leonard<br />

10 , Curtis Cooper 9 , Kevork M. Peltekian 4 , Nazia Selzner 1 , Les<br />

Lilly 1 ; 1 Multi organ Transplant, Toronto General Hospital, Toronto,<br />

ON, Canada; 2 University of Montreal, Montreal, QC, Canada;<br />

3 London Health Sciences, University of Western Ontario, London,<br />

ON, Canada; 4 Dalhousie University, Halifax, NF, Canada; 5 University<br />

of British Columbia, Vancouver, BC, Canada; 6 University of<br />

McGill, Montreal, QC, Canada; 7 University of Calgary, Calgary,<br />

AB, Canada; 8 University of Alberta, Edmonton, AB, Canada; 9 University<br />

of Ottawa, Ottawa, QC, Canada; 10 Memorial University,<br />

St. John’s, NF, Canada<br />

BACKGROUND & AIMS: The first nucleotide NS5B polymerase<br />

inhibitor, sofosbuvir (SOF), is not recommended for patients<br />

with severe renal impairment due to pharmacokinetic <strong>studies</strong><br />

demonstrating higher serum drug and metabolite levels in such<br />

patients. However, renal function in cirrhotic and non-cirrhotic<br />

patients treated with SOF-based regimens is often compromised<br />

yet poorly assessed. This analysis aims to assess the<br />

impact of SOF-based regimens on renal function in liver transplant<br />

(LT) recipients with recurrent HCV, and to assess the role<br />

of renal function on the efficacy and safety of these regimens.<br />

METHODS: 165 LT recipients across Canada with HCV recurrence<br />

were treated with SOF-based regimens from January<br />

2014 to May 2015. Mean patient age was 58±6.85 years;<br />

the majority were male, GT1, Caucasian and treatment-experienced.<br />

One third had aggressive HCV in the graft; 50% had<br />

F3/4 fibrosis. The majority were on tacrolimus based immunosuppression.<br />

Median time from LT to treatment was 27 (1-309)<br />

months. Baseline eGFR was calculated by Modification of Diet<br />

in Renal Disease formula (MDRD) and patients were stratified<br />

into 3 groups: eGFR60 ml/min. The primary<br />

outcome was post treatment changes in renal function from<br />

baseline. Secondary outcomes included sustained virological<br />

response at 12 weeks after treatment end (SVR12), anemia<br />

related adverse events (AEs), need for blood transfusions or<br />

growth factors, serious AEs, treatment discontinuation due to<br />

AE, or death. RESULTS: An improvement in post treatment renal<br />

function was seen in the majority (58%) of patients, mostly in<br />

the SVR12 group (81% vs. 19%). A decline in renal function<br />

seen in 22%; more marked in poorest renal function (eGFR<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 313A<br />

Marc Bilodeau - Advisory Committees or Review Panels: Verlyx, Bayer, Astellas;<br />

Consulting: GSK; Grant/Research Support: Merck, Synageva; Speaking and<br />

Teaching: Merck, Vertex, Abbvie, Aptalis, Roche<br />

Eric M. Yoshida - Advisory Committees or Review Panels: Hoffman LaRoche, Gilead<br />

Sciences Inc, Abbvie; Grant/Research Support: Abbvie, Hoffman LaRoche,<br />

Merck Inc, Vertex Inc, Jannsen Inc, Gilead Sciences Inc, Boeringher Ingleheim<br />

Inc, Astellas; Speaking and Teaching: Gilead Sciences Inc, Merck Inc<br />

Curtis Cooper - Advisory Committees or Review Panels: Gilead, Abbvie, MERCK;<br />

Grant/Research Support: MERCK, Gilead, Abbvie; Speaking and Teaching:<br />

MERCK, Abbvie, Gilead<br />

The following authors have nothing to disclose: Nabiha Faisal, Bandar Aljudaibi,<br />

Geri Hirsch, Tarana Hussaini, Peter Ghali, Stephen E. Congly, Mang M. Ma,<br />

Jennifer Leonard, Kevork M. Peltekian, Nazia Selzner, Les Lilly<br />

205<br />

A Randomized Controlled Trial of Sofosbuvir/GS-5816<br />

Fixed Dose Combination for 12 Weeks Compared to<br />

Sofosbuvir with Ribavirin for 12 Weeks in Genotype 2<br />

HCV Infected Patients: The Phase 3 ASTRAL-2 Study<br />

Mark S. Sulkowski 1 , Norbert Brau 2 , Eric Lawitz 3 , Mitchell L. Shiffman<br />

4 , William J. Towner 5 , Peter J. Ruane 6 , Brian Doehle 7 , Jing<br />

Wang 7 , John McNally 7 , Anu Osinusi 7 , Diana M. Brainard 7 , John<br />

G. McHutchison 7 , Nancy Reau 8 , Keyur Patel 9 ; 1 Johns Hopkins University,<br />

Baltimore, MD; 2 James J. Peters VA Medical Center, Bronx,<br />

NY and Icahn School of Medicine at Mount Sinai, New York, NY;<br />

3 Texas Liver Institute, University Of Texas Health Science Center,<br />

San Antonio, TX; 4 Liver Institute of Virginia, Richmond, VA; 5 Kaiser<br />

Permanente, Los Angeles, CA; 6 Ruane Medical, Los Angeles, CA;<br />

7 Gilead Sciences, Inc., Foster City, CA; 8 University of Chicago<br />

Medical Center, Chicago, IL; 9 Duke University School of Medicine,<br />

Durham, NC<br />

Introduction: GS-5816 is a pangenotypic HCV NS5A inhibitor<br />

that has demonstrated high efficacy in genotype 1-6 HCV-infected<br />

patients when administered for 12 weeks in combination<br />

with sofosbuvir. The Phase 3 ASTRAL-2 study compared the<br />

efficacy and safety of treatment with a fixed dose combination<br />

(FDC) of SOF/GS-5816 for 12 weeks to SOF + RBV for 12<br />

weeks in treatment-naïve and treatment-experienced genotype<br />

2 HCV-infected patients, with and without cirrhosis. Methods:<br />

Enrolled patients were randomized 1:1 to receive either SOF/<br />

GS-5816 (400 mg /100 mg daily) FDC for 12 weeks or SOF<br />

(400mg daily) with RBV (1000-1200mg daily) for 12 weeks.<br />

Randomization was stratified by prior treatment and cirrhosis<br />

status. HCV RNA was measured with the CAP/CTM HCV 2.0<br />

assay with LLOQ =15 IU/mL. The primary endpoint was sustained<br />

virologic response 12 weeks after treatment (SVR12)<br />

with a pre-specified non-inferiority margin of 10%. All patients<br />

have completed treatment and are in follow-up. Results: 266<br />

patients with genotype 2 HCV infection were randomized and<br />

treated, 134 with SOF/GS-5816 and 132 with SOF+RBV.<br />

Overall 59% were male, 88% were white, 38% had IL28B<br />

CC genotype, 15% had prior treatment failure, and 14% had<br />

compensated cirrhosis. HCV RNA declined rapidly in both<br />

treatment groups with > 90% of patients having HCV RNA<br />

< LLOQ at treatment week 4. The SVR4 rate for the SOF/<br />

GS-5816 treated patients is 99% (133/134) and for the<br />

SOF+RBV treated patients is 96% (127/132). There were no<br />

patients with relapse in the SOF/GS-5816 group compared<br />

with 4 patients with relapse in the SOF+RBV treated group<br />

through posttreatment week 4. Complete SVR12 and resistance<br />

results will be presented. Two patients discontinued treatment,<br />

one treated with SOF/GS-5816 discontinued on Day 1 for<br />

adverse events (anxiety, headache and disturbance in attention)<br />

and one treated with SOF+RBV did not return for study<br />

visits after week 10. The most common AEs were fatigue,<br />

headache, nausea, and insomnia and all were more frequently<br />

observed in the SOF+RBV treated patients. Two patients in<br />

each treatment group experienced SAEs, none were considered<br />

related to study drugs. Hemoglobin declined to < 10g/dL<br />

in 6 (5%) patients taking SOF+RBV but not in patients taking<br />

SOF/GS-5816. No other significant laboratory abnormalities<br />

were observed. Conclusions: Treatment with SOF/GS-5816<br />

FDC for 12 weeks resulted in high SVR4 rates and was well tolerated<br />

in treatment-naïve and treatment-experienced genotype<br />

2 HCV-infected patients with and without cirrhosis. Few (


314A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

206<br />

Daclatasvir plus sofosbuvir with or without ribavirin in<br />

genotype 3 patients from a large French multicenter<br />

compassionate use program<br />

Christophe Hezode 1 , Victor de Ledinghen 2 , Hélène Fontaine 3 ,<br />

Fabien Zoulim 4 , Pascal Lebray 5 , Nathalie Boyer 6 , Dominique<br />

G. Larrey 7 , Christine Silvain 8 , Danielle Botta-Fridlund 9 , Vincent<br />

Leroy 10,11 , Marc Bourlière 12 , Louis d’Alteroche 13 , Isabelle<br />

Fouchard-Hubert 14 , Dominique Guyader 15 , Isabelle Rosa 16 , Eric<br />

Nguyen-Khac 17 , Vincent Di Martino 18 , Fabrice Carrat 19 , Larysa<br />

Fedchuk 20 , Raoudha Akremi 20 , Yacia Bennai 20 , Jean-Pierre Bronowicki<br />

21 ; 1 Department of Hépato-Gastroenterology, CHU Henri<br />

Mondor, Créteil, France; 2 Liver fibrosis investigational center, Haut-<br />

Lévêque Hospital, Bordeaux University Hospital, Pessac, France;<br />

3 Department of Hepatology, Cochin Hospital, APHP, Paris, France;<br />

4 Department of Hepatology, La Croix-Rousse Hospital, Lyon Civil<br />

Hospices, Lyon, France; 5 Department of Hepato-Gastroenterology,<br />

Pitié Salpêtrière Hospital, Paris, France; 6 Department of Hepatology,<br />

Beaujon Hospital, APHP, Paris, France; 7 Department of<br />

Hepato-Gastroenterology, Saint-Eloi Hospital, CHU Montpellier,<br />

Montpellier, France; 8 Department of Hepato-Gastroenterology<br />

and of nutritional support, epithelial tissues and cytokines laboratory<br />

EA 4331, CHU Poitiers, Poitiers, France; 9 Department<br />

of Hepato-Gastroenterology, Conception Hospital, Marseille,<br />

France; 10 Grenoble University hospital, Hepato-Gastroenterology<br />

University Clinic, Grenoble, France; 11 Inserm, 823 Unity, Grenoble,<br />

France; 12 Department of Hepatology and Gastroenterology,<br />

Saint-Joseph Hospital, Marseille, France; 13 Unit of Hepatology,<br />

Department of Hepato-Gastroenterology, CHU Trousseau, Tours,<br />

France; 14 Department of Hepato-Gastroenterology, CHU Angers,<br />

Angers, France; 15 Department of liver diseases, CHU Rennes,<br />

Rennes 1 University, Rennes, France; 16 Department of Hepatology<br />

and Gastroenterology, CHI Créteil, Créteil, France; 17 Department<br />

of Hepato-Gastroenterology, CHU Amiens Nord, Amiens, France;<br />

18 Department of Hepatology and of digestive intensive care unit,<br />

CHRU Jean Minjoz, Besançon, France; 19 Inserm UMR-S 707,<br />

Pierre-et-Marie-Curie University Paris 6, Paris, France; 20 Bristol-Myers<br />

Squibb Research and Development, Rueil-Malmaison, France;<br />

21 INSERM Unité 954, CHU Nancy and Lorraine University, Vandoeuvre-lès-Nancy,<br />

France<br />

Background and aims. Treatment options for HCV genotype 3<br />

(GT3) patients are limited. The combination daclatasvir (DCV)<br />

and sofosbuvir (SOF) for 12 weeks is associated with high<br />

SVR 12<br />

rate in genotype 3 non cirrhotic patients (96%) and a<br />

lower SVR 12<br />

rate in cirrhotic (70%). GT3 cirrhotic remain a<br />

difficult to treat population and may benefit from the addition<br />

of ribavirin (RBV) or extended treatment duration. This analysis<br />

reports SVR results from an ongoing multicenter compassionate<br />

use program (ATU) in France. Methods. Patients with GT3 and<br />

advanced liver disease from 221 centers have been included<br />

since March 2014 to October 2014. Transplanted patients<br />

were excluded from this analysis. All patients received DCV+-<br />

SOF QD for 12 or 24 weeks, with RBV added at the physician’s<br />

discretion. We report interim SVR 12<br />

rates in patients who have<br />

completed treatment to date and for whom at least one follow-up<br />

visit was available. Results. 395 patients were enrolled<br />

with the following characteristics: males (75.5%), median<br />

age (54.4 years; 27-83), HCV mono-infected (80%), cirrhosis<br />

(n=322; 81.5%), including Child-Pugh A (85.5%), B (11.5%),<br />

C (3.0%) and treatment-experienced (n=288; 73.1%). Median<br />

HCV-RNA was 5.99 (1.20; 7.41) log IU/mL and median platelets<br />

count was 120.0 x10 9 /L (30-387). 71.6% of patients have<br />

albumin≥ 35 g/L. RBV was given in 20.5% of patients and<br />

treatment duration was 24 weeks in 42.0% and 12 weeks in<br />

17.5%. Data on SVR 12<br />

is available in 78 patients; SVR 12<br />

was<br />

achieved in 90% (70/78). SVR 12<br />

rates according to treatment<br />

schedule and fibrosis stage are presented in the table. The<br />

treatment failures were related to relapse in 6 patients and<br />

unknown reasons in 2 patients (only D0 and SVR 12<br />

available).<br />

No virologic breakthrough was observed. Treatment discontinuations<br />

were related to: 1 adverse event, 4 deaths not related<br />

to study drugs, 1 patient’s decision, 1 treatment failure and 4<br />

for unknown reason. Efficacy and safety data will be updated<br />

and presented in the 395 patients at the congress. CONCLU-<br />

SION: DCV+SOF±RBV regimen demonstrated high SVR 12<br />

rate<br />

in GT3 patients with advanced liver disease. The optimal treatment<br />

duration was 24 weeks in cirrhotic patients. The impact of<br />

RBV will be analyzed in the overall population and presented<br />

at the meeting.<br />

Efficacy of DCV+SOF±RBV regimens in GT3 patients<br />

Disclosures:<br />

Christophe Hezode - Speaking and Teaching: Roche, BMS, MSD, Janssen, abbvie,<br />

Gilead<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

Hélène Fontaine - Board Membership: Abbvie, Gilead, BMS, Janssen; Independent<br />

Contractor: gilead, BMS, MSD, Roche, Janssen<br />

Fabien Zoulim - Advisory Committees or Review Panels: Janssen, Gilead, Novira,<br />

Abbvie, Tekmyra, Transgene; Consulting: Roche; Grant/Research Support:<br />

Novartis, Gilead, Scynexis, Roche, Novira, Assemblypharm; Speaking and<br />

Teaching: Bristol Myers Squibb, Gilead<br />

Pascal Lebray - Grant/Research Support: Merck, astellas, Biotest, BMS; Speaking<br />

and Teaching: Janssen, MSD, Gilead<br />

Nathalie Boyer - Board Membership: MSD, JANSSEN, Gilead, Abbvie; Speaking<br />

and Teaching: BMS<br />

Dominique G. Larrey - Advisory Committees or Review Panels: BAYER, SANOFI,<br />

PFIZER, SERVIER-BIG, AEGERION, MMV, BIAL-QUINTILES, TEVA, ORION, NEG-<br />

MA-LERADS, ASTELLAS, ASTRAZENECA, DNDI, GSK, J AND J; Board Membership:<br />

BMS, GILEAD, ABBVIE, BMS, GILEAD, ITREAS, MMS; Grant/Research<br />

Support: GILEAD, MSD, BMS, ABBVIE, TIBOTEC/JANSSEN, BMS<br />

Danielle Botta-Fridlund - Consulting: GILEAD, ABBVIE, BMS, MSD<br />

Vincent Leroy - Board Membership: Abbvie, BMS, Gilead; Consulting: Janssen,<br />

MSD; Speaking and Teaching: Abbvie, BMS, Gilead, Janssen, MSD<br />

Marc Bourlière - Advisory Committees or Review Panels: Schering-Plough,<br />

Bohringer inghelmein, Schering-Plough, Bohringer inghelmein, Transgene; Board<br />

Membership: Bristol-Myers Squibb, Gilead, Idenix; Consulting: Roche, Novartis,<br />

Tibotec, Abott, glaxo smith kline, Merck, Bristol-Myers Squibb, Novartis, Tibotec,<br />

Abott, glaxo smith kline; Speaking and Teaching: Gilead, Roche, Merck, Bristol-Myers<br />

Squibb<br />

Isabelle Fouchard-Hubert - Speaking and Teaching: JANSSEN, BMS, GILEAD,<br />

ABBVIE<br />

Dominique Guyader - Advisory Committees or Review Panels: ROCHE, GILEAD,<br />

IRIS, ABBVIE; Board Membership: MERCK; Grant/Research Support: JANSSEN;<br />

Speaking and Teaching: BMS<br />

Eric Nguyen-Khac - Speaking and Teaching: Gilead, Abbvie, Janssen, Roche,<br />

MSD, BMS<br />

Vincent Di Martino - Advisory Committees or Review Panels: Gilead, France,<br />

Abbvie, BMS France; Board Membership: MSD France; Consulting: Gilead,<br />

France; Speaking and Teaching: Janssen, BMS France, Gilead France<br />

Fabrice Carrat - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Janssen, Merck, Gilead, BMS, Roche<br />

Larysa Fedchuk - Employment: Bristol-Myers Squibb<br />

Raoudha Akremi - Employment: BMS<br />

Jean-Pierre Bronowicki - Consulting: Merck, Janssen, Boehringer Ingelheim, Gilead,<br />

BMS, Bayer, Novartis, GSK, Merck, Janssen, Boehringer Ingelheim, Gilead,<br />

BMS, Bayer, Novartis, GSK, ABBVIE, ABBVIE; Speaking and Teaching: Roche,<br />

Merck, Janssen, BMS, Bayer, Roche, Merck, Janssen, BMS, Bayer, ABBVIE<br />

The following authors have nothing to disclose: Christine Silvain, Louis d’Alteroche,<br />

Isabelle Rosa, Yacia Bennai


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 315A<br />

207<br />

Antiviral Therapy for Chronic Hepatitis B (CHB) Reduces<br />

the Incidence of Hepatocellular Carcinoma (HCC)<br />

Regardless of Cirrhosis Status: Analysis with Adjustment<br />

for REACH-B Risk Score<br />

Derek Lin 2 , Hwai-I Yang 3,4 , Nghia H. Nguyen 12,1 , Joseph K.<br />

Hoang 1 , Yoona Kim 1 , Vinh D. Vu 1 , An K. Le 1 , Kevin T. Chaung 1,5 ,<br />

Vincent G. Nguyen 1,5 , Huy N. Trinh 6 , Jiayi Li 7 , Jian Q. Zhang 10 ,<br />

Ann W. Hsing 8,9 , Chien-Jen Chen 3,11 , Mindie H. Nguyen 1 ; 1 Division<br />

of Gastroenterology and Hepatology, Stanford University<br />

Medical Center, Palo Alto, CA; 2 Department of Medicine, Stanford<br />

University Medical Center, Palo Alto, CA; 3 Genomics Research<br />

Center, Academia Sinica, Taipei, Taiwan; 4 Institute of Clinical<br />

Medicine, National Yang-Ming University, Taipei, Taiwan; 5 Pacific<br />

Health Foundation, San Jose, CA; 6 San Jose Gastroenterology,<br />

San Jose, CA; 7 Palo Alto Medical Foundation, Mountain View,<br />

CA; 8 Cancer Prevention Institute of California, Fremont, CA; 9 Stanford<br />

School of Medicine, Stanford Cancer Institute, Palo Alto, CA;<br />

10 Chinese Hospital, San Francisco, CA; 11 Graduate Institute of Epidemiology<br />

and Preventive Medicine, National Taiwan University,<br />

Taipei, Taiwan; 12 Department of Medicine, University of California<br />

San Diego, San Diego, CA<br />

Objective: Benefits of antiviral therapy on HCC incidence<br />

remains controversial. We aim to examine the effect of antiviral<br />

therapy for CHB on HCC incidence after adjustment for<br />

background risks using a previously validated REACH-B risk<br />

score. Methods: We examined a cohort of 5,908 CHB patients<br />

which included primary analysis of 2,255 from a United<br />

States cohort (973 of these received antiviral therapy) and<br />

3,653 patients from the previously described community-based<br />

Taiwanese REVEAL-HBV study (none received antiviral therapy).<br />

We used Cox proportional hazards models to calculate<br />

the hazard ratios (HRs) of developing HCC with adjustment<br />

using the previously validated REACH-B risk score (17-point<br />

composite score of sex, age, ALT, HBeAg status, and HBV<br />

DNA). Results: There were a total of 273 incident cases: 12<br />

of 973 treated subjects and 263 of 4935 untreated subjects.<br />

There was a significant 77% risk reduction in HCC incidence<br />

in patients with antiviral treatment compared with those without<br />

treatment (HR 0.23; 95% CI 0.13-0.43; pULN). Conclusion: In this large and diverse<br />

cohort of clinic and community participants from two countries,<br />

antiviral therapy was significantly associated with a reduction<br />

in HCC risk after adjustment for background risk using a validated<br />

risk score inclusive of 5 major known predictors of HCC.<br />

To our knowledge, this is the largest assembled CHB cohort to<br />

examine the effect of antiviral therapy.<br />

REACH-B Adjusted HCC Incidence by Treatment<br />

Disclosures:<br />

Yoona Kim - Employment: Proteus Digital Health<br />

Huy N. Trinh - Advisory Committees or Review Panels: BMS, Gilead; Grant/<br />

Research Support: BMS, Gilead; Speaking and Teaching: BMS, Gilead; Stock<br />

Shareholder: Gilead<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Derek Lin, Hwai-I Yang, Nghia H.<br />

Nguyen, Joseph K. Hoang, Vinh D. Vu, An K. Le, Kevin T. Chaung, Vincent G.<br />

Nguyen, Jiayi Li, Jian Q. Zhang, Ann W. Hsing, Chien-Jen Chen<br />

208<br />

A Single Dose of RG-101, a GalNAc-Conjugated Oligonucleotide<br />

Targeting miR-122, Results in Undetectable<br />

HCV RNA Levels in Chronic Hepatitis C Patients at Week<br />

28 of Follow-up<br />

Meike H. van der Ree 1 , J. Marleen D. Vree 2 , Femke Stelma 1 ,<br />

Sophie Willemse 1 , Marc van der Valk 1 , Svend Rietdijk 3,1 , Richard<br />

Molenkamp 4 , Janke Schinkel 4 , Salah Hadi 5 , Marten Harbers 5 ,<br />

Andre van Vliet 5 , Joanna Udo de Haes 5 , Paul Grint 6 , Steven<br />

Neben 6 , Neil Gibson 6 , Hendrik W. Reesink 1 ; 1 Gastroenterology<br />

and Hepatology, AMC, Amsterdam, Netherlands; 2 Gastroenterology<br />

and Hepatology, UMCG, Groningen, Netherlands; 3 Gastroenterology<br />

and Hepatology, OLVG, Amsterdam, Netherlands;<br />

4 Medical Microbiology, Clinical Virology Laboratory, AMC,<br />

Amsterdam, Netherlands; 5 PRA Healthsciences, Zuidlaren, Netherlands;<br />

6 Regulus Therapeutics, San Diego, USA Minor Outlying<br />

Islands<br />

Background: MicroRNA-122 (miR-122) plays a crucial role in<br />

the hepatitis C virus (HCV) life cycle. Binding of miR-122 to<br />

the 5’UTR of the HCV genome promotes HCV RNA stability<br />

and accumulation, protects the HCV genome from degradation<br />

by cellular exonucleases and prevents induction of an innate<br />

immune response against the virus. The aim of this study was<br />

to evaluate the safety and efficacy of a single dose of RG-101,<br />

a carbohydrate conjugated oligonucleotide that targets miR-<br />

122 in hepatocytes, in chronic HCV patients infected with<br />

various HCV genotypes. Methods: In this multicenter phase<br />

1 study, we included 32 chronic HCV patients with genotype<br />

1, 3 or 4 infection. The first cohort received a single subcutaneous<br />

injection of 2 mg/kg RG-101 (n=14) or placebo (n=2),<br />

the second cohort received a single subcutaneous injection<br />

of 4 mg/kg (n=14) or placebo (n=2). Both cohorts were followed<br />

8 weeks after randomization, and patients with >2 log


316A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

decrease in HCV RNA level from baseline were included in<br />

an extended follow-up study in which patients were followed<br />

until 28 weeks after dosing. Patients were eliminated from the<br />

extended follow-up study if they had > 1 log increase in HCV<br />

RNA level from nadir. HCV RNA levels were measured using<br />

Roche COBAS AmpliPrep/COBAS Taqman HCV v2.0 assay,<br />

with a reported LLOQ of 15 IU/mL. Results: Thirty-two patients<br />

infected with HCV genotype 1 (n=16), genotype 3 (n=10) or<br />

genotype 4 (n=6) were included in the study. Twenty-three<br />

patients were treatment naïve and 9 patients were virological<br />

relapsers to prior interferon-based therapy. None of the<br />

patients had cirrhosis. At baseline, mean HCV RNA levels were<br />

comparable between RG-101 dosed patients versus placebo<br />

(6.2 versus 6.4 log 10 IU/mL, p=0.53). The mean viral load<br />

reduction at week 4 was 4.4 log 10<br />

IU/mL (range 2.3-5.8) in<br />

patients dosed with RG-101 (p200,000 IU/mL were randomized<br />

1:1 to receive either TDF from gestation week 30–32 to<br />

postpartum week 4 or no treatment, and were followed-up until<br />

postpartum week 28. All infants received immunoprophylaxis.<br />

The primary measurement was the MTCT rate, while endpoints<br />

included TDF safety, maternal HBV DNA reduction at delivery,<br />

and HBeAg or hepatitis B s antigen loss/seroconversion<br />

at postpartum week 28. Results: Among the 200 mothers<br />

enrolled in 5 regions of the country, 180 completed the study.<br />

At postpartum week 28, the MTCT rate was significantly lower<br />

in infants from TDF-treated mothers when compared to those<br />

from non-treated mothers, both on per-protocol analysis (0%<br />

vs. 6.82%, P = 0.013) and intention-to-treat analysis (5.16%<br />

vs. 18.0%, P = 0.007). The safety profile was similar between<br />

groups, with no difference in birth defect rates (2.11% with TDF<br />

exposure vs. 1.14% without exposure, P = 1.00). HBV DNA<br />

levels decreased to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 317A<br />

Baseline Values<br />

† When compared respective variables between groups, P ><br />

0.05 except HBeAg+ at birth (P < 0.001) and detectable HBV<br />

DNA at birth (P = 0.002). *3 mothers withdrew before the baseline<br />

visit.<br />

Disclosures:<br />

Calvin Q. Pan - Advisory Committees or Review Panels: Gilead; Consulting:<br />

BMS, Gilead, Abbvie, Janssen ; Grant/Research Support: BMS, Gilead, Merck;<br />

Speaking and Teaching: BMS, Gilead, Abbvie<br />

The following authors have nothing to disclose: Zhong-Ping Duan, Erhei Dai,<br />

Shuqin Zhang, Guo Rong Han, Yuming Wang, Huaihong Zhang, Huaibin Zou,<br />

Bao Shen Zhu, Wen Jing Zhao, Hong Xiu Jiang<br />

210<br />

High Efficacy of Grazoprevir/Elbasvir (GZR/EBR) in HCV<br />

Genotype 1, 4, and 6-infected Patients With HIV Coinfection:<br />

SVR24 Data From the Phase 3 C-EDGE Coinfection<br />

Study<br />

Jürgen K. Rockstroh 1 , Mark Nelson 2 , Christine Katlama 3 , Jacob P.<br />

Lalezari 4 , Josep Mallolas 5 , Mark Bloch 6 , Gail Matthews 7 , Michael<br />

Saag 8 , Philippe J. Zamor 9 , Chloe Orkin 10 , Jacqueline Gress 11 ,<br />

Melissa Shaughnessy 11 , Stephanie Klopfer 11 , Anita Y. Howe 11 ,<br />

Janice Wahl 11 , Bach-Yen T. Nguyen 11 , Eliav Barr 11 , Heather L.<br />

Platt 11 , Michael Robertson 11 , Mark S. Sulkowski 12 ; 1 Bonn Umiversity,<br />

Bonn, Germany; 2 Chelsea and Westminster Hospital, London,<br />

United Kingdom; 3 Université Pierre et Marie Curie, Paris VI and<br />

Hôpital Pitié-Salpêtrière, Paris, France; 4 Quest Clinical Research,<br />

San Francisco, CA; 5 Hospital Clinic-University of Barcelona, Barcelona,<br />

Spain; 6 Holdsworth House Medical Practice, Darlinghurst,<br />

NSW, Australia; 7 St Vincent’s Hospital, Sydney, NSW, Australia;<br />

8 University of Alabama at Birmingham, Birmingham, AL; 9 Carolinas<br />

Medical Center, Charlotte, NC; 10 Royal London Hospital,<br />

London, United Kingdom; 11 Merck & Co., Inc., Kenilworth, NJ;<br />

12 Johns Hopkins University School of Medicine, Baltimore, MD<br />

BACKGROUND: The fixed-dose combination (FDC) of grazoprevir<br />

(GZR, MK-5172; NS3/4 protease inhibitor) and elbasvir<br />

(EBR, MK-8742; NS5A inhibitor), an interferon-free, ribavirin-free,<br />

once-daily tablet has shown robust efficacy and safety<br />

in diverse populations. C-EDGE Coinfection evaluated GZR/<br />

EBR among treatment-naive, HIV/HCV coinfected patients with<br />

GT1, 4 or 6. METHODS: All patients were on stable antiretroviral<br />

(ARV) therapy (tenofovir or abacavir, and lamivudine or<br />

emtricitabine; and either raltegravir, dolutegravir or rilpivirine)<br />

with CD4 >200 cells/mm 3 and an HIV RNA 500 cells/mm 3 and<br />

HIV RNA 5× resistance to EBR in vitro (L31M, Y93S). Adverse<br />

events (AEs) were reported in 72% (157/218) of patients; serious<br />

AEs occurred in 0.9% (2/218) of patients. Adherence was<br />

>90% in the total population, including virologic failures. PK<br />

parameters were similar in patients with, and without SVR24.<br />

2 patients had transient HIV viremia; both were subsequently<br />

undetectable without a change in ARV regimen. CONCLUSION:<br />

A 12-week regimen of GZR/EBR FDC was highly effective with<br />

a favorable safety profile among HIV/HCV coinfected patients<br />

with GT1, 4 or 6 infection. SVR24 was high in all patient subgroups,<br />

including African-Americans and those with cirrhosis.<br />

Disclosures:<br />

Jürgen K. Rockstroh - Advisory Committees or Review Panels: Abbvie, BI, BMS,<br />

Merck, Roche, Tibotec, Abbvie, Bionor, Tobira, ViiV, Gilead, Janssen; Consulting:<br />

Novartis; Grant/Research Support: Merck; Speaking and Teaching: Abbott,<br />

BI, BMS, Merck, Roche, Tibotec, Gilead, Janssen, ViiV<br />

Mark Nelson - Advisory Committees or Review Panels: Janssen, MSD, BMS,<br />

ABBVIE, Viiv, Gilead; Consulting: Janssen, MSD, BMS, ABBVIE, Viiv, Gilead;<br />

Grant/Research Support: Boehringer Ingelheim, Janssen, MSD, BMS, ABBVIE,<br />

Viiv, Gilead, Roche; Speaking and Teaching: GSK, Janssen, MSD, BMS, Abbott,<br />

Viiv, Gilead<br />

Josep Mallolas - Board Membership: Boehringer, Merck, Abbvie, Gilead, Janssen<br />

Michael Saag - Grant/Research Support: BMS, Gilead, Abbvie, Merck, ViiV,<br />

Janssen<br />

Philippe J. Zamor - Grant/Research Support: AbbVie, Bristol Myers Squibb,<br />

Gilead, Merck & Co. ; Speaking and Teaching: AbbVie, Bristol Myers Squibb,<br />

Janssen<br />

Chloe Orkin - Advisory Committees or Review Panels: Viiv, gilead, bms, boehringer<br />

ingelgeim; Grant/Research Support: johnson and johnson, abbvie<br />

Jacqueline Gress - Employment: Merck, Merck<br />

Melissa Shaughnessy - Employment: Merck & Co., Inc.; Stock Shareholder:<br />

Merck & Co., Inc.<br />

Anita Y. Howe - Employment: Merck Research Laboratory<br />

Janice Wahl - Employment: Merck & Co,<br />

Bach-Yen T. Nguyen - Employment: Merck<br />

Eliav Barr - Employment: Merck<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

The following authors have nothing to disclose: Christine Katlama, Jacob P.<br />

Lalezari, Mark Bloch, Gail Matthews, Stephanie Klopfer, Heather L. Platt


318A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

211<br />

Clinical Features and Outcomes of Complementary and<br />

Alternative Therapy Medicine (CAM)-Induced Acute Liver<br />

Failure and Injury<br />

Luke Hillman 1 , Daniel Ganger 1 , Michelle Gottfried 2 , Jorge Rakela 3 ,<br />

Michael L. Schilsky 4 , William M. Lee 5 ; 1 Medicine, Northwestern<br />

University, Chicago, IL; 2 MUSC, Charleston, SC; 3 Medicine, Mayo<br />

Clinic, Scottdsdale, AZ; 4 Medicine and Surgery, Yale University,<br />

New Haven, CT; 5 Internal Medicine, University of Texas Southwestern,<br />

Dallas, TX<br />

Background: Increased use of CAM across the United States<br />

is evidenced by increased expenditures and reported use by<br />

consumers. While idiosyncratic drug-induced liver injury (DILI)<br />

from conventional prescription medicines is often severe, DILI<br />

from CAM is similarly infrequent but causes significant hepatotoxicity.<br />

AIM: We examined clinical features and outcomes<br />

among patients with acute liver failure (ALF) and acute liver<br />

injury (ALI) enrolled in the Acute Liver Failure Study Group<br />

(ALFSG) database, comparing conventional DILI injury to that<br />

due to CAMs. Patients and Methods: ALFSG has prospectively<br />

enrolled 2626 subjects with ALF/ALI since 1998 from U.S.<br />

academic liver transplant centers through January 8, 2015.<br />

We compared clinical presentation, course and outcome<br />

between those with ALF/ALI due to ‘conventional’ DILI versus<br />

CAMs. Results: We found 251 (10.4%) subjects deemed by<br />

DILIN criteria to have DILI-ALF, 210 (83.7%) due to conventional<br />

medicines and 41 (16.3%) from CAM, with a higher %<br />

of DILI cases due to CAM between 2008-2015, as compared<br />

to 1998-2007 (18.1% vs. 12.4%, p = 0.047). There was no<br />

difference in the type of liver injury (hepatocellular, cholestatic,<br />

or mixed injury) between groups as determined by R score<br />

(p=0.26). On day 1 of enrollment, conventional medicine DILI<br />

showed higher serum alkaline phosphatase levels compared<br />

to the CAM group (mean IU/L, 215.1 vs. 156.2, p=0.005);<br />

there were no differences between both cohorts in ALT/AST,<br />

INR, platelets, bilirubin or MELD score or coma grade. Despite<br />

fewer individual comorbidities were noted in the CAM population<br />

(2.03 vs. 1.15, p=0.005), the CAM group had a lower<br />

transplant-free survival at 21 days, though this was not statistically<br />

significant (24.3% vs. 38.5%, p=0.10). Within the CAM<br />

group, those surviving without transplantation had significantly<br />

lower INRs (1.9 vs. 3.8, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 319A<br />

213<br />

Safety and Tolerability of Ornithine Phenylacetate to<br />

Lower Ammonia in Acute Liver Failure: Preliminary<br />

report of the STOP-ALF Trial<br />

William M. Lee 1 , R. Todd Stravitz 2 , Robert J. Fontana 3 , A. James<br />

Hanje 4 , Holly Tillman 5 , Kristen M. Clasen 5 , Valerie L. Durkalski 5 ,<br />

Averell H. Sherker 6 , Edward Doo 6 , Brian C. Davis 7 ; 1 Division of<br />

Digestive and Liver Diseases, University of Texas Southwestern<br />

Medical Center at Dallas, Dallas, TX; 2 Gastroenterology, Virginia<br />

Commonwealth University, Richmond, VA; 3 Gastroenterology, University<br />

of Michigan, Ann Arbor, MI; 4 Gastroenterology, Hepatology<br />

and Nutrition, The Ohio State University Wexner School of<br />

Medicine, Columbus, OH; 5 Medical University of South Carolina,<br />

Charleston, SC; 6 NIDDK, NIH, Bethesda, MD; 7 Internal Medicine,<br />

UT Southwestern Medical Center at Dallas, Dallas, TX<br />

Background: No satisfactory method is currently available to<br />

lower ammonia (NH 3<br />

) to prevent or treat hepatic encephalopathy<br />

(HE) and cerebral edema in hospitalized patients with<br />

acute liver injury (ALI) or acute liver failure (ALF). Ornithine<br />

phenylacetate (OPA, OCR-002) effectively lowers NH 3<br />

in animal<br />

models of ALF by trapping it as phenylacetylglutamine<br />

(PAGN) that is excreted in urine. STOP-ALF is a phase 2a trial<br />

to develop data on safety and tolerability of a 5-day infusion<br />

of OCR-002 in the setting of ALF/ALI and elevated NH 3<br />

levels.<br />

Methods: Patients with ALF or ALI and elevated NH 3<br />

levels,<br />

>60 mmol/L, were enrolled at 8 US sites participating in the<br />

Acute Liver Failure Study Group (ALFSG) in two cohorts: normal/minimal<br />

renal dysfunction (Cr < 1.5, Cohort 1) and acute<br />

kidney injury (AKI, Cr ≥ 1.5 mg/dL, Cohort 2). Dosing of OCR-<br />

002 was open label, ascending dose, starting at 3.3 g/24hr<br />

by IV infusion, to a maximum dose of 10 gm/24 hr after 24<br />

hr of infusion; for the 1st 3 pts, final dose = 3.3 gm/24hX5<br />

days, then 6.7 gm/24hr for 3 and then 10 gm/24h for the<br />

remaining patients in each cohort. Plasma and urine PK and<br />

plasma NH 3<br />

levels were recorded, in addition to safety labs,<br />

EKGs and frequent neurological checks. Results: 30 patients<br />

have been enrolled, with current analysis based on 22 with PK<br />

data available; median age 32 yrs, 14/22 female, 16 were<br />

acetaminophen-related, 17 were considered evaluable, having<br />

received >72 hrs of OCR-002. The drug was well tolerated<br />

with only minor AE’s attributed (nausea, headache). PK results<br />

showed mean phenylacetate (PAA) plasma levels of 29.0 and<br />

62.5 mg/mL in Cohort 1 at the 2 higher dose levels, and 32.5<br />

mg/mL for Cohort 2, dose level 2. These levels of PAA are<br />

below the optimal concentration for NH 3<br />

removal. Overall survival<br />

was 73%, with deaths limited to advanced AKI patients<br />

and those unable to receive liver transplantation; 1 pt. required<br />

transplantation. NH 3<br />

levels improved in most patients: median<br />

baseline NH 3<br />

=87 mmol/L to 54 mmol/L day 4, 0 hour. This<br />

ongoing trial is currently completing the final dose level; no<br />

toxic levels of PAA or PAGN have been reached. Conclusion:<br />

OCR-002 was well tolerated without any safety signals and<br />

shows promise as adjunctive therapy in ALF. Higher doses of<br />

OCR-002 may be needed for optimal ammonia reduction in<br />

ALF. Future <strong>studies</strong> to assess safety and efficacy of higher doses<br />

of OCR-002 will be pursued.<br />

Disclosures:<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

Robert J. Fontana - Consulting: GlaxoSmithKline, CLDF; Grant/Research Support:<br />

Gilead, vertex, BMS, Jansen, Gilead<br />

A. James Hanje - Consulting: Salix pharmaceutical<br />

The following authors have nothing to disclose: R. Todd Stravitz, Holly Tillman,<br />

Kristen M. Clasen, Valerie L. Durkalski, Averell H. Sherker, Edward Doo, Brian<br />

C. Davis<br />

214<br />

Characterization of Cerebral Edema in Acute on Chronic<br />

Liver Failure<br />

Radha K. Dhiman 1 , Tarana Gupta 1 , Chirag K. Ahuja 2 , Swastik<br />

Agrawal 1 , Naveen Kalra 2 , Ajay K. Duseja 1 , Niranjan Khandelwal<br />

2 , Yogesh K. Chawla 1 ; 1 Hepatology, Postgraduate Institute of<br />

Medical Education & Research, Chandigarh, India; 2 Radiodiagnosis<br />

& Imaging, Postgraduate Institute of Medical Education &<br />

Research, Chandigarh, India<br />

Background and Aims: The significance and nature of cerebral<br />

edema in acute-on-chronic liver failure (ACLF) is not well<br />

studied. This study aimed to characterize cerebral edema in<br />

ACLF using magnetization transfer ratio (MTR) and diffusion<br />

tensor imaging (DTI). Methods: Forty consecutive patients with<br />

cirrhosis and acute decompensation were included. Patients<br />

were categorized as no-ACLF (n=11), grade 1 (n=10), grade<br />

2 (n=9) and grade 3 (n=10) ACLF as per CANONIC study criteria<br />

(Moreau et al. Gastroenterology 2013;144:1426-1437).<br />

Whole brain conventional magnetic resonance imaging (MRI)<br />

was performed on a 3.0 Tesla scanner using standard 16/32<br />

channel head coils. Apart from the conventional sequences<br />

namely T2 weighted, FLAIR weighted and T1 weighted volumetric<br />

sequences (MPRAGE/SPGR BRAVO), additional sequences<br />

performed were magnetization transfer ratio MTR and DTI.<br />

MRI of brain was performed and tumor necrosis factor α, interleukin<br />

(IL)-1β, and IL-6 were measured at baseline and after<br />

7-10 days of admission. Patients were followed up for 90-days<br />

from admission. Ten age- and sex-matched healthy subjects<br />

were also included as controls. Results: MTR values were significantly<br />

lower in patients in both no-ACLF and ACLF groups<br />

compared to controls. Mean diffusivity (MD) values progressively<br />

increased from controls to no-ACLF to ACLF grade 1, 2<br />

and 3 groups in the basal ganglia (P 8 x 10 -9 M 2 /s) in frontal white matter were independent<br />

predictors of 90-day mortality (Table). Conclusions: Our<br />

study conclusively demonstrates cerebral edema increases with<br />

severity of ACLF and is predominantly due to extracellular brain<br />

water increase. Correlation between MD values and IL-6 levels<br />

suggests pathogenic role of inflammation. MELD score and<br />

cerebral edema have prognostic significance in ACLF patients.<br />

Utility of MELD and mean diffusivity values as predictors of<br />

90-day mortality<br />

MELD, model for end-stage liver disease; MD FWM, mean diffusivity<br />

frontal white matter; PPV, positive predictive value; NPV,<br />

negative predictive value; DA, diagnostic value, AUROC, area<br />

under ROC curve<br />

Disclosures:<br />

The following authors have nothing to disclose: Radha K. Dhiman, Tarana Gupta,<br />

Chirag K. Ahuja, Swastik Agrawal, Naveen Kalra, Ajay K. Duseja, Niranjan<br />

Khandelwal, Yogesh K. Chawla


320A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

215<br />

The transportable machine perfusion Airdrive®, a novel<br />

approach to safely expand the donor pool for liver<br />

transplantation<br />

Philippe Compagnon 1,2 , Ismail Ben Mosbah 2 , Hassen Hentati 1 ,<br />

Eric Levesque 3 , Mara Disabato 1 , Jose L. Cohen 2 , Daniel Azoulay 1 ;<br />

1 Service de Chirurgie Digestive et Hépatobiliaire -Transplantation<br />

Hépatique, Hopital Henri Mondor - Assistance Publique-Hopitaux<br />

de Paris, Université Paris-Est, Creteil, France; 2 IMRB- Inserm U955,<br />

Hopital Henri Mondor, Assistance Publique-Hopitaux de Paris, Université<br />

Paris-Est créteil, Créteil, France; 3 Service de Réanimation<br />

Digestive, Hopital Henri Mondor, Assitance Publique-Hopitaux de<br />

Paris, Université Paris-Est, Créteil, France<br />

Background and Aims: Livers derived from donation after<br />

circulatory death (DCD) suffer warm ischemia (WI) and are<br />

infrequently used for transplantation; they have the potential,<br />

however, to considerably expand the donor pool. The<br />

Airdrive® is the first transportable oxygenated machine perfusion<br />

(MP) unit suitable for liver preservation. We aimed<br />

to determine whether the Airdrive® MP would improve the<br />

quality of these livers using a DCD large animal model. Methods:<br />

Female large white pigs were used as organ donors and<br />

recipients. Liver allografts procured from heart beating donors<br />

and preserved by simple cold storage (SCS) served as controls<br />

(n=6). In experimental groups, cardiac arrest was induced by<br />

IV injection of KCL. After 60min WI, livers were perfused in<br />

situ with HTK and subsequently preserved either by SCS (SCS<br />

group, n=3) or hypothermic MP (Airdrive® group, n=4) using<br />

MPS-Belzer solution. After 4hr of preservation, all livers were<br />

transplanted into recipient pigs. The main judgment criterion<br />

was animal survival at day 5 Results: All animals that received<br />

a simple cold stored liver allograft after 60min WI experienced<br />

PNF and died within 6 hours after transplantation (5-day survival<br />

= 0%). In contrast, 5-day survival in animals that received<br />

a MP liver was 100% (4/4, Airdrive® group) as well as in<br />

controls (6/6). A post-reperfusion syndrome was observed in<br />

all animals (3/3) of the SCS group but none of the control<br />

or Airdrive® groups. These phenomena caused a significant<br />

increase in fluid challenge and catecholamine needs in SCS<br />

group relative to control and Airdrive® groups. At the end of<br />

cold preservation, ATP content was significantly higher in the<br />

Airdrive® group vs. SCS group. After reperfusion, MP livers<br />

functioned better (albumin production, prothrombin time rates)<br />

and showed less hepatocellular and endothelial cell injury, in<br />

agreement with better preserved liver integrity (histology) vs.<br />

SCS group. MP livers also exhibited higher ATP recovery than<br />

SCS livers. The protective effect of the Airdrive® device was<br />

associated with a significant attenuation of oxidative stress<br />

(lower lipid peroxidation, higher catalase and superoxide<br />

dismutase activity), and a better endoplasmic reticulum adaptation<br />

leading to limit mitochondrial damage (cytochrome c,<br />

caspase 9, GLDH), and apoptosis (caspase 3 and TUNEL) Conclusions:<br />

This study demonstrates for the first time the efficacy of<br />

the transportable MP Airdrive® device to enhance donor liver<br />

viability for transplantation in a clinically relevant DCD model.<br />

Disclosures:<br />

The following authors have nothing to disclose: Philippe Compagnon, Ismail Ben<br />

Mosbah, Hassen Hentati, Eric Levesque, Mara Disabato, Jose L. Cohen, Daniel<br />

Azoulay<br />

216<br />

A Novel Percutaneous Organ Perfusion Stent Improves<br />

Liver Perfusion in a Porcine Model of Donation after<br />

Cardiac Death<br />

Bryan Tillman 1 , Youngjae Chun 2 , Nathan L. Liang 1 , Tara D. Richards<br />

1 , Anthony J. Demetris 4 , Timothy M. Maul 2 , Amit D. Tevar 3 ;<br />

1 Division of Vascular Surgery, University of Pittsburgh Medical<br />

Center, Pittsburgh, PA; 2 Swanson School of Engineering, University<br />

of Pittsburgh, Pittsburgh, PA; 3 Starzl Transplant Institute, University<br />

of Pittsburgh Medical Center, Pittsburgh, PA; 4 Department of<br />

Pathology, University of Pittsburgh Medical Center, Pittsburgh, PA<br />

Introduction: Donation after cardiac death (DCD) liver allografts<br />

are frequently discarded due to ischemia during the agonal<br />

period prior to cardiac death. A dual-chambered Organ Perfusion<br />

Stent (OPS) would isolate blood flow to the abdominal<br />

organs to maintain perfusion, while agonal cardiac flow would<br />

pass through a large central lumen. As a result, this approach<br />

would isolate perfusion to the liver from the unstable circulation<br />

of the agonal donor without affecting donor cardiac<br />

death. Methods: OPS were constructed with a central lumen<br />

for cardiac flow and an external visceral perfusion chamber. A<br />

porcine model included three control and six stent animals. The<br />

OPS was deployed through endovascular sheaths in the aorta<br />

to include the visceral branches while under cardiac monitoring.<br />

Venous blood from a separate cannula was cycled through<br />

an oxygenator and back to the isolated visceral arteries. Next,<br />

the DCD agonal period was simulated for 60 minutes with a<br />

target MAP of 40 mm Hg. Results: Angiography of the outer<br />

chamber confirmed perfusion of the visceral organs, while<br />

agonal cardiac blood flow passed through the central lumen.<br />

During stent placement, no change was observed in cardiac<br />

parameters of HR, RVEDV, CVP, MAP, or SvO2 (P > 0.05)<br />

(n=6). Cardiac output decreased by an average 11% consistent<br />

with offloading of visceral perfusion (P=0.01). Stented<br />

animals revealed an average of 4.8 fold improved oxygen<br />

delivery compared to controls. Following the agonal period,<br />

the livers of the controls revealed marked ischemic appearance<br />

contrasting to the adequately perfused livers of the stent treated<br />

animals. Conclusions: This novel OPS achieves improved oxygen<br />

delivery to visceral organs without significantly impacting<br />

cardiac physiology of the donor. A single deployment step<br />

allows for collective isolation of all four visceral branches. In<br />

summary, this novel approach may significantly increase the<br />

quality of livers available for transplantation.<br />

Disclosures:<br />

Timothy M. Maul - Advisory Committees or Review Panels: Mallinckrodt<br />

The following authors have nothing to disclose: Bryan Tillman, Youngjae Chun,<br />

Nathan L. Liang, Tara D. Richards, Anthony J. Demetris, Amit D. Tevar


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 321A<br />

217<br />

Daclatasvir and Sofosbuvir in Patients with Recurrent<br />

HCV Following Liver Transplantation and Advanced<br />

Fibrosis or Cirrhosis: United States Multicenter Treatment<br />

Protocol<br />

Paul Y. Kwo 1 , Michael W. Fried 2 , K. Rajender Reddy 3 , Consuelo<br />

Soldevila-Pico 4 , Saro Khemichian 5 , Jama M. Darling 2 , Andrew<br />

A. Napoli 6 , Beatrice Anduze-Faris 6 , Robert S. Brown 7 ; 1 Indiana<br />

University, Indianapolis, IN; 2 University of North Carolina, Chapel<br />

Hill, NC; 3 Department of Medicine, University of Pennsyslvania,<br />

Philadelphia, PA; 4 Department of Medicine, University of Florida,<br />

Gainesville, GA; 5 Keck School of Medicine, University of Southern<br />

California, Los Angeles, CA; 6 Bristol-Meyers Squibb, Plainsboro,<br />

NJ; 7 Department of Medicine, Columbia University College of Physicians<br />

& Surgeons, New York, NY<br />

Background: In the United States, there are limited treatment<br />

options for liver transplant recipients with recurrent HCV and<br />

advanced fibrosis/cirrhosis. In a Phase 3 study, the pangenotypic<br />

combination of daclatasvir (DCV) and sofosbuvir (SOF)<br />

with ribavirin demonstrated SVR rates of 94% in liver transplant<br />

recipients with HCV recurrence. This analysis reports interim<br />

findings from a U.S. Treatment Use protocol in patients with<br />

recurrent HCV following liver transplantation and advanced<br />

fibrosis or cirrhosis. Methods: In this multicenter, open-label,<br />

expanded access study, DCV was provided for use in combination<br />

with SOF, with or without RBV at the investigator’s<br />

discretion, for up to 24 weeks in genotype 1-6 patients with<br />

F3 fibrosis, cirrhosis, or fibrosing cholestatic hepatitis (FCH)<br />

following liver transplantation. The HCV TARGET consortium of<br />

academic and community medical centers and data collection<br />

methodology were utilized. This interim analysis includes 51<br />

patients who have enrolled and initiated treatment at 20 sites.<br />

Results: Of the 51 patients treated, 49 received DCV+SOF<br />

and 2 received DCV+SOF+RBV. Patients were predominantly<br />

male (77%), white (82%) and infected with HCV genotype 1<br />

(73%) or genotype 3 (24%). Fifty-seven percent had failed<br />

prior treatment with IFN-based regimens (12% included a<br />

protease inhibitor). Cirrhosis was present in 82% of patients,<br />

including 57% with evidence of post transplant hepatic decompensation,<br />

8% had a kidney transplant, and 8% had FCH. At<br />

baseline, median HCV RNA was 2.9 x 10 6 IU/ml, the mean<br />

(range) for platelets was 129 (41-239) x10 3 cells/uL, albumin<br />

3.4 (1.7-6.3) g/dL, total bilirubin 1.5 (0.3-13) mg/dL, INR 1.1<br />

(1.0-1.5), and creatinine clearance was 83 (40-190) ml/min.<br />

MELD was ≥10 in 48% of patients with available scores. Of<br />

21 patients with available data, the end of treatment virologic<br />

response was 100%. All four patients who reached post-treatment<br />

week 12 achieved SVR. Adverse events occurring in<br />

≥10% of patients were fatigue, headache, nausea, and arthralgia.<br />

Serious adverse events of interest included transplant rejection<br />

(n=1) and renal failure (n=4). Two patients discontinued<br />

treatment for adverse events; 1 due to renal failure leading to<br />

death (unrelated to treatment) and 1 due to death of unknown<br />

cause. SVR results for 31 patients will be available for presentation.<br />

Conclusion: DCV+SOF for 24 weeks was well-tolerated<br />

in this population with severe recurrent post-transplant hepatitis<br />

C. Interim data indicate effective end of treatment antiviral<br />

response and SVR in the patients receiving this regimen<br />

through compassionate use.<br />

Disclosures:<br />

Paul Y. Kwo - Advisory Committees or Review Panels: Abbvie, Abbott, Novartis,<br />

Merck, Gilead, BMS, Janssen; Consulting: BMS; Grant/Research Support:<br />

Roche, Abbvie, Merck, BMS, Abbott, Idenix, Vital Therapeutics, Gilead, Vertex,<br />

Merck, Idenix; Speaking and Teaching: Merck, Merck<br />

Michael W. Fried - Consulting: Merck, Abbvie, Janssen, Bristol Myers Squibb,<br />

Gilead; Grant/Research Support: Merck, AbbVie, Janssen, Bristol Myers Squibb,<br />

Gilead; Patent Held/Filed: HCCPlex<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

Saro Khemichian - Advisory Committees or Review Panels: Gilead, AbbVie, BMS;<br />

Speaking and Teaching: Gilead, AbbVie, Salix<br />

Jama M. Darling - Consulting: BristolMyers Squibb; Grant/Research Support:<br />

BristolMyers Squibb<br />

Andrew A. Napoli - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb<br />

Beatrice Anduze-Faris - Employment: Bristol-Myers Squibb<br />

Robert S. Brown - Consulting: Gilead, Janssen, Abbvie, Merck, BMS<br />

The following authors have nothing to disclose: Consuelo Soldevila-Pico<br />

218<br />

Rapid Decline In Interferon Stimulated Genes In Prior<br />

HCV Non-Responders Re-Treated With Direct Acting<br />

Antivirals<br />

Hawwa Alao 1 , Yun Ju Kim 1 , Elizabeth C. Wright 4 , Margaret<br />

Cam 2 , Elenita M. Rivera 1 , Nancy Fryzek 1 , David E. Kleiner 3 , Fang<br />

Zhang 1 , Edward Doo 1 , Yaron Rotman 1 , Christopher Koh 1 , Theo<br />

Heller 1 , Jay H. Hoofnagle 1 , T. Jake Liang 1 , Marc G. Ghany 1 ;<br />

1 Liver Diseases Branch, National Institutes of Health, Bethesda,<br />

MD; 2 NCI, NIH, Bethesda, MD; 3 LP, NCI, NIH, Bethesda, MD;<br />

4 Office of the Director, NIDDK, NIH, Bethesda, MD<br />

BACKGROUND: The development of direct acting antiviral<br />

(DAA) agents has revolutionized therapy of chronic hepatitis<br />

C. Prior non-responders are known to have high baseline<br />

hepatic expression of interferon stimulated genes (ISGs). The<br />

effect of potent DAAs on ISG expression is unknown. AIM: To<br />

investigate the effect of an oral DAA regimen on ISG expression<br />

in prior failures to interferon-based regimens with chronic<br />

hepatitis C (CHC). METHODS: Subjects with HCV genotype<br />

1b infection who had failed to respond to peginterferon and<br />

ribavirin were re-treated with asunaprevir anddaclatasvir for<br />

24 weeks. Subjects were monitored after therapy for 12 weeks<br />

to determine SVR12. Patients underwent paired liver biopsies,<br />

the first within 12 weeks before therapy and the second at<br />

either week 2 or 4 on therapy. Microarray analysis (Affymetrix<br />

Human Gene 2.0 ST array, Santa Clara, CA) was performed<br />

on liver biopsy tissue and on-therapy gene expression was<br />

compared to baseline. All subjects provided informed consent<br />

to participate. RESULTS: 11 patients with chronic HCV genotype<br />

1b infection had paired biopsies available for microarray<br />

analysis. The 11 patients included 6 women, mean age<br />

62 years, mean HCV RNA 6.8 log 10<br />

IU/mL, 4 with cirrhosis.<br />

The SVR rate was 64%. All 4 failures experienced virological<br />

breakthrough between weeks 4-12. At the week 2 biopsy, 5 of<br />

5 subjects had detectable virus (


322A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Hawwa Alao, Yun Ju Kim, Elizabeth<br />

C. Wright, Margaret Cam, Elenita M. Rivera, Nancy Fryzek, David E.<br />

Kleiner, Fang Zhang, Edward Doo, Yaron Rotman, Christopher Koh, Theo Heller,<br />

Jay H. Hoofnagle, T. Jake Liang, Marc G. Ghany<br />

219<br />

The Emergence of NS5B Resistant Associated Variant<br />

S282T after Sofosbuvir-Based Treatment<br />

Edward J. Gane 2 , Armando Abergel 3 , Sophie Metivier 4 , Ronald<br />

Nahass 5 , Michael Ryan 6 , Catherine A. Stedman 7 , Evguenia S.<br />

Svarovskaia 1 , Hongmei Mo 1 , Brian Doehle 1 , Hadas Dvory-Sobol 1 ,<br />

Charlotte Hedskog 1 , Ming Lin 1 , Diana M. Brainard 1 , Phillip S.<br />

Pang 1 , Jenny C. Yang 1 , John G. McHutchison 1 , Mark S. Sulkowski<br />

8 , Ziad Younes 9 , Eric Lawitz 10 ; 1 Gilead Sciences, Foster City,<br />

CA; 2 Auckland Clinical Studies, Auckland, New Zealand; 3 Centre<br />

Hospitalier Universitaire-Estaing, Clermont-Ferrand, France; 4 Centre<br />

Hospitalier Universitaire-Purpan, Toulouse, France; 5 ID Care,<br />

Inc, Hillsborough, NJ; 6 Digestive and Liver Disease Specialists,<br />

Norfolk, VA; 7 Christchurch Clinical Studies Trust, Christchurch,<br />

New Zealand; 8 Johns Hopkins University School of Medicine, Baltimore,<br />

MD; 9 Gastro One, Germantown, TN; 10 University of Texas<br />

Health Science Center, Texas Liver Institute, San Antonio, TX<br />

Introduction: Sofosbuvir (SOF), a nucleotide prodrug, and ledipasvir/sofosbuvir<br />

(LDV/SOF) are approved for the treatment of<br />

chronic HCV. In vitro HCV replicon results demonstrated that<br />

the NS5B substitution, S282T, confers a 2- to 18-fold reduced<br />

susceptibility to SOF in all genotypes while also reducing the<br />

viral replication capacity by 89% to 99% when compared to<br />

wild-type. The aim of this analysis is to evaluate the emergence<br />

and fitness of the S282T variant in virologic failure patients<br />

administered SOF-based regimens across the SOF and LDV/<br />

SOF Phase 2 and Phase 3 programs. Methods: Plasma samples<br />

collected prior to SOF treatment and at the time of virologic<br />

failure were evaluated. The HCV NS5B region was amplified<br />

and deep sequenced (cut-off at 1%). Whole HCV genome<br />

(core to NS5B) sequencing and replication capacity assays<br />

were performed for a subset of patients with S282T to investigate<br />

presence of compensatory substitutions. Results: To date,<br />

over 12,000 patients have been treated in SOF or LDV/SOF<br />

Phase 2 and 3 <strong>studies</strong>. Of these, deep sequencing was available<br />

at baseline in 6,086 subjects (70% GT1, 8% GT2, 18%<br />

GT3, and 4% GT4-6) and at virologic failure in 901 subjects.<br />

SOF regimens administered included monotherapy, SOF+RB-<br />

V±Peg-IFN, or LDV/SOF±RBV for up to 24 weeks. At baseline,<br />

no (0/6086) S282T variant was detected. At virologic failure,<br />

10 of 901 (1%) subjects had S282T detected and 6 of these<br />

subjects had additional NS5B treatment-emergent variants<br />

(L159F, E237G, M289L/I, L320V/I, and V321I) detected. In<br />

the 4 subjects with S282T for whom the entire HCV genome<br />

was successfully sequenced, no consistent amino acid substitutions<br />

were observed outside NS5B. Replication capacity testing<br />

in vitro also did not identify any restoration of the replicative<br />

defect of S282T. Subjects with S282T received SOF monotherapy<br />

(n=2), retreatment with LDV/SOF in prior LDV/SOF<br />

failures (n=3), and LDV/SOF for 8 weeks in genotype 1 (n=1)<br />

and 12 weeks in genotypes 3, 4, 5, and 6 (n=4). S282T levels<br />

declined on average by 4-fold within 2 weeks of follow-up<br />

period confirming low replication fitness of this variant. To<br />

date, 4 of these patients have been retreated: with SOF+RBV in<br />

GT2 infected patient (n=1) or with LDV/SOF+RBV in GT1 (n=2)<br />

and GT3 (n=1) infected patients; all achieved SVR12. Conclusions:<br />

The emergence of S282T variant was rare in patients<br />

who fail SOF-based regimens. S282T variants are unfit in vivo<br />

and rapidly decline over time. Successful retreatment of prior<br />

SOF-failure patients is possible in the presence of S282T variants<br />

with SOF+RBV±LDV regimens.<br />

Disclosures:<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Armando Abergel - Consulting: gilead, msd, bms; Speaking and Teaching: abbvie<br />

Sophie Metivier - Speaking and Teaching: Roche, BMS, Janssen, Merck, Gilead,<br />

abbvie<br />

Ronald Nahass - Advisory Committees or Review Panels: Gilead, MErck, Janssen,<br />

BMS; Grant/Research Support: Gilead, Merck, Janssen, BMS; Speaking and<br />

Teaching: Gilead, Merck, Janssen<br />

Catherine A. Stedman - Advisory Committees or Review Panels: MSD, Abbvie;<br />

Speaking and Teaching: Gilead, Abbvie<br />

Evguenia S. Svarovskaia - Employment: Gilead Sciences Inc.; Stock Shareholder:<br />

Gilead Sciences Inc.<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Brian Doehle - Employment: Gilead Sciences<br />

Hadas Dvory-Sobol - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Charlotte Hedskog - Consulting: Gilead Sciences<br />

Ming Lin - Employment: Gilead Sciences, Inc.<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Jenny C. Yang - Employment: Gilead Sciences, Inc<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Ziad Younes - Consulting: Gilead; Grant/Research Support: BMS, AbbVie, Gilead,<br />

Vertex, Vertex, Idenix, Merck, Janssen, Tobira; Speaking and Teaching:<br />

Gilead, AbbVie<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

The following authors have nothing to disclose: Michael Ryan<br />

220<br />

Evolution Of Resistance Associated Variants During<br />

Initial Treatment, Viral Failure, And Re-Treatment With<br />

Directly Acting Antiviral Therapy (NIH SYNERGY Trial)<br />

Sarah Kattakuzhy 2,3 , Eleanor Wilson 2,3 , Sreetha Sidharthan 3 ,<br />

Mary McLaughlin 4 , Rachel Silk 2 , Hongmei Mo 5 , Anu Osinusi 5 ,<br />

Anita Kohli 6,1 , Henry Masur 3 , Shyam Kottilil 2,3 ; 1 Laboratory<br />

of Immunoregulation, NIH, Leidos Biomedical Research Inc.,<br />

Bethesda, MD; 2 Infectious Disease, University of Maryland School<br />

of Medicine, Institute of Human Virology, Baltimore, MD; 3 Critical<br />

Care Medicine, National Institutes of Health, Bethesda, MD;<br />

4 National Institute of Allergy and Infectious Disease, Bethesda,<br />

MD; 5 Gilead Sciences, Inc., Foster City, CA; 6 Hepatology, Creighton<br />

University, Phoenix, AZ<br />

Introduction: The natural history of resistance associated<br />

variants (RAVs), and their impact on treatment in the era of<br />

directly-acting antiviral therapy (DAA) remains unclear. Recent<br />

<strong>studies</strong> have suggested that NS5A RAVs exhibit long-term persistence<br />

and may affect response to NS5A inhibitor based<br />

therapy. The current analysis seeks to outline the evolution<br />

of RAVs to the major HCV therapeutic target regions (NS3,<br />

NS5A, and NS5B), evaluate the viral evolution of these mutations<br />

over time, and assess the impact on initial treatment and<br />

retreatment outcomes Methods: In this single-center, open-la-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 323A<br />

bel, phase 2a trial, 34 HCV mono-infected participants with<br />

F0-F2 liver fibrosis, with previous viral relapse to 4-6 weeks<br />

of LDV/SOF + GS-9451 +/- GS-9669, were enrolled and<br />

re-treated with LDV/SOF for 12 weeks. Deep sequencing of<br />

NS3, NS5A, and NS5B regions was performed by Illumina<br />

next generation sequencing technology with a detection limit of<br />

1% prevalence. Samples were sequenced at three time pointsprior<br />

to initial therapy, at the time of relapse, and prior to<br />

retreatment; two-way ANOVA testing was used to evaluate<br />

significant change in RAVs Results: 33 patient samples were<br />

sequenced at each timepoint. At baseline, 10 patients (30%),<br />

9 patients (27%) and 1 patient (3%) had RAVs to NS3, NS5A,<br />

and NS5B respectively. At a median of 4 weeks post treatment,<br />

patients experienced viral relapse to short duration therapy.<br />

At viral relapse, 11 patients (33%) and 2 patients (6.1%) had<br />

NS3 and NS5B RAVs respectively. The number of patients with<br />

NS5A RAVs increased significantly to 28 patients (84%), and<br />

emergent mutations L31M (+30.3%; 95% CI: 7.8-52.7) and<br />

Q30R (+27.3%; 95% CI: 4.9-49.7) demonstrated a significant<br />

increase from baseline to viral relapse. After a median interval<br />

of 22 weeks (IQR 18-23), patients were retreated. At retreatment<br />

baseline,10 patients (30%), 28 patients (84%), and 3<br />

patients (9%) had NS3, NS5A, and NS5B RAVs respectively,<br />

with no significant change in profile of mutants compared to<br />

the time of viral relapse. There was no significant difference in<br />

number of RAVs per sample along the three timepoints. In this<br />

retreatment cohort, 31 patients out of 34 (91% ITT; 96%mITT)<br />

achieved sustained virologic response Conclusions: In this small<br />

cohort, there was a moderate baseline rate of NS3 and NS5A<br />

resistance, and minimal NS5B resistance. While failure to short<br />

duration therapy was characterized by persistent and emergent<br />

NS5A mutants, retreatment outcomes with longer duration therapy<br />

were not affected by the presence of these RAVs<br />

Disclosures:<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Anu Osinusi - Employment: gilead sciences<br />

The following authors have nothing to disclose: Sarah Kattakuzhy, Eleanor<br />

Wilson, Sreetha Sidharthan, Mary McLaughlin, Rachel Silk, Anita Kohli, Henry<br />

Masur, Shyam Kottilil<br />

221<br />

Polymorphisms in irisin-producing FNDC5 are associated<br />

with the severity of HCV genotype 3-induced steatosis<br />

Stephanie Bibert 2 , Claudio De Vito 3 , Beat Mullhaupt 4 , David<br />

Semela 5 , Jean-Francois Dufour 6 , Markus H. Heim 7 , Darius<br />

Moradpour 8 , Andreas Cerny 9 , Raffaele Malinverni 10 , Pierre-Yves<br />

Bochud 2 , Francesco Negro 1 ; 1 Gastroenterology, Hepatology and<br />

Clinical pathology, University Hospitals, Geneva-14, Switzerland;<br />

2 Service of Infectious Diseases, University Hospital, Lausanne, Switzerland;<br />

3 Clinical pathology, University Hospital, Geneva, Switzerland;<br />

4 Gastroenterology and hepatology, University Hospital,<br />

Zurich, Switzerland; 5 Gastroenterology, Cantonal Hospital, St.<br />

Gallen, Switzerland; 6 Visceral surgery and medicine, Inselspital,<br />

Berne, Switzerland; 7 Gastroenterology, Univresity Hospital, Basel,<br />

Switzerland; 8 Gastroenterology, University Hospital, Lausanne,<br />

Switzerland; 9 Epatocentro, Lugano, Switzerland; 10 Medicine,<br />

Pourtalès Hospital, Neuchatel, Switzerland<br />

Background. Steatosis, commonly observed among chronically<br />

HCV-infected patients, is associated with an increased risk of<br />

hepatocellular carcinoma development and cardiovascular morbidity.<br />

While metabolic steatosis is associated with changes in<br />

host metabolism including obesity, hyperlipidemia and insulin<br />

resistance, steatosis in HCV genotype 3 infection is triggered<br />

by the virus. Hence, both host and viral factors contribute to the<br />

development of steatosis. To date, only few host genetic polymorphisms<br />

have been associated with HCV-related steatosis.<br />

Single nucleotide polymorphisms (SNPs) within the gene coding<br />

for fibronectin type II domain-containing 5 transmembrane<br />

receptor (FNDC5) have recently been linked to NAFLD in Caucasian<br />

patients. Moreover, recent <strong>studies</strong> showed that irisin, the<br />

cleaved and secreted form of FNDC5, could prevent hepatic<br />

steatosis in NAFLD patients. Methods. We analyzed the role<br />

of 4 FNDC5 gene-tagging polymorphisms as well as PNPLA3<br />

rs738409, IFNL3 rs12980275 and TM6SF2 rs3794991 in<br />

548 Caucasian HCV-infected patients with available liver<br />

biopsy. SNPs were extracted from a genome-wide association<br />

dataset. Genetic associations with the presence and severity<br />

of steatosis were assessed by univariate and multivariate logistic<br />

regression after adjustment for relevant covariates. Results.<br />

We confirm that PNPLA3 rs738409, IFNL3 rs12980275, and<br />

TM6SF2 rs3794991 are independent predictors of steatosis<br />

in 548 patients with HCV genotypes other than 3. In addition,<br />

we found that carriage of an SNP in FNDC5 increases the<br />

risk of steatosis only in patients infected with HCV genotype<br />

3 (odds ratio [OR] = 7.08, 95% confidence interval [CI] =<br />

1.44-34.82, p = 0.016) and appears to be associated with<br />

the severity of steatosis (OR = 5.45, 95% CI = 1.46-20.40, p<br />

= 0.012), although the limited number of patients prevented us<br />

from drawing firm conclusions. Conclusion. Our data suggest<br />

that FNDC5 could be specifically involved in the development<br />

and severity of steatosis in HCV genotype 3 infection.<br />

Disclosures:<br />

Beat Mullhaupt - Consulting: MSD, Novartis, MSD, Janssen; Grant/Research<br />

Support: Bayer, Gillead<br />

Jean-Francois Dufour - Advisory Committees or Review Panels: Bayer, BMS, Gilead,<br />

AbbVie, Novartis, Sillagen, Genfit<br />

Pierre-Yves Bochud - Speaking and Teaching: MSD, Gilead<br />

Francesco Negro - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

MSD, Gilead, AbbVie; Grant/Research Support: Gilead<br />

The following authors have nothing to disclose: Stephanie Bibert, Claudio De<br />

Vito, David Semela, Markus H. Heim, Darius Moradpour, Andreas Cerny, Raffaele<br />

Malinverni<br />

222<br />

Fast and deep early HCV-RNA decay is accompanied by<br />

fast aminotransferases normalization values in cirrhotic<br />

HCV-1 patients treated with paritaprevir/r–ombitasvir<br />

and dasabuvir with ribavirin: a potential role of baseline<br />

major resistance mutations<br />

Valeria Cento 1 , Elisabetta Teti 2 , Elisa Biliotti 3 , Francesco Fiore 7 ,<br />

FrancescoPaolo Antonucci 1 , Marianna Aragri 1 , Carlo F. Magni 4 ,<br />

Daniele Di Paolo 5 , Donatella Palazzo 3 , Katia Yu La Rosa 1 , Francesca<br />

De Luca 1 , Maria Chiara Sorbo 1 , Velia Chiara Di Maio 1 ,<br />

Paolo Casalino 6 , Cesare Sarrecchia 2 , Lara Lambiase 7 , Laura<br />

Gianserra 7 , Paola Perinelli 3 , Loredana Sarmati 2 , Guido A. Gubertini<br />

4 , Mario Angelico 5 , Caterina Pasquazzi 7 , Massimo Andreoni 2 ,<br />

Gloria Taliani 3 , Carlo Federico Perno 1 , Francesca Ceccherini-Silberstein<br />

1 ; 1 Experimental Medicine and Surgery, University of Rome<br />

“Tor Vergata”, Rome, Italy; 2 Infectious Diseases, Polyclinic “Tor<br />

Vergata” Foundation, Rome, Italy; 3 Infectious and Tropical diseases,<br />

“Umberto I” Hospital, Rome, Italy; 4 Division of Infectious<br />

Disease, Hospital Sacco of Milan, Milan, Italy; 5 Hepatology, Polyclinic<br />

“Tor Vergata” Foundation, Rome, Italy; 6 Laboratory Medicine,<br />

Polyclinic “Tor Vergata” Foundation, Rome, Italy; 7 Infectious<br />

Diseases, Sant’Andrea Hospital, Rome, Italy<br />

Study purpose. This study aims to define in a real-life setting<br />

the early kinetics of HCV-RNA decay and hepatic “cell cure”<br />

during administration of paritaprevir/r-ombitasvir-dasabuvir<br />

(3D regimen) with ribavirin in cirrhotic patients (pts), and<br />

to investigate the role of pre-existing resistance associated


324A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

variants (RAVs). Methods. 22 HCV-1 (1a/1b/1g=7/14/1)<br />

Child A/B (N=20/2) cirrhotic pts received 3D plus ribavirn<br />

for 12/24 weeks. Baseline (BL) presence of NS3-NS5A-NS5B<br />

RAVs was assessed by direct sequencing. HCV-RNA and aminotransferases<br />

(AT) were quantified at BL and early time-points<br />

(4h-6/8h-24h-48h-72h-96h-1w-2w-3w-4w). Ad interim results<br />

at 4w-treatment are presented. Results. BL median (IQR) HCV-<br />

RNA was 5.7 (5.3-6.2) logIU/ml and 20 pts had high AT.<br />

Already at 8h after 3D-administration, a fast HCV-RNA decay<br />

>1 log IU/ml was observed in 13/15 (87%) pts; at that time,<br />

12% of pts had AT values already normalized, while another<br />

57% showed a remarkable drop in their value compared to<br />

BL. By 24h, the median (IQR) HCV-RNA decay was 2.8 (2.4-<br />

3.0) logIU/ml; AT normalization occurred in 75% of pts and<br />

an additional 11% showed a remarkable reduction in their<br />

values. By 1w, HCV-RNA decay slowed down (median [IQR]<br />

BL-1w decay= 3.5 [3.0-3.8] logIU/ml). Of great relevance,<br />

all pts at this time had AT within normal range. At BL, single<br />

3D-RAVs were detected in 6/18 pts ( NS5A<br />

Q30R/L31M/<br />

A92T; NS5B<br />

C316N; NS3<br />

V36L/R155K), and double 3D-RAVs in<br />

3 (2pts: NS5B<br />

L159F+C316N; 1pt: NS5B<br />

C316N+ NS5A<br />

P58S). By<br />

8h, the 2 pts with major RAVs ( NS5A<br />

L31M; NS3<br />

R155K) were the<br />

only with 0.6log HCV-RNA decay vs. median (IQR) HCV-RNA<br />

decay 1.6 (1.4-1.9) logIU/ml in other pts (p=0.03). The same<br />

trend was maintained at 24h (median [IQR] HCV-RNA decay<br />

BL-24h=2.3 [2.3-2.4] vs. 2.8 [2.5-3.0] logIU/ml; p=0.12). At<br />

1w, HCV-RNA still >1000 IU/ml was found only in the 2 pts<br />

with 2 BL major RAVs; none achieved RVR. RVR achievement<br />

was positively predicted by early HCV-RNA kinetics at 48h<br />

and 2w of 3D administration, with a trend for BL HCV-RNA<br />

(p 48h<br />

=0.010, p 2w<br />

=0.003, p BL<br />

=0.14). Conclusions. In cirrhotic<br />

HCV-1 pts, 3D treatment leads to a fast and deep HCV-RNA<br />

decay, along with a remarkably fast normalization of AT<br />

values, consistent with the hypothesis of a drug-driven rapid<br />

“cure” of infected cells (HCV has no latent reservoir), rather<br />

than their elimination. Major BL RAVs may affect the kinetics of<br />

decay, but not necessarily the virological outcome. This result<br />

warrants further <strong>studies</strong>, due to the potential importance that<br />

rare, but specific major NS3/NS5A/NS5B RAVs may have in<br />

term of treatment duration and success rate of all-oral regimens.<br />

Disclosures:<br />

Mario Angelico - Advisory Committees or Review Panels: Gilead, Janssen;<br />

Grant/Research Support: Roche; Speaking and Teaching: GSK, Roche, Gilead,<br />

Novartis, BMS, Bayer<br />

Gloria Taliani - Speaking and Teaching: ROCHE, Merck, BMS, Gilead, Jannsen,<br />

Novartis, AbbVie<br />

Francesca Ceccherini-Silberstein - Consulting: Gilead, ViiV; Grant/Research Support:<br />

Merck Sharp & Dohme; Speaking and Teaching: Janssen, Abbvie, Roche<br />

diagnosticis<br />

The following authors have nothing to disclose: Valeria Cento, Elisabetta Teti,<br />

Elisa Biliotti, Francesco Fiore, FrancescoPaolo Antonucci, Marianna Aragri,<br />

Carlo F. Magni, Daniele Di Paolo, Donatella Palazzo, Katia Yu La Rosa, Francesca<br />

De Luca, Maria Chiara Sorbo, Velia Chiara Di Maio, Paolo Casalino,<br />

Cesare Sarrecchia, Lara Lambiase, Laura Gianserra, Paola Perinelli, Loredana<br />

Sarmati, Guido A. Gubertini, Caterina Pasquazzi, Massimo Andreoni, Carlo<br />

Federico Perno<br />

223<br />

Rescue of hepatocytes from lethal mitochondrial damage<br />

by targeted transplantation of mitochondria<br />

Nidhi Gupta, David Wu, Catherine H. Wu, George Wu; UCONN<br />

Health, Farmington, CT<br />

Background: Mitochondrial defects in hepatocytes can result<br />

in severe liver dysfunction and death. There is no method to<br />

repair or replace damaged mitochondria. Hepatocytes have<br />

cell surface asialoglycoprotein receptors (AsGR) which internalize<br />

asialoglycoproteins within endosomes. Some bacteria produce<br />

endosomolytic peptides to escape endosomes. The aim of<br />

this study was to determine whether hepatocytes with destroyed<br />

mitochondria could be rescued from cell death by targeting<br />

functional mitochondria by AsGR-mediated endocytosis. Methods:<br />

An asialoglycoprotein, asialo-orosomucoid (AsOR), was<br />

fluorescently labeled and covalently linked to polylysine to create<br />

a conjugate Fl-AsOR-PL that could bind mitochondria to form<br />

complexes. Huh 7 AsGR (+) and SK Hep1 AsGR (-) cells were<br />

treated with a mitochondrial toxin to form Huh 7-mito (-) and<br />

SK Hep1-mito (-) cells which required supplemental medium for<br />

survival. An endosomolytic peptide, LLO, was conjugated to<br />

AsOR to form AsOR-LLO. Complexed mitochondria and AsOR-<br />

LLO were co-incubated with cells for 2 h without supplemental<br />

media. Fluorescence, mitochondrial and cell DNA, and mitochondrial<br />

respiration were measured at various times. Results:<br />

Fl-AsOR-PL conjugate bound stably to mitochondria. Incubation<br />

of Fl-AsOR-PL-mitochondria complexes with Huh 7-mito (-), but<br />

not SK Hep1-mito (-) cells increased fluorescence in Huh7-mito<br />

(-) cells from 6,400 at 1 h to 11,000 units at 2 h. Co-administration<br />

of complexed mitochondria with AsOR-LLO conjugate<br />

in Huh 7-mito (-) cells increased fluorescence from 14,000 at<br />

1 h to >30,000 units at 2 h (p 9,700-fold over control at 7 d<br />

(p90% of<br />

controls by 10 d. Conclusions: This is the first demonstration of<br />

normalization of mitochondrial respiration, and rescue of mitochondria-damaged<br />

hepatocytes by targeted uptake of normal<br />

mitochondria by receptor-mediated endocytosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nidhi Gupta, David Wu, Catherine<br />

H. Wu, George Wu<br />

224<br />

Predicting In Vivo Hepatotoxicity for a Pair of Related<br />

Compounds with a Mechanistic Model of Drug-induced<br />

Livery Injury<br />

Diane M. Longo 1,2 , Yuching Yang 1,2 , Brett A. Howell 1,2 , Scott Q.<br />

Siler 1,2 , Paul B. Watkins 1,2 ; 1 The Hamner-UNC Institute for Drug<br />

Safety Sciences, Research Triangle Park, NC; 2 DILIsym® Services,<br />

Research Triangle Park, NC<br />

Background: Tolcapone and entacapone are catechol-O-methyltransferase<br />

(COMT) inhibitors developed for the treatment<br />

of Parkinson’s disease. While clinical <strong>studies</strong> have revealed<br />

the potential of tolcapone to cause hepatotoxicity, no toxicity<br />

issues have been reported with entacapone. In vitro assays<br />

have shown that both compounds are capable of uncoupling<br />

the mitochondria proton gradient in a dose-dependent manner.<br />

In addition, both drugs cause modest inhibition of the bile salt<br />

export pump (BSEP). Aim: To integrate pharmacokinetic data<br />

and in vitro toxicity data via the use of a multi-scale, mechanistic<br />

model to simulate the in vivo response in humans to tolcapone<br />

and to entacapone. Methods: DILIsym ® , a predictive, mechanistic,<br />

mathematical model of drug-induced liver injury (DILI), was


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 325A<br />

used to simulate liver drug concentrations and DILI responses<br />

after administration of tolcapone (200mg, 3 times per day)<br />

or entacapone (200mg, 8 times per day) in a virtual human<br />

population sample (SimPops) of 229 simulated individuals.<br />

The DILIsym ® model consists of various smaller submodels that<br />

are mathematically integrated to simulate an organism-level<br />

response. The current work utilized several components of<br />

DILIsym®, including a physiologically-based pharmacokinetic<br />

(PBPK) submodel representing drug distribution, submodels of<br />

mitochondrial dysfunction and toxicity, bile acid physiology<br />

and pathophysiology, hepatocyte life cycle, and liver injury<br />

biomarker submodels. The DILIsym ® model is developed and<br />

maintained through the DILI-sim Initiative, which is a public-private<br />

partnership involving scientists in academia, industry, and<br />

the FDA. Results: In the simulated population of humans, tolcapone<br />

administration resulted in hepatocellular ATP reductions<br />

and increases in serum alanine transaminase >3x the upper<br />

limit of normal in 2% of the population. In contrast, entacapone<br />

administration did not elicit any hepatotoxicity in the<br />

simulated human population. This difference in hepatotoxic<br />

potential between tolcapone and entacapone is consistent with<br />

clinical observations. Conclusions: Our results demonstrate the<br />

utility of applying mechanistic modeling to distinguish hepatotoxic<br />

drugs from safer alternatives. This study also illustrates<br />

the capability of DILIsym ® to combine clinical data, in vitro<br />

data, predicted liver compound exposure, and inter-patient<br />

differences to provide an account of how compound exposure,<br />

biological variability, and multiple hepatotoxicity mechanisms<br />

may come together to result in DILI.<br />

Disclosures:<br />

Paul B. Watkins - Consulting: Abbott, Actelion, Boerringer-Ingelheim, Cempra,<br />

Alecra, Roche, Merck, Reservlogix, Intercept, Janssen, Novartis, Otsuka, Pfizer,<br />

Sanolfi, Takeda, UCB, Bristol-Myers Squibb, GSK<br />

The following authors have nothing to disclose: Diane M. Longo, Yuching Yang,<br />

Brett A. Howell, Scott Q. Siler<br />

225<br />

HLA-A*33:01 is strongly associated with drug-induced<br />

liver injury (DILI) due to terbinafine and several other<br />

unrelated compounds<br />

Guruprasad P. Aithal 1 , Paola Nicoletti 2 , Einar Bjornsson 3 , M. I.<br />

Lucena 23,4 , Raul J. Andrade 23,4 , Jane Grove 1 , C. Stephens 23 , Par<br />

Hallberg 5 , Anke H. Maitland-van der Zee 28 , Jennifer H. Martin 6,7 ,<br />

Ingolf Cascorbi 8 , John F. Dillon 9 , Tarja Laitinen 10 , Dominique<br />

G. Larrey 11 , Mariam Molokhia 12 , Gerd A. Kullak-Ublick 13 , Luisa<br />

Ibanez 14 , Munir Pirmohamed 15 , Shengying Qin 16 , Ashley Sawle 2 ,<br />

Fernando Bessone 21 , Nelia Hernández 22 , Andrew Stolz 24 , Naga<br />

P. Chalasani 25 , Jose Serrano 26 , Huiman X. Barnhart 27 , Robert J.<br />

Fontana 17 , Paul Watkins 18 , Thomas J. Urban 19 , Ann K. Daly 20 ;<br />

1 NIHR Nottingham Digestive Diseases Biomedical Research Unit,<br />

Nottingham University Hospitals NHS Trust, Nottingham, United<br />

Kingdom; 2 Columbia University, New York, NY; 3 Landspitali University<br />

Hospital, Reykjavik, Iceland; 4 CIBERehd, Madrid, Spain;<br />

5 Uppsala University, Uppsala, Sweden; 6 University of Queensland,<br />

Brisbane, QLD, Australia; 7 Princess Alexandra Hospital, Brisbane,<br />

QLD, Australia; 8 University Hospital Schleswig-Holstein, Kiel, Germany;<br />

9 Ninewells Hospital and medical school, Dundee, United<br />

Kingdom; 10 Helsinki University Central Hospital, Helsinki, Finland;<br />

11 Hôpital Saint Eloi, Montpellier, France; 12 Kings College London,<br />

London, United Kingdom; 13 University of Zurich, Zurich, Switzerland;<br />

14 Hospital Universitari Vall d’Hebron, Barcelona, Spain;<br />

15 University of Liverpool, Liverpool, United Kingdom; 16 Shanghai<br />

Jiao Tong university, Shanghai, China; 17 University of Michigan,<br />

Ann Arbor, MI; 18 Hamner-UNC institute for Drug Safety Sciences,<br />

Durham, NC; 19 UNC Eshelman School of Pharmacy, Chapel<br />

Hill, NC; 20 Newcastle University, Newcastle, United Kingdom;<br />

21 Facultad de ciencias Medicas, Universidad nacional de Rosario,<br />

Rosario, Argentina; 22 Universidad de la Republica, Mentevideo,<br />

Uruguay; 23 University of Malaga, Malaga, Spain; 24 University of<br />

Southern California, Los Angeles, CA; 25 Indiana University, Indianapolis,<br />

IN; 26 National Institute of Diabetes and Digestive and<br />

Kidney Diseases, Bethesda, MD; 27 Duke Clinical Research Institute,<br />

Durham, NC; 28 Utrecht University, Utrecht, Netherlands<br />

Background: Recent genome wide association (GWA) <strong>studies</strong><br />

have revealed a number of HLA alleles as risk factors for DILI<br />

from specific drugs. However, untill now, the only genomewide<br />

significant signal seen in large DILI cohorts containing<br />

cases related to a range of different drugs have been HLA<br />

signals due to DILI from amoxicillin-clavulanate and flucloxacillin.<br />

Aim: To perform a GWA analysis of DILI cases enrolled by<br />

iDILIC and DILIN investigators that were not related to amoxicillin-clavulanate<br />

or flucloxacillin. Methods: Principal component<br />

analysis showed that 878 Caucasian cases (mean age 53<br />

years, 338 males; 540 females) clustered in Italians, Spanish<br />

and Northern Europeans groups could be matched with<br />

10605 population based controls. Cases and controls were<br />

genotyped at different times using three platforms (Illumina 1M,<br />

HCE and OMX). To compare the different cohorts, the genotyped<br />

datasets were phased and imputed. For each individual,<br />

HLA alleles were also predicted. After applying quality checks,<br />

5,429,170 variants and 217 HLA alleles were tested for association.<br />

We set the genomewide significance p-value threshold<br />

to 5*10 -8 . Results: An asociation was demonstrated between a<br />

single nucleotide polymorphism (SNP) that tags HLA-A*33:01<br />

and DILI treated as a single phenotype with an odds ratio<br />

(OR) of 2.65 [95%CI 1.87-3.76] (p=3.9x10 -8 ). This association<br />

was consistent across Italian, Spanish and Northern Europeans<br />

populations. When subdivided by DILI phenotype, the<br />

genomewide significant signal was seen only for a cholestatic/<br />

mixed DILI group and not for hepatocellular DILI. Further analysis<br />

showed that HLA-A*33:01 was a strong risk factor for<br />

DILI due to the anti-fungal agent terbinafine (n=14, OR 40.53


326A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

[95%CI 12.51-288.9] (p=6.7x10 -10 )). Associations were also<br />

found between the risk allele and DILI from fenofibrate (n=7,<br />

OR 58.7[95%CI 12.31-279.8](p=3.2x10-7)) and ticlopidine<br />

(n=5, OR 163.1[95%CI 16.2-1642.0] (p=0.00002)). Suggestive<br />

signals were found also for sertraline (n=5), enalapril<br />

(n=4), methyldopa (n=4) and erythromycin (n=10). We confirmed<br />

by sequence-based typing 99% of the predictions for 34<br />

of 46 samples whose DILI was due to those seven drugs. HLA<br />

typing of a limited set of additional cases (n=13) showed 4<br />

further HLA-A*33 positive cases relating to terbinafine and sertraline.<br />

Conclusions: The novel genome-wide significant association<br />

of HLA-A*33:01 with all cause DILI coupled with its<br />

strong association with liver injury from terbinafine and other<br />

unrelated compounds confirms an important role for adaptive<br />

immunity in DILI pathogenesis and widens the range of drugs<br />

associated with idiosyncratic DILI for which HLA is a risk factor.<br />

Disclosures:<br />

Dominique G. Larrey - Advisory Committees or Review Panels: BAYER, SANOFI,<br />

PFIZER, SERVIER-BIG, AEGERION, MMV, BIAL-QUINTILES, TEVA, ORION, NEG-<br />

MA-LERADS, ASTELLAS, ASTRAZENECA, DNDI, GSK, J AND J; Board Membership:<br />

BMS, GILEAD, ABBVIE, BMS, GILEAD, ITREAS, MMS; Grant/Research<br />

Support: GILEAD, MSD, BMS, ABBVIE, TIBOTEC/JANSSEN, BMS<br />

Mariam Molokhia - Grant/Research Support: SAEC<br />

Fernando Bessone - Advisory Committees or Review Panels: Schering Plough,<br />

Gilead, Glaxo, MSD, Janssen; Speaking and Teaching: Bristol Myers Squibb,<br />

Janssen, Bayer, Gilead, Abbvie<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

Robert J. Fontana - Consulting: GlaxoSmithKline, CLDF; Grant/Research Support:<br />

Gilead, vertex, BMS, Jansen, Gilead<br />

The following authors have nothing to disclose: Guruprasad P. Aithal, Paola<br />

Nicoletti, Einar Bjornsson, M. I. Lucena, Raul J. Andrade, Jane Grove, C. Stephens,<br />

Par Hallberg, Anke H. Maitland-van der Zee, Jennifer H. Martin, Ingolf<br />

Cascorbi, John F. Dillon, Tarja Laitinen, Gerd A. Kullak-Ublick, Luisa Ibanez,<br />

Munir Pirmohamed, Shengying Qin, Ashley Sawle, Nelia Hernández, Andrew<br />

Stolz, Jose Serrano, Huiman X. Barnhart, Paul Watkins, Thomas J. Urban, Ann<br />

K. Daly<br />

226<br />

Sodium 4-phenylbutyric acid protects against mice from<br />

acetaminophen-induced liver injury through attenuation<br />

of endoplasmic reticulum stress.<br />

Hiromi Kusama, Kazuyoshi Kon, Kenichi Ikejima, Akira Uchiyama,<br />

Kumiko Arai, Shunhei Yamashina, Sumio Watanabe; Gastroenterology,<br />

Juntendo University School of Medicine, Tokyo, Japan<br />

Background: Acetaminophen (AAP) overdose consequently<br />

elicits massive necrosis of hepatocytes, resulting in acute liver<br />

failure. Recent lines of reports have suggested that endoplasmic<br />

reticulum (ER) stress is contributed to AAP hepatotoxicity; however,<br />

therapeutic targeting toward ER stress has not been well<br />

evaluated. Therefore, here we investigate the effect of 4-phenylbutyric<br />

acid (PBA), a chemical chaperone, on AAP-induced<br />

liver injury in mice. Methods: Male, 8 week-old C57Bl/6 mice<br />

were exposed to AAP (450 mg/kg BW, i.p.), and serially sacrificed<br />

up to 12 h later. Some mice were repeatedly injected<br />

with PBA (120 mg/kg BW, i.p.) every 3 h starting at 0.5 h<br />

after AAP challenge, since PBA demonstrates relatively rapid<br />

decay in vivo. Liver pathology was evaluated by H-E staining,<br />

and serum aminotransferases levels were determined.<br />

Oxidative stress was assessed by immunohistological staining<br />

for 4-hydroxynonenal (HNE). Total glutathione (GSH) content<br />

in the liver was measured colorimetrically. Apoptotic cell<br />

death was visualized by TdT-mediated dUTP nick end labeling<br />

(TUNEL). Phosphorylation of c-jun N—terminal kinase (p-JNK/<br />

JNK) and cleavage of activating transcription factor (ATF) 6<br />

were detected by Western blotting. Results: A single injection<br />

of AAP caused massive necrosis of hepatocytes predominantly<br />

in pericentral area in 12 h as expected; however, subsequent<br />

repeated injections of PBA dramatically ameliorated these<br />

pathological changes in the liver. Indeed, the ALT levels were<br />

elevated to 11200 ± 455 IU/L at 6 h after AAP, whereas the<br />

values only reached 4510 ± 434 IU/L in mice treated with<br />

repeated PBA (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 327A<br />

showed enhanced mitochondrial translocation of JNK accompanied<br />

by an increase in the release of mitochondrial enzymes,<br />

such as apoptosis-inducing factor and endonuclease G, into<br />

the cytosol, which is indicative of increased mitochondrial dysfunction<br />

and subsequent nuclear DNA fragmentation. Finally,<br />

in vitro experiments showed that Gab1-deficient hepatocytes<br />

were more susceptible to APAP-induced mitochondrial dysfunction<br />

and cell death, suggesting that hepatocyte Gab1 is a<br />

direct target of APAP-induced hepatotoxicity. Conclusion: Our<br />

current data demonstrate that hepatocyte Gab1 plays a critical<br />

role in controlling the balance between hepatocyte death and<br />

compensatory hepatocyte proliferation during APAP-induced<br />

liver injury. Thus, hepatocyte Gab1 would be a potential therapeutic<br />

target for APAP-induced liver injury.<br />

Disclosures:<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yuichi Yoshida, Kunimaro Furuta,<br />

Satoshi Ogura, Tomohide Kurahashi, Mayumi Egawa, Shinichi Kiso, Yoshihiro<br />

Kamada<br />

228<br />

Lipotoxicity Accounts for Liver Injury due to Etomoxir:<br />

Application of Systems Toxicology Modeling of a Patient<br />

Population<br />

Scott Q. Siler 1,2 , Yuching Yang 2,1 , Diane M. Longo 2,1 , Brett A.<br />

Howell 2,1 , Paul Watkins 2 ; 1 DILI-sim, Castro Valley, CA; 2 Institute<br />

for Drug Safety Sciences, The Hamner Institutes-University of North<br />

Carolina, Research Triangle Park, NC<br />

Background: Etomoxir reduces fatty acid oxidation by inhibiting<br />

mitochondrial carnitine palmitoyltransferase 1 (CPT1).<br />

Etomoxir was originally developed for treatment of type 2 diabetes<br />

and congestive heart failure (CHF). In the “Etomoxir for<br />

the Recovery of Glucose Oxidation” (ERGO) clinical trial for<br />

CHF, elevations in serum ALT > 5 X ULN were observed in<br />

4 patients receiving active treatment, causing termination of<br />

the trial. It has remained unclear whether etomoxir hepatotoxicity<br />

resulted from direct mitochondrial toxicity or lipotoxicity<br />

due to the accumulation of lipids. Aim: To identify the<br />

most likely mechanistic determinant of etomoxir hepatotoxicity<br />

via simulations with DILIsym®. Methods: DILIsym®, a predictive,<br />

mechanistic, mathematical model of drug-induced liver<br />

injury (DILI), was used to simulate liver drug concentrations<br />

and DILI responses after administration of etomoxir (80 mg<br />

q.d.) in a virtual human population sample (SimPops) of 229<br />

simulated individuals. Quantitative fatty acid inhibition effects<br />

were based on in vitro data from the literature (Agius 1991).<br />

The DILIsym® model consists of various sub-models that are<br />

mathematically integrated to simulate patient response. The<br />

current work utilized several components of DILIsym®, including<br />

a physiologically-based pharmacokinetic (PBPK) sub-model<br />

representing drug distribution, sub-models of mitochondrial<br />

dysfunction and lipotoxicity, hepatocyte life cycle, and liver<br />

injury biomarkers. SimPops has been created by varying<br />

key parameters consistent with experimental data. DILIsym®<br />

is developed and maintained through the DILI-sim Initiative,<br />

a public-private partnership involving scientists in academia,<br />

industry, and the FDA. Results: In simulations where the model<br />

parameters excluded contributions of lipotoxicity to liver injury,<br />

no simulated patients were predicted to experience serum ALT<br />

increases. However, simulations including lipotoxicity showed<br />

an incidence of DILI that was similar to that seen in the ERGO<br />

trial, with 2-3% of simulated patients predicted to have ALT >5x<br />

ULN. The predicted ALT increases did not occur until after 4-6<br />

weeks of dosing, as was observed in the ERGO trial. Conclusions:<br />

The DILIsym® simulation results suggest that hepatotoxicity<br />

observed in the ERGO trial resulted from accumulation of<br />

lipids leading to lipotoxicity and not direct mitochondrial toxicity.<br />

This study illustrates the capability of DILIsym® to combine<br />

clinical data, in vitro data, predicted liver compound exposure,<br />

and inter-patient differences to provide insight into underlying<br />

mechanisms causing DILI.<br />

Disclosures:<br />

The following authors have nothing to disclose: Scott Q. Siler, Yuching Yang,<br />

Diane M. Longo, Brett A. Howell, Paul Watkins<br />

229<br />

HCV Core Gene Mutations with Heightened Liver Cancer<br />

Risk: Impact on Cellular Gene Expression<br />

Ahmed M. Elshamy 1 , Francis J. Eng 1 , Erin H. Doyle 1 , Arielle L. Klepper<br />

1 , Xiaochen Sun 1 , Angelo Sangiovanni 2 , Massimo Iavarone 2 ,<br />

Massimo Colombo 2 , Robert E. Schwartz 3 , Yujin Hoshida 4 , Andrea<br />

D. Branch 1 ; 1 Liver Diseases, Icahn School of Medicine at Mount<br />

Sinai, New York, NY; 2 M. & A. Migliavacca Center for Liver Disease<br />

and 1st Division of Gastroenterology, Fondazione IRCCS<br />

Cà Granda Ospedale Maggiore Policlinico, University of Milan,<br />

Milan, Italy; 3 Department of Medicine, Weill Cornell Medical College,<br />

Department of Physiology, Biophysics, and Systems Biology,<br />

Weill Cornell Medical College, New York, NY; 4 Liver Cancer<br />

Program, Tisch Cancer Institute, Division of Liver Diseases, Department<br />

of Medicine, Icahn School of Medicine at Mount Sinai, New<br />

York, NY<br />

Background and Aims: Patients infected by genotype-1b hepatitis<br />

C virus (HCV) with Q 70 and/or M 91 core gene mutations<br />

have an almost five-fold increased risk of developing hepatocellular<br />

carcinoma (HCC) and an increased susceptibility to<br />

insulin resistance. The elevated HCC risk persists despite a sustained<br />

virological response to treatment. These compelling clinical<br />

data prompted us to seek laboratory data corroborating<br />

the oncogenic effects of the Q 70 /M 91 mutations. Methods: The<br />

HCC-associated virus (Q 70 /M 91 ) and the control virus (R 70 /<br />

L 91 ) were studied in Huh-7 cells. Infected cells were cultured<br />

in media containing adult human serum to promote differentiation.<br />

Quantitative RT-PCR and genome-wide transcriptome<br />

profiling (HumanHT12 beadarray, Illumina) were used for RNA<br />

analysis. Modulated molecular pathways were determined by<br />

Gene Set Enrichment Analysis. For clinical validation, transcriptomes<br />

of infected cells were compared to those of liver<br />

biopsies of patients with early-stage HCV-related cirrhosis who<br />

were followed prospectively for up to 23 yr to monitor for<br />

HCC and other clinical outcomes (n=216). Results: The Huh-7<br />

cells cultured in human serum acquired several features of differentiated<br />

hepatocytes, including a cobble-stone morphology<br />

and significantly enhanced expression of albumin and other<br />

hepatocyte-specific genes. Their expression of albumin was<br />

similar to that of cultured primary human hepatocytes and ~<br />

20-fold higher than that of Huh-7 cells maintained in media<br />

containing fetal bovine serum. Surprisingly, although the two<br />

viruses were virtually identical in virological fitness, they had<br />

markedly different effects on cellular gene expression. The highrisk<br />

virus (Q 70 /M 91 ) enhanced expression of pathways associated<br />

with cancer and type II diabetes, while the control virus<br />

(R 70 /L 91 ) enhanced pathways associated with oxidative phosphorylation.<br />

Of special clinical relevance, the transcriptome of<br />

cells replicating the high-risk virus correlated significantly with<br />

an HCC-associated transcriptome in patients (Bonferroni-corrected<br />

P=0.03), but not with any other outcome-related transcriptome.<br />

The transcriptome of cells replicating the control<br />

virus did not correlate with any outcome-related transcriptome.


328A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Conclusions: These results provide direct experimental evidence<br />

of sequence-specific oncogenic effects of HCV, recapitulating<br />

epidemiologic data and emphasizing the link between HCV<br />

sequences and clinical outcomes. These findings justify the<br />

assessment of HCV mutations as prognostic indicators. This<br />

simple cell-based system may be useful for investigating mechanisms<br />

of hepatocarcinogenesis and potential interventions<br />

(DA031095, DK090317).<br />

Disclosures:<br />

Massimo Colombo - Advisory Committees or Review Panels: BRISTOL-MEY-<br />

ERS-SQUIBB, SCHERING-PLOUGH, ROCHE, GILEAD, BRISTOL-MEYERS-SQUIBB,<br />

SCHERING-PLOUGH, ROCHE, GILEAD, Janssen Cilag, Achillion; Grant/<br />

Research Support: BRISTOL-MEYERS-SQUIBB, ROCHE, GILEAD, BRISTOL-MEY-<br />

ERS-SQUIBB, ROCHE, GILEAD; Speaking and Teaching: Glaxo Smith-Kline,<br />

BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH, ROCHE, NOVARTIS, GILEAD,<br />

VERTEX, Glaxo Smith-Kline, BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH,<br />

ROCHE, NOVARTIS, GILEAD, VERTEX, Sanofi<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

The following authors have nothing to disclose: Ahmed M. Elshamy, Francis J.<br />

Eng, Erin H. Doyle, Arielle L. Klepper, Xiaochen Sun, Angelo Sangiovanni, Massimo<br />

Iavarone, Robert E. Schwartz, Yujin Hoshida<br />

230<br />

Voluntary exercise opposes accelerated hepatocarcinogenesis<br />

in obese, diabetic mice by improving metabolic<br />

regulation and fatty liver disease, not via AMPK/<br />

mTORC1<br />

Evi Arfianti 1 , Vanessa Barn 1 , W. G. Haigh 2 , Matthew M. Yeh 3 ,<br />

George N. Ioannou 2 , Narci C. Teoh 1 , Geoffrey C. Farrell 1 ; 1 Liver<br />

Research Group, Australian National University (ANU) Medical<br />

School at The Canberra Hospital, Canberra, ACT, Australia; 2 Division<br />

of Gastroenterology, University of Washington, Seattle, WA;<br />

3 Division of Pathology, University of Washington, Seattle, WA<br />

Background Obesity and diabetes are independent risk factors<br />

for HCC associated with non-alcoholic fatty liver disease<br />

(NAFLD), HCV and other disorders. Diethylnitrosamine<br />

(DEN)-induced hepatocarcinogenesis is enhanced in obese,<br />

diabetic Alms1 mutant (foz/foz) mice which develop NASH.<br />

Such accelerated hepatocarcinogenesis is associated with<br />

hyperinsulinemia, hyperglycemia, and Akt/mTORC1 activation,<br />

but rapamycin fails to prevent HCC. We investigated<br />

whether exercise sufficient to slow weight gain reduces growth<br />

of dysplastic hepatocytes and HCC development in mice<br />

genetically predisposed to obesity and diabetes. Methods<br />

Male foz/foz and wild-type (Wt) littermates were injected with<br />

DEN (10mg/kg i.p.) or vehicle (saline) at age 12 days. From<br />

4wks, they were randomly assigned to cages provided with an<br />

exercise wheel (EW) until age 12 or 24 wks, or without EW.<br />

Dysplastic hepatocytes were identified by GST-pi immunohistochemistry<br />

(IHC), protein expression by immunoblotting, gene<br />

expression by real-time PCR, glucose tolerance by i.p. glucose<br />

tolerance test, hepatic lipid content using HPLC. Results: foz/<br />

foz and Wt mice provided with in-cage exercise wheels ran<br />

an average of 4 km/day. In association with such physical<br />

activity, foz/foz weight gain slowed to that of Wt until 12<br />

wks; weight gain then accelerated but body weight remained<br />

lower than non-exercising mice. At 12wks, there was a significant<br />

reduction in GST-pi positive preneoplastic hepatocytes in<br />

exercising vs sedentary foz/foz mice. At 24 wks, more obese<br />

foz/foz mice had HCCs than lean Wt littermates (67% vs. 0,<br />

P6 tumors; 41.7%/41.7%/16.7% vs<br />

9.1%/18.1%/72.7%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 329A<br />

sion rate, and malignant macroscopic classification. Conclusion:<br />

CTGF is up-regulated through Ras/Mek/Erk pathway in<br />

HCC. CTGF promotes the progression of HCC and related with<br />

its malignant characteristics. CTGF could be a new therapeutic<br />

target against HCC.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yuki Makino, Takahiro Kodama,<br />

Minoru Shigekawa, Yugo Kai, Yasutoshi Nozaki, Tasuku Nakabori, Yoshinobu<br />

Saito, Satoshi Tanaka, Ryotaro Sakamori, Naoki Hiramatsu, Tomohide Tatsumi<br />

232<br />

p62 is a Key Regulator for Initiation and Development<br />

of Hepatocellular carcinoma<br />

Atsushi Umemura 1,2 , Yoshito Itoh 1 , Maria T. Diaz-Meco 3 , Jorge<br />

Moscat 3 , Michael Karin 2 ; 1 Department of Molecular Gastroenterology<br />

and Hepatology, Graduate School of Medical Science,<br />

Kyoto Prefectural University of Medicine, Kyoto, Japan; 2 Laboratory<br />

of Gene Regulation and Signal Transduction, Department of<br />

Pharmacology, University of California San Diego, San Diego,<br />

CA; 3 Sanford-Burnham Medical Research Institute, San Diego, CA<br />

p62, a scaffold protein interacting with various<br />

signaling pathways, accumulates and forms aggregates<br />

frequently observed in chronic liver diseases including NASH<br />

(non-alcoholic steatohepatitis), ASH (alcoholic steatohepatitis),<br />

and their progressed conditions, liver cirrhosis and hepatocellular<br />

carcinoma (HCC). p62 has also been reported as a<br />

critical oncoprotein involved in tumorigenesis and the development<br />

of many types of cancer including lung, breast, and<br />

kidney, whereas the tumorigenic role of p62 in the liver still<br />

remains unknown. It is reported that loss of p62 suppresses<br />

adenoma development in the autophagy knockout liver, importantly,<br />

however, this liver does not develop HCC. <br />

To elucidate the impact of p62 on HCC initiation and development,<br />

we decided to use two different animal models. First off,<br />

we injected a chemical carcinogen diethylnitrosamine (DEN),<br />

which induces multiple HCCs 7 months later, into liver-specific<br />

p62 knockout mice (p62KO) and their control mice (CON)<br />

to evaluate HCC development. We then established liver-specific<br />

TSC1 (tuberous sclerosis complex1)/p62 double-knockout<br />

mice (DKO). It is reported that liver-specific TSC1 knockout<br />

mice (TSC1KO) develop multiple HCCs spontaneously around<br />

10-12 months old. This means that DKO mice are useful in<br />

analyzing the role of p62 in hepatocarcinogenesis . CON mice developed multiple HCCs 7 months<br />

after DEN injection, but p62KO mice showed much lower incidence<br />

and smaller size of DEN-induced HCCs. Loss of TSC1 in<br />

the liver induced multiple HCCs as expected, but surprisingly,<br />

DKO mice did not developed HCCs nor even small tumors. In<br />

addition, liver injury and fibrosis observed in TSC1KO mice<br />

were attenuated in DKO livers. Among a number of signaling<br />

pathways, which p62 may be involved, Nrf2 was clearly activated<br />

in the livers of both DEN-injected CON and TSC1KO<br />

mice, and this activation was diminished in DEN-injected<br />

p62KO and DKO mice, respectively. Here we<br />

show that p62 plays important roles in HCC initiation as well<br />

as its development using two different HCC models. The Nrf2<br />

signaling pathway seems to be the most crucial in both models.<br />

Furthermore, additional p62 ablation ameliorated liver injury<br />

and fibrosis in TSC1KO mice. p62 is a well-known regulator of<br />

Nrf2 via its interaction with keap1. Since Nrf2 activation was<br />

suppressed by deletion of p62, the p62/keap1/Nrf2 signaling<br />

axis may profoundly contribute to hepatocarcinogenesis. In<br />

conclusion, p62 has a strong oncogenic function in the liver<br />

and is a promising target for HCC therapy.<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

The following authors have nothing to disclose: Atsushi Umemura, Maria T. Diaz-<br />

Meco, Jorge Moscat, Michael Karin<br />

233<br />

Sitagliptin, a Dipeptidyl Peptidase 4 inhibitor, Suppressed<br />

Tumor Progression with Down-regulation of Nrf<br />

Nuclear Expression in a Mouse Model of Non-alcoholic<br />

Steatohepatitis-related Hepatocellular Carcinoma<br />

Takumi Kawaguchi, Eitaro Taniguchi, Minoru Itou, Tetsuharu Oriishi,<br />

Takuji Torimura; Division of Gastroenterology, Department of<br />

Medicine, Kurume University School of Medicine, Kurume, Japan<br />

Background and Aims: Sitagliptin is an anti-diabetic agent classified<br />

as a dipeptidyl peptidase 4 (DPP4) inhibitor. Although<br />

sitagliptin is known to improve non-alcoholic steatohepatitis<br />

(NASH), the effects of sitagliptin on NASH-related hepatocellular<br />

carcinoma (HCC) remains unclear. In addition, it is still<br />

unclear whether sitagliptin affects Nrf, a transcription factor<br />

regulating tumor proliferation via reprogramming of glucose<br />

metabolism. The aims of this study are to investigate effects of<br />

sitagliptin on HCC progression and nuclear expression of Nrf<br />

in a mouse model of NASH-related HCC. Material and Methods:<br />

A mouse model of NASH-related HCC without genetic<br />

modification (STAM mice) was employed in this study. Eightweek<br />

old STAM mice were orally administrated with either<br />

sitagliptin (30 mg/kg/day: DPP4 group; n=7) or distilled<br />

water (CON group; n=7) for 10 weeks. Then, the incidence,<br />

number, and volume of HCC were evaluated by contrast-enhanced<br />

computed tomography and histopathological examination.<br />

Nuclear expression of Nrf in HCC and non-tumor tissues<br />

were evaluated by immunostaining and quantified by Image<br />

J (U. S. National Institutes of Health, Bethesda, MD). Results:<br />

All mice survived until the end of this study. No significant<br />

difference was seen in body weight and the amount of food<br />

intake between the DPP4 and CON groups during the course<br />

of this study. While, liver-to-body weight ratio was significantly<br />

lower in the DPP4 group than in the CON group (7.3±1.2% vs.<br />

11.0±1.8%, P


330A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Takumi Kawaguchi, Eitaro Taniguchi,<br />

Minoru Itou, Tetsuharu Oriishi, Takuji Torimura<br />

234<br />

Two-step forward genetic screens of transposon mutagenesis<br />

and pooled shRNA library identify novel tumor<br />

suppressor genes and potential therapeutic target in<br />

HCC<br />

Takahiro Kodama, Nancy Jenkins, Neal Copeland; Cancer Biology,<br />

Houston Methodist Research Institute, Houston, TX<br />

Background: The use of emerging sequencing and genomics<br />

technologies has recently identified many mutated and/or differentially<br />

expressed genes in HCC. However, it is still difficult<br />

to identify true driver genes for HCC because of the large number<br />

of passenger mutations and other gene-regulatory mechanism<br />

such as epigenetic regulation, SNP and non-coding RNA.<br />

To overcome this difficulty, we performed two-step forward<br />

genetic screens and aimed to discover novel tumor suppressor<br />

genes (TSGs) in HCC. Methods: A transposon mutagenesis<br />

screen was performed using mice that harbor the conditional<br />

SB system (loxP-STOP-loxP(LSL)-SB11 transposase and a T2/<br />

Onc2 or T2/Onc3 concatemer) and Alb-cre with or without<br />

sensitizing mutation of HBs-Ag transgene (Liver-SB/HBV or Liver-SB/noHBV).<br />

To validate candidate TSGs in a high-throughput<br />

manner, we subsequently performed in vivo small-scale<br />

pooled shRNA library screen using mouse immortalized liver<br />

progenitor cells (iLPCs). Results: We collected 334 and 146<br />

tumors from the Liver-SB/HBV and Liver-SB/noHBV mice,<br />

respectively, and identified 3523 and 2177 candidate cancer<br />

genes, respectively. List of 1917 candidate genes common<br />

in two transposon screening were narrowed down into 250<br />

genes that had non-synonymous mutation or mRNA downregulation<br />

reported in human HCC. Fifteen lentiviral pooled shRNA<br />

libraries (50 shRNAs/pool) targeting these 250 genes were<br />

introduced into iLPCs individually and assessed for their effects<br />

on subcutaneous tumor growth of transduced iLPCs in Nude<br />

mice. Eleven pooled libraries showed significant increase in<br />

tumor formation rate of iLPCs compared to negative control<br />

shRNA (16.7% - 32.1% vs 4.8%, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 331A<br />

236<br />

Clinical and Metabolic Effects Associated With Weight<br />

Loss and Obeticholic Acid In Nonalcoholic Steatohepatitis<br />

(NASH)<br />

Bilal Hameed 1 , Norah Terrault 1 , Ryan M. Gill 1 , Rohit Loomba 2 ,<br />

Naga P. Chalasani 3 , Jay H. Hoofnagle 4 , Mark L. Van Natta 5 ;<br />

1 UCSF (University of California San Francisco), San Francisco, CA;<br />

2 University of California San Diego, San Diego, CA; 3 Indiana Univeristy,<br />

Indianapolis, IN; 4 NIDDK, Bethesda, MD; 5 John Hopkins<br />

University, Baltimore, MD<br />

Background/Aims: In a 72-week, randomized controlled trial<br />

(FLINT), obeticholic acid (OCA) was superior to placebo in<br />

improving serum ALT levels and liver histology in 283 patients<br />

with NASH. OCA therapy also reduced weight (mean net loss<br />

= 2.3 kg). Because weight loss can improve hepatic histology,<br />

we have assessed the role of weight loss and OCA treatment<br />

in improving clinical and metabolic features of NASH. Methods:<br />

This secondary analysis was limited to 200 patients who<br />

underwent a baseline and end-of-treatment liver biopsy. Liver<br />

histology was graded using the standardized NASH CRN Scoring<br />

System for NASH activity. Weight loss was defined as ≥ 2<br />

kg loss by the end of treatment. Results: Weight loss occurred<br />

in 43 of 102 (42%) OCA vs 29 of 98 (30%) placebo-recipients<br />

(p=0.08). NAS improved more with OCA in patients<br />

who lost vs did not lose weight (-2.4 vs -1.2, p


332A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

The following authors have nothing to disclose: Kellie Young, Maria Aguilar, Taft<br />

Bhuket, Benny Liu, Robert J. Wong<br />

238<br />

Application of transient elastography and FibroMeter in<br />

patients with nonalcoholic fatty liver disease<br />

Vincent W. Wong 1 , Grace LH Wong 1 , Sally Shu 1 , Angel ML<br />

Chim 1 , Anthony W. Chan 2 , Paul C. Choi 2 , Henry Lik-Yuen Chan 1 ;<br />

1 Department of Medicine and Therapeutics, The Chinese University<br />

of Hong Kong, Hong Kong, China; 2 Department of Anatomical<br />

and Cellular Pathology, The Chinese University of Hong Kong,<br />

Hong Kong, China<br />

Background: Nonalcoholic fatty liver disease (NAFLD) has<br />

emerged as a significant cause of cirrhosis and hepatocellular<br />

carcinoma, but the majority of patients have relatively minor<br />

disease. Hence, initial non-invasive assessments are preferred.<br />

Liver stiffness measurement (LSM) by transient elastography<br />

may fail in obese NAFLD patients. We therefore evaluated<br />

the use of LSM and serum-based FibroMeter tests in NAFLD<br />

patients. Methods: Consecutive liver biopsies for NAFLD were<br />

evaluated using the Kleiner’s system. LSM and blood tests were<br />

performed one day prior to liver biopsy. LSM was considered<br />

reliable if 10 valid acquisitions were obtained and the interquartile<br />

range-to-median ratio of LSM was ≤0.3. Cutoff values<br />

for LSM (7.9 and 9.6 kPa for ≥F3 disease) and FibroMeter-NA-<br />

FLD (0.62 for ≥F3 disease) were based on prior publications.<br />

Results: 293 examinations were analyzed (age 50 years [SD<br />

10], 54% males, body mass index 27.7 kg/m 2 [SD 4.2]).<br />

The distribution of fibrosis stages was 38% F0, 30% F1, 9%<br />

F2, 11% F3 and 12% F4. FibroMeter score was valid in all<br />

patients. Among 233 patients undergoing LSM, 224 (96%)<br />

had 10 valid acquisitions and 203 (87%) had reliable measurements.<br />

In the overall population, the area under the receiver-operating<br />

characteristics curve (AUROC) for FibroMeter was<br />

0.72 (95% CI 0.66-0.78) for ≥F2, 0.73 (95% CI 0.67-0.80)<br />

for ≥F3 and 0.79 (95% CI 0.72-0.87) for F4. At the cutoff of<br />

0.62, the sensitivity, specificity, positive and negative predictive<br />

values of FibroMeter for ≥F3 disease were 25%, 91%,<br />

46% and 81%, respectively. In patients with reliable LSM, the<br />

corresponding AUROC was 0.83 (95% CI 0.77-0.89), 0.91<br />

(95% CI 0.87-0.95) and 0.91 (95% CI 0.86-0.97), respectively.<br />

The LSM cutoff of 7.9 kPa had 93% sensitivity and 98%<br />

negative predictive value in excluding ≥F3 disease. On the<br />

other hand, the cutoff of 9.6 kPa had 86% specificity and<br />

only 60% positive predictive value in ruling in ≥F3 disease. If<br />

FibroMeter was performed in 113 patients with unreliable LSM<br />

or LSM ≥7.9 kPa, 61 of 98 (62%) patients with a score


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 333A<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

Manal F. Abdelmalek - Consulting: Islet Sciences; Grant/Research Support:<br />

Tobira, Gilead Sciences, NIH/NIDDK, Synageva, Genfit Pharmaceuticals,<br />

Immuron, Galmed, TaiwanJ Pharma, Intercept, NGM Pharmaceuticals<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Brent A. Neuschwander-Tetri - Advisory Committees or Review Panels: Nimbus<br />

Therapeutics, Bristol Myers Squibb, Janssen, Mitsubishi Tanabe, Conatus,<br />

Scholar Rock<br />

Reshma Shringarpure - Employment: Intercept<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: Arthur J. McCullough, Saswati<br />

Hazra, Xiaohong Yan, Leigh MacConell<br />

240<br />

Omega-3 fatty acids improve proteomic and lipidomic<br />

markers of endoplasmic reticulum (ER) stress and mitochondrial<br />

dysfunction in a randomized controlled trial in<br />

subjects with Nonalcoholic Steatohepatitis<br />

Livia Rodrigues 1 , Claudia P. Oliveira 1 , José Tadeu Stefano 1 ,<br />

Monize A. Nogueira 1 , Ismael D. Silva 2 , Edson G. Lo Turco 3 ,<br />

Venancio Avancini F. Alves 4 , Flair J. Carrilho 1 , Puneet Puri 5 , Dan<br />

Waitzberg 1 ; 1 Gastroenterology, University of São Paulo School of<br />

Medicine, São Paulo, Brazil; 2 Gynecology, Universidade Federal<br />

de São Paulo, São Paulo, Brazil; 3 Surgery/Urology, Universidade<br />

Federal de São Paulo, São Paulo, Brazil; 4 Patology, University of<br />

São Paulo School of Medicine, São Paulo, Brazil; 5 Gastroenterology,<br />

Hepatology and Nutrition, Virginia Commonwealth University-VCU,<br />

Richmond, VA<br />

Background/Aims: There is no effective FDA approved treatment<br />

for nonalcoholic steatohepatitis (NASH). Increased<br />

n-6/n-3 polyunsaturated fatty acid (PUFA) ratio can induce<br />

endoplasmic reticulum (ER) stress and mitochondrial dysfunction<br />

that characterize NASH. We hypothesized that n-3 PUFA supplementation<br />

would improve these abnormalities. We aimed to<br />

define the hepatic proteomic and plasma lipidomic profiles following<br />

n-3 PUFA therapy in subjects with NASH and relate it to<br />

markers of ER stress and mitochondrial dysfunction. Methods:<br />

A 6-month double blind randomized controlled trial in subjects<br />

with biopsy proven NASH was conducted with n-3 PUFA (945<br />

mg [α linolenic acid/ 64%, eicosapentaenoic acid (EPA)/16%<br />

and docosahexaenoic acid (DHA)/21%]), 3 capsules/day and<br />

matched placebo. A 6-month follow up liver biopsy was performed<br />

per IRB protocol. Hepatic proteomics from formalin-fixed<br />

paraffin embedded liver tissue and plasma lipidomics were<br />

performed at baseline and 6-months in the n-3 PUFA group<br />

using mass spectrometry. The proteins and lipids were matched<br />

to UniProt and LIPID MAPS database respectively. Cytoscape<br />

software was used to analyze functional pathways. Results:<br />

Age and gender matched 60 NASH subjects (n=32, n-3 PUFA<br />

treatment group; n=28, placebo group) completed the study. A<br />

6-month n-3 PUFA therapy significantly (all p 2x baseline and > 5x<br />

ULN, occurring within the first 24 weeks of treatment were<br />

evaluated in context of subsequent clinical outcomes through<br />

Week 48. Results: Overall, 29/172, 25/170, 3/174 and<br />

25/163 patients, respectively, from Arms A-D met criteria for<br />

ALT flare; of those, 18, 16, 2, and 14 patients were baseline<br />

HBeAg-positive. A greater proportion of patients in Arms<br />

A, B and D who experienced ALT flares achieved subsequent<br />

HBeAg loss, HBsAg decline ≥ 1log 10<br />

IU/ml, or HBsAg loss<br />

compared to those who did not experience an ALT flare (table).


334A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Comparing two combination therapies, Arm A achieved higher<br />

rates of HBsAg loss (P=0.016) following flares than Arm B.<br />

This is partly driven by HBeAg-negative patients in Arm A who<br />

attained disproportionately increasing rates of HBsAg loss and<br />

HBsAg decline ≥ 1log 10<br />

IU/ml which was not observed in<br />

Arm B. Conclusion: Treatment-associated ALT flares, especially<br />

with TDF+PEG combination therapy, are associated with clinical<br />

endpoints related to HBV immune control. ALT flares were<br />

higher in the TDF+PEG x48w group and associated with the<br />

highest HBsAg loss.<br />

Table: Proportion of Patients who Achieved Clinical Endpoints by<br />

ALT flare status (N=679)<br />

*Includes subjects with baseline HBeAg-positive only.<br />

Disclosures:<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

Phillip Dinh - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Eduardo B. Martins - Employment: Gilead Sciences, Inc.; Stock Shareholder:<br />

Gilead Sciences, Inc.<br />

Leland J. Yee - Employment: Gilead Science<br />

Prista Charuworn - Stock Shareholder: Gilead Sciences<br />

Mani Subramanian - Employment: Gilead Sciences<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Seng Gee Lim - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals, Gilead<br />

Pharmaceuticals, Roche Pharmaceuticals; Speaking and Teaching: Bristol-Myers<br />

Squibb, Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals,<br />

Gilead Pharmaceuticals<br />

Scott Fung - Advisory Committees or Review Panels: Gilead Sciences, Merck,<br />

Vertex, Roche; Speaking and Teaching: BMS, Gilead, AbbVie, Janssen, Merck<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

Giovanni B. Gaeta - Advisory Committees or Review Panels: Janssen, Merck,<br />

Abbvie, Roche; Speaking and Teaching: BMS, Gilead<br />

George V. Papatheodoridis - Advisory Committees or Review Panels: Merck<br />

Sharp & Dohme, Novartis, Abbvie, Boerhinger Ingelheim, Bristol-Meyer Squibb,<br />

Gilead, Roche, Janssen, GlaxoSmith Kleine; Grant/Research Support: Roche,<br />

Gilead, Bristol-Meyer Squibb, Abbvie, Janssen; Speaking and Teaching: Merck<br />

Sharp & Dohme, Bristol-Meyer Squibb, Gilead, Roche, Janssen, Abbvie<br />

Robert Flisiak - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, Janssen, Merck, Roche, Abbvie; Consulting: Janssen, Bristol Myers<br />

Squibb, Gilead, Merck, Abbvie; Grant/Research Support: Roche; Speaking and<br />

Teaching: Janssen, Roche, Bristol Myers Squibb, Gilead, Merck, Abbvie<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

The following authors have nothing to disclose: Sang Hoon Ahn, Fehmi Tabak,<br />

Rajiv Mehta, Kyungpil Kim, Aric Josun Hui<br />

242<br />

Nucleos(t)ide Analogues against Hepatitis B Virus<br />

Reduces the Risk of Various Major Malignancies: a<br />

Nationwide Cohort Study Based on the Health Insurance<br />

Review and Assessment Service Database.<br />

Donghyeon Lee 1 , Jeong-Hoon Lee 1 , Yuri Cho 1 , Jeong-Ju Yoo 1 ,<br />

Minjong Lee 1 , Young Youn Cho 1 , Hyeki Cho 1 , Hongkeun Ahn 1 ,<br />

June Sung Lee 2 , Eun Ju Cho 1 , Su Jong Yu 1 , Yoon Jun Kim 1 , Jung-<br />

Hwan Yoon 1 ; 1 Seoul national university hospital, Seoul, Korea (the<br />

Republic of); 2 Inje University, Ilsan Paik Hospital, Ilsan, Korea (the<br />

Republic of)<br />

Background: Preceding epidemiology <strong>studies</strong> reported that<br />

chronic hepatitis B virus (HBV) infection is a risk factor for various<br />

malignancies including hepatocellular carcinoma (HCC),<br />

cholangiocarcinoma, non-Hodgkin lymphoma, and stomach<br />

cancer. These associations have been partially explained by<br />

the HBV-inducing immune tolerance to malignant cells. It has<br />

been well proven that antiviral treatment in chronic hepatitis<br />

B (CHB) patients reduces the risk of HCC. However, it is still<br />

unclear whether nucleos(t)ide analogues (NAs) against HBV<br />

could reduce the risk of various malignancies except for HCC.<br />

In this study, we evaluated the association of anti-HBV treatment<br />

with NAs and the risk of various cancers in CHB patients.<br />

Methods: The National Claims Database from Health Insurance<br />

Review and Assessment Service (HIRA) was used. From<br />

this database, we included CHB patients who received anti-<br />

HBV NAs (i.e., lamivudine, telvibudine, clevudine, adefovir,<br />

entecavir, and tenofovir) for > 30 days between 2009 and<br />

2013 (the NA group) or no antiviral treatment (the control<br />

group). Date of data cutoff was December 31, 2014. The<br />

hazard ratio (HR) for NAs was calculated by Cox regression<br />

model. Results: A total of 299,740 patients with CHB were<br />

included and 39,900 patients (13.3%) were involved in the<br />

NA group. During the study period 14,186 patients developed<br />

any kind of malignancy. In multivariate analysis, the NA<br />

group showed significantly lower risk of developing new malignancies<br />

(vs the control group; HR=0.511, 95% confidence<br />

interval=0.466-0.561, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 335A<br />

243<br />

Biochemical, Virologic, and Serologic Response Following<br />

Discontinuation of Long Term TDF Therapy<br />

Maria Buti 1 , David K. Wong 2 , Edward J. Gane 3 , Robert Flisiak 4 ,<br />

Michael P. Manns 5 , Kelly D. Kaita 6 , Lanjia Lin 7 , Kathryn M. Kitrinos<br />

7 , John F. Flaherty 7 , Anuj Gaggar 7 , Patrick Marcellin 8 , Harry<br />

L. Janssen 9 ; 1 Hospital General Universitari Vall d’Hebron and<br />

Ciberehd, Barcelona, Spain; 2 University of Toronto, Toronto, ON,<br />

Canada; 3 Auckland City Hospital, Auckland, New Zealand; 4 Medical<br />

University of Bialystok, Bialystok, Poland; 5 Medical School of<br />

Hannover, Hannover, Germany; 6 University of Manitoba, Winnipeg,<br />

MB, Canada; 7 Gilead Sciences, Inc., Foster City, CA; 8 Hôpital<br />

Beaujon, Clichy, France; 9 ErasmusMC, Rotterdam, Netherlands<br />

Introduction: Treatment of HBV with tenofovir disoproxil fumarate<br />

(TDF) is associated with durable viral suppression, liver<br />

fibrosis improvement, and no documented resistance in a<br />

cohort of patients followed for up to 8 years. We characterized<br />

the responses following treatment discontinuation. Methods:<br />

Patients from 2 phase 3 <strong>studies</strong> of TDF in HBeAg-negative and<br />

HBeAg-positive patients were followed for up to 24 weeks after<br />

discontinuing TDF. Clinical and laboratory parameters, and<br />

pre-specified requirements to restart therapy were monitored<br />

during the treatment-free follow-up (TFFU) period. Results: To<br />

date, 123 patients entered TFFU with mean duration of 12<br />

weeks and 71 patients completing 24 weeks of TFFU. 98 (80%)<br />

patients were HBeAg negative at the time of TDF withdrawal,<br />

including 21 that seroconverted in the parent study. Serious<br />

adverse events (AEs) were reported in 10 subjects (8%), 7<br />

of which involved ALT elevation, and Grade 3-4 AEs were<br />

observed in 12% of subjects. 89% of patients had increase in<br />

HBV DNA to >69 IU/mL and 28% had Grade 3-4 elevations in<br />

ALT after cessation of TDF. At last TFFU visit, 44 patients (36%)<br />

had normal ALT, 59 (48%) maintained HBV DNA


336A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

were younger (49.9 vs. 537; P90%.<br />

Conclusions: The results demonstrate that serum HBV RNA levels<br />

could be used early during PegIFNα-2a therapy, and that<br />

they appear superior to conventional biomarkers for predicting<br />

non-response in HBeAg-positive patients.<br />

Biomarker Cutoff Performance Characteristics at Treatment Week<br />

12<br />

Disclosures:<br />

Florian van Bömmel - Advisory Committees or Review Panels: Gilead Sciences<br />

Inc., Roche Pharmaceutics; Grant/Research Support: Biopredictive; Speaking<br />

and Teaching: Bristol-Myers Squibb, Janssen, Roche Pharmaceutics, Gilead Sciences<br />

Inc., Abbvie<br />

Hua He - Employment: Roche<br />

Cynthia Wat - Employment: Roche Products Ltd<br />

Vedran Pavlovic - Employment: Roche Products Ltd; Stock Shareholder: Roche<br />

Products Ltd<br />

Lei Yang - Employment: Roche (China) Holding Ltd.<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

The following authors have nothing to disclose: Alena van Bömmel, Alexander<br />

Krauel, Danilo Deichsel, Stephan Boehm<br />

246<br />

The Predictive Values of Hepatocellular Carcinoma Risk<br />

Scores are not Affected from the Antiviral Therapy in<br />

Patients with Chronic Hepatitis B Infection<br />

Ulus S. Akarca, Fulya Gunsar, Ilker Turan, Galip Ersoz, Zeki<br />

Karasu, Omer Ozutemiz; Division of Gastroenterology, Ege University<br />

Faculty of Medicine, Izmir, Turkey<br />

Background and Aims: Hepatocellular carcinoma (HCC)<br />

remains the major complication of chronic hepatitis B (CHB).<br />

Several scoring systems including CU-HCC, GAG-HCC, and<br />

REACH-B have been developed to assess the risk for development<br />

of HCC. In addition to the patients who do not receive<br />

antiviral treatment, the predictive value of these scores for HCC<br />

was found to be good in patients on treatment. In the present<br />

study, we aimed to assess the impact of antiviral therapy for the<br />

predictive values of HCC risk scores in Caucasian CHB patients<br />

and compare the predictive values in patients on treatment and<br />

without treatment. Methods: We investigated the predictive<br />

values of CU-HCC, GAG-HCC, and REACH-B scores in Caucasian<br />

HBsAg-positive patients who had baseline labs including<br />

HBV DNA at presentation. Patients who had (1) HCC at presentation;<br />

(2) HCV, HDV and HIV coinfection; (3) missing data<br />

for the calculation of the scores were excluded. Results: We<br />

analyzed 2331 patients (M/F:1370/961, mean age 42±13).<br />

In this study group, 623 patients were inactive HBV carrier,<br />

1454 patients were CHB, 163 patients had cirrhosis and 91<br />

patients had HBsAg seroconversion during follow-up. HCC<br />

developed in 37 patients during median 4.9 year follow-up.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 337A<br />

Three hundred forty-one patients were on lamivudine, 60 were<br />

on adefovir, 188 were on entecavir, 467 were on tenofovir,<br />

and 32 were on combination therapies. ROC analysis for the<br />

predictive values of HCC scores showed that AUROC values<br />

were 0.804 for CU-HCC, 0.809 for REACH-B and 0.882 for<br />

GAG-HCC. These values were similar in patients without therapy<br />

and with entecavir and tenofovir therapies as in the literature.<br />

The most predictive GAG-HCC score had sensitivity of<br />

97.3% and negative predictive value of 99.9%. Conclusion:<br />

In this large cohort of Caucasian HBV patients, all of these<br />

scores had good predictive values for HCC development both<br />

in patients on treatment and in patients who did not receive<br />

antiviral treatment.<br />

Disclosures:<br />

Ulus S. Akarca - Advisory Committees or Review Panels: GILEAD, BMS, MSD,<br />

AbbVie<br />

The following authors have nothing to disclose: Fulya Gunsar, Ilker Turan, Galip<br />

Ersoz, Zeki Karasu, Omer Ozutemiz<br />

247<br />

High Rates of SVR in Treatment-Experienced Patients<br />

with Genotype 1 HCV Infection and Cirrhosis After<br />

Treatment with Ledipasvir/Sofosbuvir and Vedroprevir<br />

with or Without Ribavirin for 8 weeks<br />

Eric Lawitz 1 , Fred Poordad 1 , Robert H. Hyland 2 , Lin Liu 2 , Hadas S.<br />

Dvory 2 , Phillip S. Pang 2 , Diana M. Brainard 2 , Julio A. Gutierrez 1 ;<br />

1 Texas Liver Institute, San Antonio, TX; 2 Gilead Sciences, Inc, Foster<br />

City, CA<br />

Background: Ledipasvir/sofosbuvir (LDV/SOF) fixed dose combination<br />

± ribavirin (RBV) for 12 weeks in patients with HCV<br />

genotype 1 and cirrhosis is safe and effective. We previously<br />

showed that LDV/SOF plus the non-nucleotide NS5B inhibitor<br />

GS-9669 for 8 weeks resulted in 82-91% SVR12 in patients<br />

with cirrhosis and genotype 1 HCV infection. In the current<br />

study, we evaluated the safety and efficacy of 8 weeks of LDV/<br />

SOF plus the HCV NS3 protease inhibitor vedroprevir (VDV,<br />

GS-9451) ± RBV in genotype 1 HCV-infected patients with<br />

cirrhosis. Methods: Patients, previously treated with pegylated<br />

interferon+RBV, with compensated cirrhosis and chronic HCV<br />

genotype 1 infection were randomized equally to receive LDV/<br />

SOF+VDV or LDV/SOF+VDV+RBV for 8 weeks. Deep sequencing<br />

was performed for all patients who didn’t achieved SVR12.<br />

Results: 46 patients were randomized and treated. The majority<br />

were male (65%), Caucasian (96%), and had HCV genotype<br />

1a (70%), and were IL28B non-CC (87%). The median<br />

HCV RNA was 6.0 log 10<br />

IU/ml. Rates of on-treatment Week 4<br />

viral suppression, SVR12, and relapse are shown in the table.<br />

NS5A RAVs were detected in all 3 subjects who relapsed.<br />

LDV/SOF+VDV±RBV was safe and well tolerated. There were<br />

no SAEs. One patient was lost to follow up having completed<br />

only 4 weeks of therapy. There were no discontinuations due<br />

to adverse events. Adverse events were generally mild, and<br />

Grade 3/4 laboratory abnormalities were infrequent. The most<br />

frequent adverse events were headache, rash, and sinusitis.<br />

Conclusions: Ledipasvir/sofosbuvir +VDV for 8 weeks achieved<br />

high SVR rates in genotype 1 HCV-infected treatment-experienced<br />

patients with cirrhosis irrespective of the addition of RBV.<br />

Both regimens were safe and well tolerated.<br />

Disclosures:<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Robert H. Hyland - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Lin Liu - Employment: Gilead Sciences, Inc.<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

The following authors have nothing to disclose: Hadas S. Dvory, Julio A. Gutierrez<br />

248<br />

High SVR4 Rates Achieved With the Next Generation<br />

NS3/4A Protease Inhibitor ABT-493 and NS5A Inhibitor<br />

ABT-530 in Non-Cirrhotic Treatment-Naïve and Treatment-Experienced<br />

Patients With HCV Genotype 3 Infection<br />

(SURVEYOR-2)<br />

Paul Y. Kwo 1 , Michael Bennett 2,3 , Stanley Wang 4 , Hugo E.<br />

Vargas 5 , David L. Wyles 6 , J. Scott Overcash 7 , Peter J. Ruane 8 ,<br />

Benedict Maliakkal 9 , Asma Siddique 10 , Bal R. Bhandari 11 , Fred<br />

Poordad 12 , Ran Liu 4 , Chih-Wei Lin 4 , Teresa Ng 4 , Federico J.<br />

Mensa 4 , Jens Kort 4 ; 1 Division of Gastroenterology, Indiana University<br />

Medical Center, Indianapolis, IN; 2 San Diego Digestive<br />

Disease Consultants, Inc., San Diego, CA; 3 Medical Associates<br />

Research Group, Inc., San Diego, CA; 4 AbbVie Inc., North Chicago,<br />

IL; 5 Mayo Clinic, Phoenix, AZ; 6 University of California San<br />

Diego, La Jolla, CA; 7 eStudySite, San Diego, CA; 8 Ruane Medical<br />

& Liver Health Institute, Los Angeles, CA; 9 University of Rochester<br />

Medical Center, Rochester, NY; 10 Virginia Mason Hospital and<br />

Medical Center, Seattle, WA; 11 Gastroenterology & Nutritional<br />

Medical Services, Bastrop, LA; 12 Texas Liver Institute, University of<br />

Texas Health Science Center, San Antonio, TX<br />

Background: The next generation HCV direct acting antivirals<br />

(DAAs) ABT-493, an NS3/4A protease inhibitor identified by<br />

AbbVie and Enanta, and ABT-530, an NS5A inhibitor, are<br />

characterized by potent pan-genotypic in vitro antiviral activity<br />

and a high barrier to resistance selection with retention<br />

of activity against key known resistance-associated variants.<br />

ABT-493 and ABT-530 each provided a mean 4 log 10<br />

IU/mL<br />

decline from baseline in HCV plasma viral load after 3 day<br />

monotherapy in genotype (GT) 1-infected subjects and have<br />

comparable in vitro antiviral potency against GT3a. Purpose:<br />

To evaluate the efficacy and safety of ABT-493 and ABT-530<br />

with or without ribavirin (RBV) in non-cirrhotic GT3-infected<br />

treatment-naïve (TN) and pegylated interferon/RBV (pegIFN/<br />

RBV) treatment experienced (TE) subjects Methods: Subjects<br />

received 12 weeks of ABT-493 300 mg+ABT-530 120 mg<br />

(Arm D), ABT-493 200 mg+ABT-530 120 mg (Arm E), ABT-<br />

493 200 mg+ABT-530 120 mg+RBV (Arm F), or ABT-493 200<br />

mg+ABT-530 40 mg (Arm G). DAAs were dosed once daily;<br />

weight-based RBV (1000 or 1200 mg) was dosed twice daily.<br />

The primary efficacy endpoint is sustained virologic response<br />

12 weeks after the last dose of study drug (SVR12); we present<br />

post-treatment week 4 data (SVR4) here. Safety was evaluated<br />

by monitoring adverse events (AEs), laboratory tests, and vital<br />

signs. Results: 120 GT3-infected subjects were treated in Arms


338A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

D (n=30), E (n=29), F (n=31), or G (n=30). Subjects were<br />

male, 56%; median age, 52.0 years; GT3a, 98%; TN, 92%;<br />

TE, 8%; fibrosis >F2, 15%; median baseline HCV RNA log 10<br />

IU/mL, 6.7. There has been 1 virologic failure thus far in each<br />

treatment arm (n=4), 3 of which were in pegIFN/RBV TE subjects.<br />

One subject in Arm G was lost to follow-up at the week<br />

2 visit. The SVR4 rate was 96% (27/28), 96% (27/28), 97%<br />

(29/30) and 93% (27/29) in Arms D, E, F, and G, respectively,<br />

among subjects with available post-treatment Week 4<br />

data. AEs were mostly mild, with most common DAA-related<br />

AEs being fatigue, nausea, and headache. There were no serious<br />

DAA-related AEs; 1 subject discontinued due to DAA- and<br />

RBV-related AEs of abdominal pain and heat sensation. Typical<br />

reductions in hemoglobin were observed in the RBV containing<br />

arm (Arm F). Conclusions: ABT-493+ABT-530 treatment for<br />

12 weeks with or without RBV in TN or TE noncirrhotic HCV<br />

GT3-infected subjects was well tolerated. Promising SVR4 rates<br />

of 93%–96% were achieved without RBV. SVR12 data and<br />

further details regarding virologic failures will be available<br />

at the time of the meeting. These encouraging results warrant<br />

further study of this once-daily ABT-493/ABT-530 regimen in<br />

GT3 patients, including patients with cirrhosis.<br />

Disclosures:<br />

Paul Y. Kwo - Advisory Committees or Review Panels: Abbvie, Abbott, Novartis,<br />

Merck, Gilead, BMS, Janssen; Consulting: BMS; Grant/Research Support:<br />

Roche, Abbvie, Merck, BMS, Abbott, Idenix, Vital Therapeutics, Gilead, Vertex,<br />

Merck, Idenix; Speaking and Teaching: Merck, Merck<br />

Stanley Wang - Employment: AbbVie<br />

David L. Wyles - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Janssen, Merck; Grant/Research Support: Gilead, Merck, Bristol Myers Squibb,<br />

AbbVie, Tacere<br />

Peter J. Ruane - Advisory Committees or Review Panels: BMS; Consulting: Gilead,<br />

Abbvie, Janssen; Grant/Research Support: BMS, Gilead, Merck, Abbvie, Idenix,<br />

Idenix, Janssen, Viiv; Speaking and Teaching: Gilead, Merck, Abbvie, Abbvie,<br />

Janssen; Stock Shareholder: Gilead, Gilead<br />

Benedict Maliakkal - Speaking and Teaching: Valeant Pharma (Salix), Gilead,<br />

AbbVie, Bristol Myers Squibb<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Ran Liu - Employment: Abbvie<br />

Chih-Wei Lin - Employment: Abbvie<br />

Teresa Ng - Employment: AbbVie; Patent Held/Filed: AbbVie; Stock Shareholder:<br />

AbbVie<br />

Federico J. Mensa - Employment: Abbvie Inc.; Stock Shareholder: Abbvie Inc.<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

The following authors have nothing to disclose: Michael Bennett, Hugo E. Vargas,<br />

J. Scott Overcash, Asma Siddique, Bal R. Bhandari<br />

249<br />

Sofosbuvir/GS-5816 Fixed Dose Combination for 12<br />

Weeks Compared to Sofosbuvir with Ribavirin for 24<br />

Weeks in Genotype 3 HCV Infected Patients: The Randomized<br />

Controlled Phase 3 ASTRAL-3 Study<br />

Alessandra Mangia 1 , Stuart K. Roberts 2 , Stephen Pianko 3 , Alex<br />

J. Thompson 4 , Curtis Cooper 5 , Brian Conway 6 , Marc Bourlière<br />

7 , Tarik Asselah 8 , Thomas Berg 9 , Stefan Zeuzem 10 , William<br />

M. Rosenberg 11 , Kosh Agarwal 12 , Edward J. Gane 13 , Catherine<br />

A. Stedman 14 , Francesco Mazzotta 15 , Tram T. Tran 16 , Stuart<br />

C. Gordon 17 , Evguenia S. Svarovskaia 18 , Lingling Han 18 ,<br />

John McNally 18 , Anu Osinusi 18 , Diana M. Brainard 18 , John G.<br />

McHutchison 18 , Nezam H. Afdhal 19 , Graham R. Foster 20 ; 1 Casa<br />

Sollievo della Sofferenza Hospital, San Giovanni Rotondo, Italy;<br />

2 Alfred Health and Monash University, Melbourne, VIC, Australia;<br />

3 Monash Health and Monash University, Melbourne, VIC, Australia;<br />

4 St Vincent’s Hospital, Melbourne, VIC, Australia; 5 Ottawa<br />

General Hospital, Ottawa, ON, Canada; 6 Vancouver Infectious<br />

Diseases Centre, Vancouver, BC, Canada; 7 Hôpital Saint Joseph,<br />

Marseilles, France; 8 University Paris Diderot, INSERM U773, Paris,<br />

France; 9 University Hospital Leipzig, Leipzig, Germany; 10 Johann<br />

Wolfgang Goethe University, Frankfurt, Germany; 11 University<br />

College London, London, United Kingdom; 12 Kings College<br />

Hospital, London, United Kingdom; 13 Auckland Clinical Studies,<br />

Auckland, New Zealand; 14 Christchurch Clinical Studies Trust and<br />

University of Otago, Christchurch, New Zealand; 15 Santa Maria<br />

Annunziata Hospital, Florence, Italy; 16 Cedars-Sinai Medical Center,<br />

Los Angeles, CA; 17 Henry Ford Health System, Detroit, MI;<br />

18 Gilead Sciences, Inc., Foster City, CA; 19 Beth Israel Deaconess<br />

Medical Center, Boston, MA; 20 Queen Mary University of London,<br />

London, United Kingdom<br />

Introduction: GS-5816 is a pangenotypic HCV NS5A inhibitor.<br />

In Phase 2 <strong>studies</strong>, the combination of sofosbuvir (SOF) and<br />

GS-5816 for 12 weeks demonstrated high efficacy in patients<br />

with genotype 3 HCV. This Phase 3 study compared treatment<br />

with a fixed dose combination (FDC) of SOF/GS-5816 for<br />

12 weeks to standard of care, SOF+RBV for 24 weeks, in<br />

patients with genotype 3 HCV. Methods: Patients at 75 sites<br />

in North America, Europe, Australia and New Zealand were<br />

randomized 1:1 to received SOF/GS-5816 (400 mg /100<br />

mg daily) FDC for 12 weeks or SOF (400mg daily) with RBV<br />

(1000-1200mg daily) for 24 weeks. HCV RNA was measured<br />

with the CAP/CTM HCV 2.0 assay with LLOQ =15 IU/mL The<br />

primary endpoint was sustained virologic response 12 weeks<br />

after treatment (SVR12) with a pre-specified non-inferiority margin<br />

of 10%. Results: Of 552 patients randomized and treated,<br />

62% were male, 89% were white, 39% had IL28B CC genotype,<br />

26% had prior treatment failure, and 30% had cirrhosis.<br />

Eight patients discontinued treatment due to adverse events,<br />

all of which were in the SOF +RBV treatment group. The most<br />

common AEs (>10% in either treatment group) are presented<br />

below. Five patients in the SOF/GS-5816 treatment group<br />

and 8 patients in the SOF+RBV treatment group experienced<br />

SAEs. One SAE, acute generalized exanthematous pustulosis,<br />

in a SOF+RBV treated patient was assessed as related to study<br />

drugs by the investigator. Hemoglobin decline and total bilirubin<br />

increases were more commonly observed in the group<br />

treated with SOF +RBV consistent with RBV–induced hemolysis.<br />

No other significant lab abnormalities were observed. HCV<br />

RNA declined rapidly in both treatment groups with 92% and<br />

88% of patients achieving HCV RNA < LLOQ after 4 weeks<br />

of treatment in the SOF/GS-5816 and SOF+RBV treatment<br />

groups, respectively. SVR12 data for all patients will be presented.<br />

Conclusions: The once daily, all-oral, single tablet regimen<br />

of SOF/GS-5816 was well tolerated in treatment-naïve<br />

and treatment-experienced genotype 3 HCV-infected patients


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 339A<br />

with and without cirrhosis. There were no discontinuations due<br />

to adverse events and a lower incidence of fatigue, insomnia<br />

and irritability in patients treated with SOF/GS-5816 for<br />

12 weeks compared to patients treated with SOF+RBV for 24<br />

weeks.<br />

Nezam H. Afdhal - Advisory Committees or Review Panels: Trio Helath Care;<br />

Board Membership: Journal Viral hepatitis; Consulting: Merck, EchoSens, BMS,<br />

Achillion, GlaxoSmithKline, Springbank, Gilead, AbbVie; Grant/Research Support:<br />

Gilead; Stock Shareholder: Springbank<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

The following authors have nothing to disclose: Francesco Mazzotta, Lingling<br />

Han<br />

Disclosures:<br />

Alessandra Mangia - Advisory Committees or Review Panels: ROCHE, Janssen,<br />

MSD, ROCHE, Janssen, MSD, Boheringer ; Consulting: Gilead; Grant/Research<br />

Support: Shering-Plough, Shering-Plough<br />

Stuart K. Roberts - Board Membership: AbbVie, Gilead<br />

Stephen Pianko - Advisory Committees or Review Panels: Roche, Novartis, GIL-<br />

EAD, Roche, Novartis; Consulting: GILEAD; Speaking and Teaching: JANSSEN<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

Curtis Cooper - Advisory Committees or Review Panels: Gilead, Abbvie, MERCK;<br />

Grant/Research Support: MERCK, Gilead, Abbvie; Speaking and Teaching:<br />

MERCK, Abbvie, Gilead<br />

Brian Conway - Advisory Committees or Review Panels: Vertex Pharmaceuticals,<br />

Merck, Boehringer Ingelheim, Jannsen Pharmaceuticals; Grant/Research Support:<br />

Vertex Pharmaceuticals, Merck, Boehringer Ingelheim, Jannsen Pharmaceuticals,<br />

AbbVie, Gilead Sciences, Gilead Sciences<br />

Marc Bourlière - Advisory Committees or Review Panels: Schering-Plough,<br />

Bohringer inghelmein, Schering-Plough, Bohringer inghelmein, Transgene; Board<br />

Membership: Bristol-Myers Squibb, Gilead, Idenix; Consulting: Roche, Novartis,<br />

Tibotec, Abott, glaxo smith kline, Merck, Bristol-Myers Squibb, Novartis, Tibotec,<br />

Abott, glaxo smith kline; Speaking and Teaching: Gilead, Roche, Merck, Bristol-Myers<br />

Squibb<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

William M. Rosenberg - Advisory Committees or Review Panels: Janssen, Merk,<br />

Gilead, Merk, Gilead, GSK; Board Membership: iQur Limited, iQur Limited;<br />

Consulting: siemens; Speaking and Teaching: siemens, Roche<br />

Kosh Agarwal - Advisory Committees or Review Panels: Gilead, BMS, Novartis,<br />

Janssen, AbbVie, Gilead; Consulting: MSD, Janssen; Grant/Research Support:<br />

Roche, Gilead, BMS, BMS; Speaking and Teaching: Astellas<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Catherine A. Stedman - Advisory Committees or Review Panels: MSD, Abbvie;<br />

Speaking and Teaching: Gilead, Abbvie<br />

Tram T. Tran - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb; Consulting: Gilead, AbbVie, Janssen; Grant/Research Support: Bristol<br />

Myers Squibb; Speaking and Teaching: Gilead<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

Evguenia S. Svarovskaia - Employment: Gilead Sciences Inc; Stock Shareholder:<br />

Gilead Sciences Inc<br />

John McNally - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Anu Osinusi - Employment: gilead sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

250<br />

High SVR4 Rates Achieved With the Next Generation<br />

NS3/4A Protease Inhibitor ABT-493 and NS5A Inhibitor<br />

ABT-530 in Non-Cirrhotic Treatment-Naïve and Treatment-Experienced<br />

Patients With HCV Genotype 2 Infection<br />

(SURVEYOR-2)<br />

David L. Wyles 2 , Mark S. Sulkowski 3 , Stanley Wang 1 , Michael<br />

Bennett 4 , Hugo E. Vargas 5 , J. Scott Overcash 6 , Benedict Maliakkal<br />

7 , Asma Siddique 8 , Bal R. Bhandari 9 , Fred Poordad 10 , Sandra<br />

S. Lovell 1 , Chih-Wei Lin 1 , Teresa Ng 1 , Federico J. Mensa 1 , Jens<br />

Kort 1 ; 1 AbbVie, Inc., North Chicago, IL; 2 University of California<br />

San Diego, La Jolla, CA; 3 Johns Hopkins University School of Medicine,<br />

Baltimore, MD; 4 San Diego Digestive Disease Consultants,<br />

Inc., and Medical Associates Research Group, Inc., San Diego,<br />

CA; 5 Mayo Clinic, Pheonix, AZ; 6 eStudySite, San Diego, CA;<br />

7 University of Rochester Medical Center, Rochester, NY; 8 Virginia<br />

Mason Hospital and Medical Center, Seattle, WA; 9 Gastroenterology<br />

& Nutritional Medical Services, Bastrop, LA; 10 Texas Liver Institute,<br />

University of Texas Health Science Center, San Antonio, TX<br />

Background: The next-generation HCV direct acting antivirals<br />

(DAAs) ABT-493, an NS3/4A protease inhibitor identified by<br />

AbbVie and Enanta, and ABT-530, an NS5A inhibitor, are<br />

characterized by potent pan-genotypic in vitro antiviral activity<br />

and a high barrier to resistance selection with retention of<br />

activity against key resistance-associated variants. ABT-493<br />

and ABT-530 each provided a mean 4 log 10<br />

IU/mL decline<br />

from baseline in HCV plasma viral load after 3 day monotherapy<br />

in genotype (GT) 1-infected subjects and have comparable<br />

in vitro antiviral potency against GT2. Purpose: To<br />

evaluate the efficacy and safety of ABT-493 and ABT-530<br />

with or without ribavirin (RBV) in non-cirrhotic GT2-infected<br />

treatment-naïve (TN) and pegylated interferon/RBV (pegIFN/<br />

RBV) treatment experienced (TE) subjects Methods: Subjects<br />

received 12 weeks of ABT-493 300 mg+ABT-530 120 mg<br />

(Arm A), ABT-493 200 mg+ABT-530 120 mg (Arm B), or ABT-<br />

493 200 mg+ABT-530 120 mg+RBV (Arm C). DAAs were<br />

dosed once daily; weight-based RBV (1000 or 1200 mg) was<br />

dosed twice daily. Subjects were then followed for 24 weeks.<br />

The primary efficacy endpoint is sustained virologic response<br />

12 weeks after the last dose of study drug (SVR12); we present<br />

post-treatment week 4 data (SVR4) here. Safety was evaluated<br />

by monitoring adverse events (AEs), laboratory tests,<br />

and vital signs. Demographics and safety are presented on<br />

all randomized subjects. Efficacy is presented on all subjects<br />

confirmed to have GT2 infection. Results: 75 subjects were<br />

treated in Arms A–C (n=25 each); 74 had GT2, and 1 subject<br />

initially randomized to Arm B was determined to have GT3a<br />

infection. Subjects were male, 63%; median (range) age, 57.0<br />

(20.0–69.0) years; GT2b, 81%; TN, 88%; TE, 12%; F0–F2,<br />

87%; F3, 13%; median (range) baseline HCV RNA log 10<br />

IU/<br />

mL, 7.1 (4.7–7.8). No subjects have experienced virologic<br />

failure. One subject in Arm A prematurely discontinued study<br />

drugs and is lost to follow-up. The SVR4 rates (ITT analysis) are<br />

96% (24/25), 100% (24/24), and 100% (25/25) in Arms


340A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

A, B, and C, respectively. Most AEs were mild, with the most<br />

common DAA-related AEs being fatigue, nausea, headache,<br />

and diarrhea. There were no serious DAA-related AEs. Typical<br />

reductions in hemoglobin were observed in the RBV-containing<br />

arm. Conclusions: ABT-493+ABT-530 with or without RBV for<br />

12 weeks was highly effective and well tolerated, achieving<br />

SVR4 rates of 96%–100% and no virologic failure to-date;<br />

SVR12 data will be available at the time of the meeting. These<br />

data warrant further study of this once daily RBV-free ABT-493/<br />

ABT-530 combination in GT2 infection, including patients with<br />

cirrhosis.<br />

Disclosures:<br />

David L. Wyles - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Janssen, Merck; Grant/Research Support: Gilead, Merck, Bristol Myers Squibb,<br />

AbbVie, Tacere<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Stanley Wang - Employment: AbbVie<br />

Benedict Maliakkal - Speaking and Teaching: Valeant Pharma (Salix), Gilead,<br />

AbbVie, Bristol Myers Squibb<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Sandra S. Lovell - Employment: AbbVie<br />

Chih-Wei Lin - Employment: Abbvie<br />

Teresa Ng - Employment: AbbVie; Patent Held/Filed: AbbVie; Stock Shareholder:<br />

AbbVie<br />

Federico J. Mensa - Employment: Abbvie Inc.; Stock Shareholder: Abbvie Inc.<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

The following authors have nothing to disclose: Michael Bennett, Hugo E. Vargas,<br />

J. Scott Overcash, Asma Siddique, Bal R. Bhandari<br />

251<br />

High Efficacy of Grazoprevir and Elbasvir With or Without<br />

Ribavirin in 103 Treatment-Naive and Experienced<br />

Patients With HCV Genotype 4 Infection: A Pooled Analysis<br />

Tarik Asselah 1 , Hendrik W. Reesink 2 , Jan Gerstoft 3 , Victor de Ledinghen<br />

4 , Paul J. Pockros 5 , Michael Robertson 6 , Peggy Hwang 6 ,<br />

Bach-Yen T. Nguyen 6 , Janice Wahl 6 , Eliav Barr 6 , Rohit Talwani 6 ,<br />

Lawrence Serfaty 7 ; 1 Hôpital Beaujon, Clichy, France; 2 Academical<br />

Medical Center, Amsterdam, Netherlands; 3 Rigshospitalet, University<br />

of Copenhagen, Copenhagen, Denmark; 4 CHU Bordeaux,<br />

Pessac, France; 5 Scripps Clinic, La Jolla, CA; 6 Merck & Co., Inc.,<br />

Kenilworth, NJ; 7 Saint-Antoine Hospital, Paris, France<br />

Background: Genotype (GT) 4 accounts for around 15% of the<br />

170 million HCV infections worldwide, with high prevalence<br />

in Middle East and sub-Saharan Africa, but also increasing in<br />

Europe. Data on the efficacy of direct acting antiviral regimens<br />

in the GT4 population are limited. We evaluated the efficacy<br />

of the protease inhibitor grazoprevir 100 mg (GZR) and NS5A<br />

inhibitor elbasvir 50 mg (EBR), with and without (±) ribavirin<br />

(RBV) in GT4 infected patients enrolled in the GZR/EBR<br />

phase 2/3 clinical program. Methods: Two pooled datasets<br />

were evaluated: (1) treatment-naive (TN) patients who received<br />

12 weeks of GZR/EBR ± RBV and (2) treatment-experienced<br />

(TE) patients, defined as those who had previously failed<br />

peginterferon + RBV (PR) ± first-generation protease inhibitor<br />

who received 12, 16, or 18 weeks of GZR/EBR ± RBV. TE<br />

patients were further classified as failures due to prior relapse<br />

or prior on-treatment virologic failure (OTF) (eg, null or partial<br />

response; virologic breakthrough). Efficacy was defined as sustained<br />

virologic response (HCV RNA < assay-specified lower<br />

limit of quantification) 12 weeks after end of therapy (SVR12)<br />

in the full analysis set (all patients who received at least 1 dose<br />

of study drug). Results: Data from a total of 103 HCV-GT4 (47<br />

GT4a, 47 GT4d, 9 GT4-other)–infected patients, 66 TN and<br />

37 TE, including 17 cirrhotic patients, were analyzed. Overall<br />

64/66 (97%) TN (including 6/6 cirrhotic patients) and 32/37<br />

(86%) TE GT4 patients achieved SVR12. Among TE patients,<br />

9/9 prior relapsers and 23/28 (82%) prior OTFs achieved<br />

SVR12. Among TE cirrhotic patients, 13/17 (77%) including<br />

3/3 prior relapsers and 4/4 treated with GZR/EBR+RBV for<br />

16 weeks achieved SVR12. For prior OTFs, longer durations<br />

and use of RBV increased the response rate. Additional stratification<br />

of treatment responses according to regimen is included<br />

in Table 1. Conclusion: A 12-week regimen of GZR/EBR was<br />

highly effective among 103 GT4-infected TN and prior relapse<br />

patients. A regimen of GZR/EBR + RBV for 16 weeks may<br />

result in improved efficacy among prior OTF patients, although<br />

sample size was small. Further <strong>studies</strong> among subpopulations<br />

are needed.<br />

Table 1<br />

a Other: Non-SVR, but did not meet virologic failure criteria<br />

Disclosures:<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

Hendrik W. Reesink - Consulting: Abbvie, Alnylam, BMS, Gilead, Janssen-Cilag,<br />

Merck, PRA-International, Regulus, Replicor, Roche, R-Pharm; Grant/Research<br />

Support: AbbVie, BMS, Boehringer Ingelheim, Gilead, Janssen-Cilag, Merck,<br />

PRA-International, Regulus, Replicor, Roche<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

Paul J. Pockros - Advisory Committees or Review Panels: Janssen, Merck, BMS,<br />

Gilead, AbbVie; Consulting: Lumena, Beckman Coulter; Grant/Research Support:<br />

Intercept, Janssen, BMS, Gilead, Lumena, Beckman Coulter, AbbVie, RMS,<br />

Merck; Speaking and Teaching: AbbVie, Janssen, Gilead<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Peggy Hwang - Employment: Merck, Merck<br />

Bach-Yen T. Nguyen - Employment: Merck<br />

Janice Wahl - Employment: Merck & Co,<br />

Eliav Barr - Employment: Merck<br />

Rohit Talwani - Employment: Merck<br />

Lawrence Serfaty - Advisory Committees or Review Panels: MSD, Janssen, Roche,<br />

Gilead, BMS, Abbvie; Speaking and Teaching: Aptalis<br />

The following authors have nothing to disclose: Jan Gerstoft


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 341A<br />

252<br />

Daclatasvir in combination with sofosbuvir with or without<br />

ribavirin is safe and efficacious in liver transplant<br />

recipients with HCV recurrence: Interim results of a multicenter<br />

compassionate use program<br />

Kerstin Herzer 1 , Tania M. Welzel 2 , Peter Ferenci 3 , Joerg Petersen 4 ,<br />

Michael Gschwantler 5 , Markus Cornberg 6 , Thomas Berg 7 , Ulrich<br />

Spengler 8 , Ola Weiland 9 , Marc van der Valk 10 , Hartwig Klinker 11 ,<br />

Jürgen K. Rockstroh 8 , Eckart Schott 12 , Markus Peck-Radosavljevic 3 ,<br />

Yue Zhao 13 , Maria Jesus Jimenez Exposito 14 , Stefan Zeuzem 2 ;<br />

1 Universitätsklinikum Essen (AöR), Essen, Germany; 2 Universitätsklinikum<br />

der Johann Wolfgang Goethe Universität, Frankfurt, Germany;<br />

3 Medizinische Universität Wien, Vienna, Austria; 4 IFI Institut<br />

für Interdisziplinäre Medizin, Hamburg, Germany; 5 Wilhelminenspital,<br />

Vienna, Austria; 6 Medizinische Hochschule Hannover, Hannover,<br />

Germany; 7 Universitätsklinikum Leipzig, Leipzig, Germany;<br />

8 Universitätsklinikum Bonn, Bonn, Germany; 9 Karolinska Institutet,<br />

Solna, Sweden; 10 Academic Medical Center, University of Amsterdam,,<br />

Amsterdam, Netherlands; 11 Universitätsklinikum Würzburg,<br />

Würzburg, Germany; 12 Charité Universitätmedizin Berlin, Berlin,<br />

Germany; 13 Global Biostastistics, Bristol-Myers Squibb, Hopewell,<br />

NJ; 14 Research and Development, Bristol-Myers Squibb, Princeton,<br />

NJ<br />

BACKGROUND: Interferon (IFN)-based therapy for HCV postliver<br />

transplant (LT) has been associated with suboptimal<br />

response and poor tolerability. New IFN-free regimens have<br />

the potential to improve response rates and safety. The all-oral,<br />

pangenotypic regimen of daclatasvir (DCV) and sofosbuvir<br />

(SOF) with ribavirin (RBV) was well tolerated and achieved<br />

high SVR rates in patients with post-LT recurrence or advanced<br />

cirrhosis in the ALLY-1 study. We report here a detailed subanalysis<br />

of LT recipients with HCV recurrence from a compassionate<br />

use program (CUP) of DCV+SOF±RBV. METHODS: This<br />

multicenter CUP, opened in 5 European countries, enrolled<br />

adult patients with chronic HCV infection who are at a high<br />

risk of hepatic decompensation or death within 12 months if<br />

left untreated, and who had no available treatment options.<br />

Patients received DCV 60mg + SOF 400mg once daily for<br />

24 weeks, with RBV added at the physician’s discretion. This<br />

interim analysis includes 87 post-LT patients enrolled in the CUP<br />

with available data on March 16, 2015, of whom 38 patients<br />

have SVR12 data available to date. RESULTS: Most patients<br />

were male (69%), white (93%), with a median age of 58 years<br />

(39–75). Mean (SD) baseline HCV RNA 5.8 (1.7) log 10<br />

IU/<br />

mL. Most patients were infected with HCV GT1 (87%) (1a,<br />

n=28; 1b, n=37; 1 unknown, n=11) or GT3 (8%). Median<br />

time since LT to treatment initiation was 3.3 years. Most frequent<br />

immunosuppressant includes tacrolimus (n=47) and cyclosporine<br />

(n=11). Cirrhosis was present in 35 (40%) patients;<br />

among cirrhotic patients, 13 (37%) were Child-Pugh class B/C,<br />

and 8 (23%) had MELD scores ≥15. Twenty-six (30%) patients<br />

received RBV as part of the regimen. Baseline characteristics<br />

and results for each treatment group are described in the Table.<br />

Overall, of the 38 patients with available HCV RNA values at<br />

posttreatment week 12 to date, all achieved virologic response<br />

(SVR12 100%). Updated SVR12 data will be presented. Serious<br />

AEs occurred in 18 patients (21%), 6 patients discontinued<br />

treatment due to AEs and 3 patients died. CONCLUSION: In<br />

this preliminary analysis of patients in a real-world setting,<br />

DCV+SOF±RBV achieved high rates of SVR12 and was well<br />

tolerated in this difficult-to-treat post-liver transplant patient population.<br />

The use of ribavirin did not affect efficacy outcomes.<br />

Disclosures:<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

Peter Ferenci - Advisory Committees or Review Panels: Idenix, Gilead, MSD,<br />

Janssen, Salix, AbbVie, BMS; Patent Held/Filed: Madaus Rottapharm; Speaking<br />

and Teaching: Gilead, Roche<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

Michael Gschwantler - Advisory Committees or Review Panels: Janssen, BMS,<br />

Gilead, AbbVie; Speaking and Teaching: Janssen, BMS, Gilead, AbbVie<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Ola Weiland - Advisory Committees or Review Panels: Merk, BMS, Medivir, Gilead,<br />

AbbVie; Grant/Research Support; Speaking and Teaching: Merk, Roche,<br />

BMS, Novartis, Janssen, Medivir, Gilead, AbbVie<br />

Marc van der Valk - Advisory Committees or Review Panels: gilead, msd, bms,<br />

abbvie, janssen cilag, viiV, roche<br />

Jürgen K. Rockstroh - Advisory Committees or Review Panels: Abbvie, BI, BMS,<br />

Merck, Roche, Tibotec, Abbvie, Bionor, Tobira, ViiV, Gilead, Janssen; Consulting:<br />

Novartis; Grant/Research Support: Merck; Speaking and Teaching: Abbott,<br />

BI, BMS, Merck, Roche, Tibotec, Gilead, Janssen, ViiV<br />

Eckart Schott - Advisory Committees or Review Panels: Gilead, Roche, Bayer,<br />

BMS, Abbvie; Speaking and Teaching: Gilead, Novartis, Roche, MSD, Bayer,<br />

Falk, BMS, Janssen, Abbvie<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Maria Jesus Jimenez Exposito - Employment: Bristol-Myers Squibb<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

The following authors have nothing to disclose: Kerstin Herzer, Ulrich Spengler,<br />

Hartwig Klinker, Yue Zhao<br />

253<br />

Molecular heterogeneity in multinodular Hepatocellular<br />

Carcinoma<br />

Daniela Sia 1,2 , Andrew Harrington 1 , Zhongyang Zhang 3 , Genis<br />

Camprecios 1 , Sara Toffanin 1 , Agrin Moeini 1,2 , M. Isabel Fiel 1 ,<br />

Ke Hao 3 , Monica Higuera 2 , Oriana Miltiadous 1 , Laia Cabellos 2 ,<br />

Helena Cornella 2 , Yujin Hoshida 1 , Sander S. Florman 1 , Myron E.<br />

Schwartz 1 , Josep M. Llovet 1,2 ; 1 Liver Cancer Program, Dept of Liver<br />

Diseases, Icahn School of Medicine at Mount Sinai, New York,<br />

NY; 2 Liver Cancer Translational Research Laboratory, IDIBAPS,<br />

Hospital Clinic, Barcelona, Spain; 3 Icahn Institute for Multiscale<br />

Biology, Icahn School of Medicine at Mount Sinai, New York, NY<br />

Introduction: Whether multinodular hepatocellular carcinomas<br />

(HCCs) result from intrahepatic metastasis (IM or clonal tumors)


342A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

or de novo cancers (synchronic tumors) has direct implications<br />

in treatment decision-making. We aimed to assess the genomic<br />

heterogeneity of multifocal HCC using single-nucleotide polymorphism<br />

(SNP), gene expression and mutations analyses.<br />

Methods: Among 544 HCCs consecutively transplanted at<br />

Mount Sinai Hospital, 18 patients with 2-3 non-satellite nodules/patient<br />

were selected for this study. Formalin-fixed tissue<br />

blocks and clinico-pathological parameters were collected for<br />

each tumor. SNP array and gene expression profiling was generated.<br />

Clonality was defined by measuring SNP profiles similarity<br />

between two nodules. Prediction of previously published<br />

signatures was performed using Nearest Template Prediction.<br />

Mutational status of TERT promoter, ARID1A, CTNNB1, TP53,<br />

AXIN1-2 genes was checked using TruSeqAmplicon. Results:<br />

A total of 42 tumors have been analyzed. Most patients were<br />

male (17/18, 95%) with HCV (10/18, 56%) or HBV infection<br />

(6/18, 33%). Median tumor size was 3 cm (range 1.5-<br />

6.5), satellites and vascular invasion were present in 3 (17%)<br />

and 11 patients (61%), respectively. CNV profiles predicted<br />

clonal tumors in 38% (6/16) and non-clonality in 63% of cases<br />

(10/16), while the remaining 2 cases were not informative.<br />

Clonal tumors were significantly associated with HCV infection<br />

(5/6 vs 3/10, p=0.01), whereas all HBV-induced HCC were<br />

synchronic tumors (0/6 vs 6/10, p=0.03). Clonal tumors were<br />

significantly associated with presence of satellites and early<br />

recurrence. Gene expression-based unsupervised clustering<br />

revealed that each clonal tumor showed genetic proximity to its<br />

paired tumor and clustered around the same node (6/6,100%)<br />

as opposed to non-clonal (2/9, 22%). When exploring molecular<br />

subclasses, while half of clonal tumors retained the molecular<br />

fingerprint, the other half switched to more aggressive<br />

subclasses. Conversely, all non-clonal tumors belonged to distinct<br />

molecular subclasses. In order to explore if clonal tumors<br />

retain the same driver trunk mutations, we tested five common<br />

driver mutations in all HCC nodules. Identical trunk mutations<br />

were identified only in clonal tumors. Conclusions: Among<br />

multinodular HCC tumors, 40% of cases were classified as<br />

clonal tumors (true IM) and the remaining 60% as non-clonal<br />

synchronic tumors. Clonal tumors were significantly associated<br />

with HCV infection, satellites and recurrence. Despite molecular<br />

subclasses prediction revealed heterogeneous profile in<br />

both clonal and non-clonal tumors, genetic proximity and trunk<br />

driver mutations were similar in clonal tumors.<br />

Disclosures:<br />

Josep M. Llovet - Consulting: Bayer Pharmaceuticals, Bristol Myer Squibb, Boehringer-Ingelheim,<br />

Eli Lilly Pharmaceuticals, Celsion, Biocompatibles, Novartis,<br />

GlaxoSmithKline, Blueprint Medicines; Grant/Research Support: Bayer Pharmaceuticals,<br />

Bristol Myers Squibb, Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Daniela Sia, Andrew Harrington,<br />

Zhongyang Zhang, Genis Camprecios, Sara Toffanin, Agrin Moeini, M. Isabel<br />

Fiel, Ke Hao, Monica Higuera, Oriana Miltiadous, Laia Cabellos, Helena Cornella,<br />

Yujin Hoshida, Sander S. Florman, Myron E. Schwartz<br />

254<br />

Mixed hepatocellular-cholangiocarcinoma tumors:<br />

stem-cell subtype and cholangiolocellular carcinoma are<br />

unique molecular entities<br />

Agrin Moeini 1,2 , Daniela Sia 2,1 , Zhongyang Zhang 3 , Genis<br />

Camprecios 2 , M. Isabel Fiel 2 , Oriana Miltiadous 2 , Xiaochen<br />

Sun 2 , Ke Hao 3 , Yujin Hoshida 2 , Swan N. Thung 2 , Augusto Villanueva<br />

2 , Myron E. Schwartz 2 , Josep M. Llovet 2,1 ; 1 Liver Cancer<br />

Translational Research Laboratory, Liver Unit, Institut d’Investigacions<br />

Biomèdiques August Pi i Sunyer (IDIBAPS), Hospital Clínic,<br />

CIBERehd, Universitat de Barcelona, Barcelona, Spain; 2 Mount<br />

Sinai Liver Cancer Program, (Divisions of Liver Diseases, Department<br />

of Medicine, Department of Pathology, Recanati Miller<br />

Transplantation Institute), Tisch Cancer Institute, Icahn School<br />

of Medicine at Mount Sinai, New York, NY; 3 Icahn Institute for<br />

Genomics and Multiscale Biology, Icahn School of Medicine at<br />

Mount Sinai, New York, NY<br />

Purpose: Provide a comprehensive analysis of the molecular<br />

alterations of mixed hepatocellular-cholangiocarcinoma (HCCiCCA)<br />

tumors, a rare type of primary liver cancer, by integrating<br />

histological and genomic data. Methods: 19 patients<br />

with mixed HCC-iCCA were classified based on histological<br />

features and immunohistochemistry for specific hepatocyte<br />

(HEP1, GPC3), biliary (CK7, CK19) and stem-cell markers<br />

(EpCAM, NCAM, SALL4). Whole-genome gene-expression<br />

and DNA copy number alterations (CNA) analysis were performed<br />

(Illumina). Prevalent oncogenic mutations in HCC (TERT<br />

promoter, TP53, CTNNB1) or iCCA (FGFR2-fusions, IDH1/2,<br />

KRAS, BRAF) were screened. Results: Following WHO criteria,<br />

HCC-iCCA samples were classified as classical (n=4), stem-cell<br />

feature (n=9) and cholangiolocellular carcinoma (CLC) (n=6).<br />

From the pathological standpoint, classical HCC-iCCA were<br />

mainly EpCAM positive (3/4, 75%) with expression of hepatocyte<br />

and biliary markers in the HCC-like and iCCA-like components,<br />

respectively. Stem-cell subtypes were characterized<br />

by SALL4 positive staining (7/9 vs 0/10 others, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 343A<br />

Disclosures:<br />

Josep M. Llovet - Consulting: Bayer Pharmaceuticals, Bristol Myer Squibb, Boehringer-Ingelheim,<br />

Eli Lilly Pharmaceuticals, Celsion, Biocompatibles, Novartis,<br />

GlaxoSmithKline, Blueprint Medicines; Grant/Research Support: Bayer Pharmaceuticals,<br />

Bristol Myers Squibb, Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Agrin Moeini, Daniela Sia,<br />

Zhongyang Zhang, Genis Camprecios, M. Isabel Fiel, Oriana Miltiadous, Xiaochen<br />

Sun, Ke Hao, Yujin Hoshida, Swan N. Thung, Augusto Villanueva, Myron<br />

E. Schwartz<br />

255<br />

Loss of adherens junctions proteins disrupts blood<br />

bile barrier to cause progressive familial intrahepatic<br />

cholestasis (PFIC)-like phenotype<br />

Kari Nejak-Bowen, Lili Zhou, An Jiang, Minakshi Poddar, Satdarshan<br />

(Paul) S. Monga; University of Pittsburgh, Pittsburgh, PA<br />

Tight junction (TJ) proteins are traditionally thought to be the<br />

chief regulators of blood-bile barrier in the liver. In fact, loss<br />

of TJP2 has been shown to be associated with a subset of<br />

PFIC cases. PFIC is traditionally linked to loss of function mutations<br />

in various bile acid transporters such as BSEP, MDR3 and<br />

ATP8B1, although the molecular basis of a notable subset of<br />

these cases remains elusive. Further, the role of adherens junctions<br />

(AJ) in regulating blood-bile barrier or in PFIC pathogenesis<br />

remains unknown. AJ regulate cell-cell adhesion by linking<br />

E-cadherin of neighboring cells to respective actin cytoskeleton<br />

through β-catenin and α-catenin. Intriguingly, we showed that<br />

conditional loss of β-catenin in hepatocytes was well compensated<br />

by upregulation of γ-catenin, a desmosomal protein that<br />

maintained the E-cadherin-actin cytoskeleton link. To characterize<br />

this redundancy in catenin proteins, we generated<br />

liver-specific β-catenin-γ-catenin double knockout (DKO) mice<br />

using cre-lox. DKO show failure to thrive and most animals<br />

die by 1 month. Kaplan Meier curve shows significantly lower<br />

survival as compared to littermate controls (Con) (p


344A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

257<br />

Hepatic XBP1 deletion promotes endoplasmic reticulum<br />

stress-induced liver injury and apoptosis<br />

Shantel Olivares, Anne Henkel; Medicine, Northwestern University,<br />

Chicago, IL<br />

Endoplasmic reticulum (ER) stress triggers activation of the<br />

unfolded protein response (UPR), a highly conserved signaling<br />

cascade that functions to alleviate stress and promote cell survival.<br />

If, however, the cell is unable to restore homeostasis, the<br />

UPR activates pathways that promote apoptotic cell death. The<br />

molecular mechanisms governing this critical transition from restoration<br />

of homeostasis to initiation of apoptosis remain poorly<br />

understood. We aim to determine the role of hepatic XBP1, a<br />

key regulator of the UPR, in controlling cell fate in response<br />

to ER stress. Liver-specific XBP1 knockout mice (XBP1 LKO ) and<br />

XBP1 fl/fl littermate controls were subjected to varying levels and<br />

duration of pharmacologic ER stress. XBP1 LKO mice and XBP1 fl/<br />

fl<br />

controls showed robust and equal induction of the UPR acutely<br />

as evidenced by equivalent hepatic expression of Grp78/BIP,<br />

CHOP, and ATF4 six hours after exposure to pharmacologic<br />

ER stress. By 72 hours, XBP1 fl/fl control mice showed complete<br />

resolution of UPR activation and no liver injury suggesting complete<br />

restoration of homeostasis. Conversely, XBP1 LKO mice<br />

showed ongoing UPR activation at 3 and 7 days after induction<br />

of ER stress associated with progressive liver injury histologically.<br />

Consistent with enhanced hepatic injury, plasma ALT<br />

levels were markedly elevated in XBP1 LKO mice at days 3 and 7<br />

(283±41 vs 49±12 IU/L at day 3 and 217±3 vs 47±11 IU/L at<br />

day 7 vs XBP1 fl/fl controls, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 345A<br />

259<br />

Midodrine and Tolvaptan in Patients with Cirrhosis and<br />

Refractory or Recurrent Ascites : A Randomized Pilot<br />

Study<br />

Virendra Singh 1 , Nitish Rai 2 , Baljinder Singh 3 , Rajesh Vijayvergiya<br />

4 , Navneet Sharma 2 , Ashish Bhalla 2 ; 1 Hepatology,<br />

Postgraduate Institute of Medical Education and Research, Chandigarh,<br />

India; 2 Internal Medicine, Postgraduate Institute of Medical<br />

Education and Research, Chandigarh, India; 3 Nuclear Medicine,<br />

Postgraduate Institute of Medical Education and Research, Chandigarh,<br />

India; 4 Cardiology, Postgraduate Institute of Medical Education<br />

and Research, Chandigarh, India<br />

Background: Splanchnic arterial vasodilatation and subsequent<br />

activation of antinatriuretic and vasoconstrictive mechanisms<br />

resulting in sodium and water retention play an important role<br />

in cirrhotic ascites. Midodrine and tolvaptan have been used<br />

in these patients, separately. However, there are no reports<br />

on the use of combination of midodrine and tolvaptan in the<br />

control of ascites. The aim of this study was to evaluate the<br />

effects of midodrine,tolvaptan and their combination on systemic<br />

hemodynamics, renal function and control of ascites in<br />

patients with cirrhosis and refractory or recurrent ascites. Methods:<br />

Fifty cirrhotic patients with refractory or recurrent ascites<br />

were prospectively studied after long-term administration of<br />

midodrine (7.5 mg three times a day; n=13) or tolvaptan (15<br />

mg twice a day; n=12) or both (n=13) plus standard medical<br />

therapy (SMT) or SMT alone (n=12) in a randomized controlled<br />

trial (NCT02173288) at a tertiary center. Results: A<br />

significant increase in urinary volume (midodrine:1077±274<br />

vs 1440±270 ml/day, p=0.004; tolvaptan:1150±344 vs<br />

2250±688 ml/day, p=0.001; combination:1130±233<br />

vs 3110±1170 ml/day, p=0.0001 at 1 and 3 months,<br />

p


346A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Peter C. Hayes - Advisory Committees or Review Panels: Roche, MSD, Jannsen,<br />

Gilead, ?ONO, Norgine; Grant/Research Support: Novartis; Speaking and<br />

Teaching: Roche, MSD, Pfiser, Gore, Falk, Ferring<br />

The following authors have nothing to disclose: Neil J. Lachlan, Scott I. Semple,<br />

Dilip Patel, David Carr, John P. Iredale, Jonathan A. Fallowfield<br />

261<br />

Ascites neutrophil dysfunction in patients with decompensated<br />

cirrhosis is linked to ascites protein content<br />

Cornelius Engelmann 1 , Christina Becker 2 , Andreas Boldt 2 , Sandra<br />

Krohn 1 , Michael Bartels 3 , Thomas Berg 1 , Ulrich Sack 2 ; 1 Section<br />

of Hepatology; Department of Internal Medicine, Neurology, Dermatology,<br />

University Hospital Leipzig, Leipzig, Germany; 2 Institute<br />

of Clinical Immunology, University of Leipzig, Leipzig, Germany;<br />

3 Department of Visceral, Vascular, Thoracic and Transplant Surgery,<br />

University Hospital Leipzig, Leipzig, Germany<br />

Background: Spontaneous bacterial peritonitis is a major and<br />

life threatening complication in patients with decompensated<br />

cirrhosis. Immune paralysis and neutrophil dysfuntion has been<br />

regarded as a main driver for the increased risk of bacterial<br />

infections in these patients. It remains unknown whether this<br />

feature can be extrapolated to immune cells in the peritoneal<br />

cavity. We therefore evaluated the cell function of neutrophils<br />

(PMN) in ascites fluid. Material and Methods: In total 71 blood<br />

and ascites samples of mainly male patients (n=48) with cirrhosis<br />

(n=63) and without cirrhosis (malignant n=6, cardiac n=1,<br />

Budd-Chiari syndrome n=1) were collected between August<br />

2014 and May 2015. PMNs were identified by flow-cytometry.<br />

Phagocytic activity (PA) as well as oxidative burst activity<br />

(OBA) of both blood and ascites PMNs were evaluated after<br />

in vitro stimulation with E.coli cells. Fluorescence signal was<br />

measured by flow cytometry and PA and OBA was expressed<br />

as percentage of viable PMNs. Function tests of ascites PMNs<br />

were repeated in human albumin and autologous blood<br />

plasma. Results: As compared to their blood counterparts, ascites<br />

PMNs showed a significantly impaired PA and OBA with a<br />

much broader activity range (median blood PA (n=70) 98.1%<br />

(86.8-99.8) vs. median ascites PA (n=68) 52.95% (0.4-97.3),<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 347A<br />

263<br />

TIPS with PTFE-covered stent improves liver transplant<br />

free survival in patients with cirrhosis and recurrent<br />

ascites : results of a multicentre randomized trial<br />

Christophe Bureau 1,2 , Dominique Thabut 3 , Frederic Oberti 4 ,<br />

Sebastien Dharancy 5 , Pauline Cabarrou 1 , Nicolas Carbonell 3 ,<br />

Antoine Bouvier 4 , Philippe Mathurin 5 , Philippe Otal 1 , Jean-Marie<br />

Peron 1,2 , Jean-Pierre Vinel 1,2 ; 1 Service d’Hépatogastroenterologie,<br />

CHU Toulouse, Toulouse, France; 2 Université Paul Sabatier,<br />

Toulouse, France; 3 APHP, Paris, France; 4 CHU Angers, Angers,<br />

France; 5 CHRU Lille, Lille, France<br />

Introduction Transjugular intrahepatic portosystemic shunt<br />

(TIPS) proved to be effective in the treatment of refractory ascites.<br />

However, its impact on survival remains controversial. The<br />

high rate of shunt dysfunction with the use of uncovered stents<br />

could attenuate the beneficial effects of TIPS. It has been shown<br />

that the use of PTFE covered stents decreases the risk of shunt<br />

dysfunction and improves clinical outcome. The main objective<br />

was to compare the liver transplantation free survival of<br />

cirrhotic patients and recurrent ascites between those treated<br />

by covered TIPS and those treated by large volume paracentesis<br />

+ albumin infusion (P+A). Methods Between august 2005<br />

and november 2012, all cirrhotic patients with at least two<br />

large volume paracentesis were considered for inclusion. The<br />

main end point was 1-year liver transplantation free survival.<br />

Secondary end points were the rates of: ascites recurrence<br />

and treatment failure, overt hepatic encephalopathy and portal<br />

hypertension (PHT) related complications; and the number of<br />

days in hospitalization over a 1-year period after inclusion.<br />

Results 62 patients were included (29 in TIPS group and 33 in<br />

P+A group). The baseline characteristics of the patients were<br />

similar in both groups. The median follow up of the whole<br />

cohort was 365 days. One patient in each group was lost to<br />

follow up. Actuarial rates of transplantation free survival was<br />

better in TIPS group compared to that observed in the P+A<br />

group: 93 % [IC 95 % 82%-100 %] and 52 % [IC 95 % 34%-<br />

60 %], respectively (p=0.003). During the 1-year of follow<br />

up, the total number of paracentesis was 32 (1.0 ± 1.6 per<br />

patient) in TIPS group and 320 in P+A (10.1 ± 7.0 per patient).<br />

Total albumin infusion was 2061 g in TIPS group compared to<br />

17727 g in P+A group. In P+A group, 15 patients were treated<br />

by TIPS during follow up, 14 because of the need of more than<br />

6 large volume paracentesis within less than 3 months and 1<br />

because of associated recurrent variceal bleeding. The 1-year<br />

probability of remaining free of encephalopathy was 65 %<br />

in both groups. The rates of PHT related bleeding and hernia<br />

related complications were significantly higher in P+A group.<br />

The number of days in hospitalizations was doubled in P+A<br />

group compared to TIPS group (17 ± 28 vs 35 ± 40 p = 0.04).<br />

Conclusion: the use of covered TIPS in selected patients with<br />

recurrent ascites improved transplant free survival as compared<br />

to patients treated by repeated large volume paracentesis +<br />

albumin infusion.<br />

Disclosures:<br />

Christophe Bureau - Grant/Research Support: Gore; Speaking and Teaching:<br />

Gore<br />

Frederic Oberti - Stock Shareholder: Bio Live Scale<br />

Sebastien Dharancy - Advisory Committees or Review Panels: CHIESI; Board<br />

Membership: NOVARTIS; Speaking and Teaching: ASTELLAS<br />

Philippe Mathurin - Board Membership: MSD, Janssen-Cilag, BMS, Gilead,<br />

Abvie, Oncozyme; Consulting: Roche, Bayer<br />

Jean-Marie Peron - Board Membership: BAYER; Consulting: BMS, GILEAD, BOS-<br />

TON SCIENTIFIC<br />

Jean-Pierre Vinel - Grant/Research Support: Roche, Gore, LFB<br />

The following authors have nothing to disclose: Dominique Thabut, Pauline<br />

Cabarrou, Nicolas Carbonell, Antoine Bouvier, Philippe Otal<br />

264<br />

Diagnostic accuracy of the Periscreen TM reagent strip<br />

in diagnosis of spontaneous bacterial peritonitis in cirrhotic<br />

patients (PerDRISLA study)<br />

Thierry Thevenot 1 , Charline Briot 1 , Vincent Macé 2,22 , Hortensia<br />

Lison 3 , Laure Elkrief 4 , Alexandra Heurgué-Berlot 5 , Christophe<br />

Bureau 6 , Caroline Jezequel 7 , Ghassan Riachi 8 , Alexandre Louvet<br />

9 , Isabelle Ollivier-Hourmand 10 , Hélène Labadie 11 , Nicolas<br />

Carbonell 12 , Karim Aziz 13 , Denis Grasset 14 , Rodolphe Anty 15 ,<br />

Eric Nguyen-Khac 16 , Mehdi Kaasis 17 , Thomas Mouillot 18 , Arnaud<br />

Pauwels 19 , Florence Tanné 20 , Si Nafa Si Ahmed 21 , Jean Paul<br />

Cervoni 1 , Jean françois D. Cadranel 3 , Matthieu Schnee 2 ; 1 Hepatogastroenterology,<br />

CHU Minjoz, Besançon, France; 2 Hepatogastroenterology,<br />

CHD-Vendée, La Roche sur Yon, France;<br />

3 Hepatogastroenterology, Höpital Laennec, Creil, France; 4 Hepatology,<br />

CHU- Beaujon, Clichy, France; 5 Hepatogastroenterology,<br />

CHU Reims, Reims, France; 6 Hepatogastroenterology, CHU Purpan,<br />

Toulouse, France; 7 Hepatogastroenterology, CHU Pontchaillou,<br />

Rennes, France; 8 Hepatogastroenterology, CHU Charles<br />

Nicolle, Rouen, France; 9 Hepatology, CHU Claude Huriez, Lille,<br />

France; 10 Hepatogastroenterology, CHU Caen, Caen, France;<br />

11 Hepatogastroenterology, CH de Saint-Denis, Saint-Denis,<br />

France; 12 Hepatogastroenterology, Hôpital Saint-Antoine, Paris,<br />

France; 13 Hepatogastroenterology, CH de Saint-Brieuc, Saint-<br />

Brieuc, France; 14 Hepatogastroenterology, CH Bretagne-Atlantique,<br />

Vannes, France; 15 Hepatogastroenterology, CHU Archet,<br />

Nice, France; 16 Hepatogastroenterology, CHU Amiens, Amiens,<br />

France; 17 Hepatogastroenterology, CH de Cholet, Cholet, France;<br />

18 Hepatogastroenterology, Hôpital du Bocage, Dijon, France;<br />

19 Hepatogastroenterology, CH de Gonesse, Gonesse, France;<br />

20 Hepatogastroenterology, CHU La-Cavale-Blanche, Brest, France;<br />

21 Hepatogastroenterology, CHR d’Orléans, Orléans, France;<br />

22 Hepatogastroenterology, CHU de Nantes, Nantes, France<br />

Background & Aim: The diagnostic accuracy of spontaneous<br />

bacterial peritonitis (SBP) using reagent strips is currently disappointing<br />

(1) but an experimental work using the Periscreen<br />

TM strip was more optimistic (2). We aim to assess the<br />

performance of the Periscreen TM strip for the diagnosis of SBP.<br />

Methods: This study involved all consecutive cirrhotic patients<br />

with ascites between June 2014 and March 2015 among<br />

22 French centers. Ascitic fluid was tested independently by<br />

two investigators using Periscreen TM at 3 minutes as reported<br />

(2). The strip has 4 colorimetric graduations (negative, trace,<br />

small or large). The diagnosis of SBP was based on neutrophils<br />

>250/mm 3 , regardless the culture of ascites. We excluded<br />

bloody, chylous or bilious ascitic fluid, or concurrent imipenem<br />

treatment. Results: The characteristics of the 387 cirrhotic<br />

patients included were: mean age 61.9±10.4 years, 75.4%<br />

males, 73.1% alcohol, mean Child-Pugh 9.6 (range: 6-15). A<br />

total of 1190 paracenteses were analyzed: 468 (39.3%) were<br />

collected in patients hospitalized for an acute decompensation<br />

of cirrhosis and 722 (60.7%) in outpatients. For inter-investigator<br />

agreement, the Kappa value was 0.79 (IC95%: 0.76-<br />

0.83). A total of 72 (5.8%) samples had SBP, 58 in inpatient<br />

(11.5%) and 14 (1.9%) in outpatient setting (P


348A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(64.9-83.4), PPV 34.3% (19.1-52.2) and VPN 97.2% (90.2-<br />

99.7). In this study, the prevalence of SIRS was 13.2%. In<br />

5 samples, neutrophil count was above 250/mm 3 (385 to<br />

13020/ mm 3 ) but the strip was considered as negative. When<br />

setting “trace” as the first positive result, Se and NPV were both<br />

of 100% for outpatients. Conclusions: This study shows, for the<br />

first time, in a large sample size of cirrhotic patients that the<br />

Periscreen TM strip is a robust screening tool for the detection of<br />

SBP. More importantly, this new leukocyte esterase strip has<br />

an excellent accuracy to exclude SBP in the outpatient setting.<br />

(1) Nousbaum JB, et al. Hepatology 2007;45:1275-1281. (2)<br />

Mendler MH, J Hepatol 2010;53:477-483.<br />

Disclosures:<br />

Alexandra Heurgué-Berlot - Consulting: Abbvie; Speaking and Teaching: Gilead<br />

Christophe Bureau - Grant/Research Support: Gore; Speaking and Teaching:<br />

Gore<br />

Isabelle Ollivier-Hourmand - Speaking and Teaching: gilead, jansen, bayer<br />

Eric Nguyen-Khac - Speaking and Teaching: Gilead, Abbvie, Janssen, Roche,<br />

MSD, BMS<br />

Si Nafa Si Ahmed - Board Membership: La Roche Hoffmann; Grant/Research<br />

Support: Janssen; Speaking and Teaching: BMS, Gilead, Abbvie<br />

The following authors have nothing to disclose: Thierry Thevenot, Charline Briot,<br />

Vincent Macé, Hortensia Lison, Laure Elkrief, Caroline Jezequel, Ghassan Riachi,<br />

Alexandre Louvet, Hélène Labadie, Nicolas Carbonell, Karim Aziz, Denis Grasset,<br />

Rodolphe Anty, Mehdi Kaasis, Thomas Mouillot, Arnaud Pauwels, Florence<br />

Tanné, Jean Paul Cervoni, Jean françois D. Cadranel, Matthieu Schnee<br />

265<br />

Deleterious effect of beta-blockers in cirrhotic patients<br />

with refractory ascites: potential role of myocardial dysfunction<br />

Valerio Giannelli 1,2 , Olivier Roux 2 , Pierre-Emmanuel Rautou 2 ,<br />

Emmanuel Weiss 2 , Richard Moreau 2 , Didier Lebrec 2 , Francois<br />

Durand 2 , claire francoz 2 ; 1 gastroenterology, university of rome<br />

“sapienza”, Rome, Italy; 2 Service d’Hépatologie, Hôpital Beaujon,<br />

Assistance Publique-Hôpitaux de Paris, Clichy, France; INSERM<br />

U1149,CRI, Paris, France<br />

Introduction: Recent <strong>studies</strong> have suggested that non selective<br />

beta-blockers (NSBB) could be deleterious in cirrhotic patients<br />

with refractory ascites (RA) by increasing transplant-free mortality.<br />

It has been proposed that NSBB may increase the risk<br />

of paracentesis-induced circulatory dysfunction but the mechanisms<br />

involved are still unclear. The aim of this study was to<br />

explore the impact of NSBB on systemic and splanchnic hemodynamics<br />

in cirrhotic patients with RA evaluated for liver transplantation<br />

(LT). Patients and methods: 583 cirrhotic patients<br />

evaluated for a first LT in a single center between 1997 and<br />

2013 were studied. There were 74% males, mean age was<br />

52 years. The cause of cirrhosis was alcohol in 43%, HCV<br />

in 24% and HBV in 11%. 29% of patients had hepatocellular<br />

carcinoma. 196 patients (34%) had RA, and among them, 51<br />

% received NSBB. Physiological MELD score was comparable<br />

in RA and no—RA groups (17 vs 16, ns) as well as the proportion<br />

of patients receiving NSBB (51 vs 51 %, ns). In addition<br />

to standard preLT workup, all patients had measurement of<br />

both splanchnic and systemic hemodynamics by right heart<br />

catheterization with assessment of cardiac work (Left and Right<br />

Ventricular Stroke Work Index, LVSWI and RVSW). Results:<br />

Waiting list mortality was higher in RA compared to no-RA<br />

group, (62 vs 38 % at 1 year, p=0.001). RA patients had significantly<br />

lower mean arterial pressure (MAP) lower heart rate<br />

(HR) and higher hepatic venous pressure gradient (HVPG) than<br />

no-RA patients. Cardiac index (CI) as well as systemic vascular<br />

resistance (SVRI) were similarcomparable in both groups. By<br />

contrast, Left and Right Ventricular Stroke Work Index (LVSWI<br />

and RVSWI) were significantly lower in RA patients compared<br />

to no-RA. In both RA and no-RA patients, NSBB resulted in a<br />

significant decrease in HR, MAP, and CI. In the no-RA group,<br />

NSBB did not affect LVSWI (57+15 vs 60+21 g.m/m 2 , ns). By<br />

contrast, in the RA group, NSBB were associated with a significant<br />

decrease in LVSWI (48+15 vs 56+15, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 349A<br />

48%, and 33%, respectively. In multivariate analysis, AKI<br />

stage1B or greater (HR=2.32; p


350A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

269<br />

Hepatiq Measure Of Hepatic Function: Threshold Function<br />

For Ascites And Death In Patients With Chronic Liver<br />

Disease (CLD)<br />

John C. Hoefs 1,2 , Lien Tran 2 , Dipu Ghosh 3 ; 1 Medicine, Univ of<br />

California, Orange, CA; 2 Medicine, Liver Specialtly Center, Irvine,<br />

CA; 3 HEPATIQ_LLC, Irvine, CA<br />

INTRODUCTION:Functional quantitative tests have been<br />

largely ignored in the non-invasive staging of CLD. The perfused<br />

hepatic mass (PHM) is a precise measure of liver function<br />

correlating with the functional hepatic mass (r2 = .905)(AmJ-<br />

Gastro;92:2054) and adverse clinical outcomes (ACO). PHM<br />

= 95 (HALT-C<br />

2012;Hepat;55:1019). HEPATIQ is an automated measure<br />

of PHM to make quantitative image analysis of SPECT images<br />

easier. METHODS: Sequential SPECT scans from 384 patients<br />

were processed by HEPATIQ for PHM to assess the relationship<br />

to ascites and death. There were normals (N) 29 and patients<br />

with CLD: HBV 61, HCV 138, NASH 49, PBC/ACAH/PSC<br />

34, ALD 13, abnormal AST/ALT44, post liver transplant (LT)<br />

6, and miscellaneous 16. Any ascites that required treatment<br />

was recorded. Patients with recovery from prior ACO from<br />

CLD associated with ascites (R-ACO: requiring no treatment ><br />

2 yrs) or before liver transplant (LT) and prior variceal bleeding<br />

without ascites (VB) were included along with treatable ascites<br />

(RxA), Refractory ascites (RfA), and death (D). SCAN: Patients<br />

were fed prior to IV injection of 5-6 mCi 99Tc sulfur colloid with<br />

subsequent SPECT /planar images and processed by HEPA-<br />

TIQ. RESULTS: 306 patients never had ACO, 5 had a variceal<br />

bleed with no ascites (VB) and 28 patients had current ascites:<br />

responsive ascites RxA 10, refractory ascites (RfA) 12, and<br />

death 6. PHM in N was 104.3+/-2.8 and 306 patients never<br />

having ACO 102.3+/-4.6. In 6 LT patients PHM 99.2+/-4.5<br />

and R-ACO 91.9+/-6.4. PHM was 77.7+/-6.9 with RxA; RfA<br />

64.2+/-11.6 and 6 deaths 49.9+/-16.1 (< 60 for 4 liver failure,<br />

2 HCC). For PHM threshold for ACO PHM < 95, RxA 85,<br />

RfA 78, and death < 60 (see figure) CONCLUSIONS: 1. PHM<br />

using HEPATIQ is a precise measure of CLD severity correlating<br />

with clinical outcomes regardless of CLD cause. 2. Hepatic<br />

function by PHM thresholds for RxA, RfA, and death are relatively<br />

precise, 3. PHM with HEPATIQ is a hepatic function test,<br />

valuable for non-invasive staging and monitoring of CLD.<br />

Disclosures:<br />

John C. Hoefs - Management Position: HEPATIQ_LLC; Speaking and Teaching:<br />

Gilead, Jannsen, Abbvie<br />

Dipu Ghosh - Management Position: HEPATIQ LLC<br />

The following authors have nothing to disclose: Lien Tran<br />

270<br />

The Trigger Matters – Outcome of Specifically-Triggered<br />

AKI versus HRS in Patients with Cirrhosis and Ascites<br />

Theresa Bucsics, Philipp Schwabl, Mattias Mandorfer, Simona<br />

Bota, Bernhard Scheiner, Wolfgang Sieghart, Arnulf Ferlitsch,<br />

Michael Trauner, Markus Peck-Radosavljevic, Thomas Reiberger;<br />

Div. of Gastroenterology and Hepatology, Department of Internal<br />

Medicine III, Medical University of Vienna, Vienna, Austria<br />

Background: Hepatorenal syndrome (HRS) is considered a<br />

severe and often lethal type of acute kidney injury (AKI) in<br />

patients with cirrhosis and ascites. Recently, the International<br />

Club of Ascites (ICA) proposed new criteria for HRS, loosening<br />

some of the stricter criteria in order to enhance diagnostic<br />

sensibility. However, HRS diagnosis still requires the exclusion<br />

of certain triggers of AKI to ensure “hepatic” etiology of AKI.<br />

Aims: To assess survival of patients with HRS according to the<br />

new ICA-HRS guidelines as compared with patients with “specifically-triggered”<br />

AKI. Methods: A cohort of 497 consecutive<br />

patients with cirrhosis and ascites was longitudinally screened<br />

for renal dysfunction. AKI and HRS were diagnosed according<br />

to new ICA criteria. Specific triggers for severe (grade 2/3) AKI<br />

episodes that are considered exclusion criteria for HRS were<br />

recorded (sAKI). Transplant-free survival (TFS) of HRS and sAKI<br />

was compared using log-rank test. Results: Among all patients,<br />

64 cases of HRS and 124 cases of sAKI were recorded. No<br />

differences were found regarding patient characteristics or TFS:<br />

Median TFS was 24 days in HRS (IQR 4-405 days) vs. 16 days<br />

(3-217) in sAKI (p=0.364). sAKI with manifest hypovolemia as<br />

trigger was associated with the shortest TFS [n=70; 8 (3-40)<br />

days], followed by preceding surgery or intervention [n=6;<br />

TFS: 32 (3-109) days] and infections as triggers [n=30; TFS:<br />

44 (2-187) days]. Conversely, patients with acute-on-chronic<br />

renal failure [n=7; TFS: 250 (14-end of follow-up) days] and<br />

nephrotoxic trigger [n=11; TFS: 217 (7-387) days] showed a<br />

favorable TFS. Conclusion: Patients with HRS according to the<br />

new ICA guidelines and sAKI showed similar characteristics<br />

and outcomes. However, since the specific cause of sAKI significantly<br />

influences prognosis, identifying the trigger of sAKI<br />

represents an important clinical priority in cirrhotic patients<br />

with ascites.<br />

Disclosures:<br />

Mattias Mandorfer - Consulting: Janssen; Speaking and Teaching: AbbVie, Gilead,<br />

Janssen, Boehringer Ingelheim, Bristol-Myers Squibb, Roche<br />

Simona Bota - Speaking and Teaching: Janssen Pharmaceutica, Bristol-Myers<br />

Squibb<br />

Wolfgang Sieghart - Grant/Research Support: Bayer Schering Pharma, Bayer<br />

Schering Pharma, Bayer Schering Pharma, Bayer Schering Pharma; Speaking<br />

and Teaching: Bayer Schering Pharma, Bayer Schering Pharma, Bayer Schering<br />

Pharma, Bayer Schering Pharma<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

The following authors have nothing to disclose: Theresa Bucsics, Philipp Schwabl,<br />

Bernhard Scheiner, Arnulf Ferlitsch


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 351A<br />

271<br />

Chronic Rifaximin Therapy Decreases the Risk and<br />

Severity of Hepatorenal Syndrome in Cirrhosis<br />

Tien S. Dong, Andrew I. Aronsohn, Nancy Reau, K. Gautham<br />

Reddy, Helen S. Te; Medicine, University of Chicago, Chicago, IL<br />

BACKGROUND: The intestinal bacterial flora plays an important<br />

and complex role in patients with cirrhosis. Studies in alcohol-related<br />

cirrhotic patients showed that rifaximin reduced the<br />

frequency of acute kidney injury (AKI) and hepatorenal syndrome<br />

(HRS). AIM: To determine the effect of long-term rifaximin<br />

use on the renal function of cirrhotic patients. METHODS:<br />

A retrospective chart review was performed to identify cirrhotic<br />

patients who had at least two visits in the Liver Clinic from<br />

1/1/11 to 12/31/14. The study group consisted of patients<br />

who were on rifaximin for ≥90 days, who were then matched<br />

by age, gender, and baseline MELD score to a control group.<br />

Patients with a diagnosis of hepatocellular carcinoma and<br />

renal replacement therapy (RRT) at baseline were excluded.<br />

Demographic and laboratory data, concurrent midodrine use,<br />

incidence of AKI, HRS, need for RRT, infections, frequency of<br />

hospitalizations and length of stay (LOS) were collected. AKI<br />

was defined as an increase in serum Cr (SCr) by >0.3 mg/<br />

dl within 48 hours or an increase in SCr >1.5 folds above<br />

baseline (International Society of Nephrology criteria). Data<br />

was censored at the last follow-up, initiation of RRT, death, or<br />

liver transplant, and analyzed using logistic regression, oneway<br />

ANOVA, and Pearson’s chi-squared test with the Stata<br />

software. RESULTS: 88 rifaximin cases were identified and<br />

matched to 88 control cases. Age, gender, race, duration of<br />

follow up, baseline MELD score, SCr, and GFR, and etiology<br />

of cirrhosis were not significantly different between the two<br />

groups. There were more patients on chronic midodrine (>90<br />

days use) in the rifaximin group as compared to control (17 %<br />

vs 4.5%, p= 0.008). There was no significant difference in the<br />

frequency of infections, deaths, liver transplants, hospitalizations<br />

or mean LOS between the two groups. Despite a similar<br />

frequency of AKI between the two groups, there was a trend<br />

towards less likelihood for HRS as a cause for AKI while controlling<br />

for midodrine use (OR 0.272 [0.067-1.086], p=0.06).<br />

In addition, the rifaximin group also had a lower likelihood of<br />

having at least one HRS episode (OR 0.257 [0.0696-0.950],<br />

p=0.04) and lower likelihood of requiring RRT (OR 0.257<br />

[0.067-0.950], p=0.04). The number needed to treat to prevent<br />

one episode of RRT is 10. CONCLUSIONS: Chronic use<br />

of rifaximin is associated with a decreased risk of HRS and a<br />

decrease likelihood of requiring RRT for AKI in a general population<br />

of cirrhotic patients. This finding should be validated in<br />

a prospective study in a larger patient population.<br />

Disclosures:<br />

Nancy Reau - Advisory Committees or Review Panels: Jannsen, Merck, AbbVie,<br />

Intercept, Salix, BMS, Jannsen; Grant/Research Support: Merck, Gilead, BMS,<br />

AbbVie, Jannsen, BI<br />

K. Gautham Reddy - Advisory Committees or Review Panels: AASLD Transplant<br />

Hepatology Pilot Steering Committee, ACG Training Committee, Program Director’s<br />

Caucus Steering Committee, Gilead, Inercept, Bristol-Myers Squibb; Grant/<br />

Research Support: Intercept, Ocera, Merck, Lumena<br />

Helen S. Te - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Abbvie, Conatus, BMS<br />

The following authors have nothing to disclose: Tien S. Dong, Andrew I. Aronsohn<br />

272<br />

Ascites Neutrophil Gelatinase-Associated Lipocalin<br />

(NGAL) Identifies Spontaneous Bacterial Peritonitis and<br />

Predicts Mortality in Hospitalized Patients with Cirrhosis<br />

Elizabeth C. Verna 1 , Giuseppe Cullaro 1 , Grace Kim 1 , Mona Gossmann<br />

1 , Thresiamma Lukose 1 , Connie Kim 2 , Sofia Nigar 3 , Marcus<br />

Pereira 1 , Robert S. Brown 1 ; 1 Columbia University Medical Center,<br />

New York, NY; 2 Stony Brook University School of Medicine, Stony<br />

Brook, NY; 3 The Brooklyn Hospital Center, Brooklyn, NY<br />

Background: Neutrophil gelatinase-associated lipocalin<br />

(NGAL) levels in the urine strongly predict acute kidney injury<br />

(AKI) and mortality in patients with cirrhosis. Ascites NGAL<br />

(aNGAL) levels may identify peritonitis in patients on peritoneal<br />

dialysis but have not been tested in cirrhosis. We hypothesized<br />

that aNGAL level is a marker of spontaneous bacterial<br />

peritonitis (SBP) and mortality risk in hospitalized patients with<br />

cirrhosis. Methods: Hospitalized patients with cirrhosis and<br />

ascites undergoing diagnostic paracentesis were prospectively<br />

enrolled and followed until death or discharge. Exclusion criteria<br />

included secondary peritonitis, history of transplantation<br />

and active colitis. NGAL was measured in the ascites, serum<br />

(sNGAL) and urine (uNGAL) by ELISA (Bioporto, Denmark).<br />

aNGAL level was evaluated as a marker of SBP (defined as<br />

ascites absolute neutrophil count >250) and predictor of inpatient<br />

mortality. Results: 79 patients were enrolled in this ongoing<br />

study. The median age was 58.4, 59% male gender, 53%<br />

non-Hispanic white, 21% Hispanic, and 11% black. 44% had<br />

HCV as their primary liver disease, 35% alcohol and 10%<br />

NAFLD. The median MELD was 20, 26% had AKI, defined as<br />

an increase of 0.3 mg/dL from baseline, and median (range)<br />

follow up was 11 (3-374) days. 16 patients (20%) had SBP.<br />

Baseline characteristics were similar between subjects with and<br />

without SBP. Median (IQR) aNGAL was significantly higher<br />

in patients with SBP compared to those without SBP [297.0<br />

(365.0) v. 155.0 (162.5) ng/mL, p= 0.008]. In ROC analysis,<br />

aNGAL had an AUC of 0.72 (95% CI 0.59-0.86). Median<br />

sNGAL (454.5 v. 409.7 ng/mL, p=0.41) and uNGAL (133.6<br />

v. 99.3 ng/mL, p=0.07) were similar between groups. aNGAL<br />

correlated relatively well with sNGAL (r=0.74) and less so with<br />

uNGAL (r=0.40). 8 (10%) patients died in the hospital. aNGAL<br />

(OR 1.02 per 10 units, p=0.02), MELD (1.13, p=0.02), SBP<br />

(OR 4.91, p=0.04) and bacteremia (OR 6.2, p=0.02) were<br />

predictive of in hospital mortality. In the final multivariable<br />

model, aNGAL (OR 1.03 per 10 units, p=0.01) remained<br />

predictive of in hospital mortality, controlling for SBP (OR 9.05,<br />

p=0.03) and AKI (8.82, p=0.04). Conclusions: aNGAL level<br />

may be a biomarker of peritonitis in hospitalized patients with<br />

ascites, and importantly, a predictor of short term in hospital<br />

mortality even controlling for SBP and AKI. Larger <strong>studies</strong> are<br />

needed to validate these findings.<br />

Disclosures:<br />

Elizabeth C. Verna - Advisory Committees or Review Panels: Gilead; Grant/<br />

Research Support: Salix, Merck<br />

Robert S. Brown - Consulting: Gilead, Janssen, Abbvie, Merck, BMS<br />

The following authors have nothing to disclose: Giuseppe Cullaro, Grace Kim,<br />

Mona Gossmann, Thresiamma Lukose, Connie Kim, Sofia Nigar, Marcus Pereira


352A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

273<br />

Use of non-selective beta blockers in cirrhotic patients<br />

with ascites is associated with increased risk of acute<br />

kidney injury at hospital admission<br />

Juliana R. de Carvalho, Rodrigo Luz, Henrique Sergio M. Coelho,<br />

Cristiane Villela-Nogueira, Renata M. Perez; Hepatology Service,<br />

Federal University of Rio de Janeiro, Rio de Janeiro, Brazil<br />

Introduction Use of non-selective beta blockers (NSBB) in<br />

patients with advanced cirrhosis has recently become an issue<br />

of debate, with concerns about their safety and possible association<br />

with complications such as acute kidney injury (AKI).<br />

Objective The aim of this study was to evaluate the association<br />

between use of NSBB and the presence of AKI in the first 48<br />

hours of hospital admission in cirrhotic patients with ascites.<br />

Methods Hospitalizations of cirrhotic patients with ascites in<br />

an university hospital were retrospectively reviewed. AKI at<br />

hospital admission was defined as a difference of at least 0.3<br />

mg/dL in two serum creatinine values obtained in the first 48<br />

hours of admission. Association between use of NSBB and AKI<br />

was evaluated by logistic regression analysis. Results This study<br />

included 328 hospitalizations of cirrhotic patients with ascites,<br />

with a mean age of 59 (SD: ±12.7) years. Men represented<br />

59% of cases. At admission, 133 patients (41%) were classified<br />

as Child B and 195 (59%) as Child C. History of previous<br />

gastrointestinal bleeding episode was present in 144 patients<br />

(44%) and 200 (61%) were on NSBB. AKI at hospital admission<br />

occurred in 196 patients (60%), being present in 66% of<br />

patients on use of NSBB and 50% of those not on use of the<br />

drug (p=0.004). In Child B patients, AKI occurred in 61% of<br />

those on NSBB and 51% of those without the drug (p=0.24).<br />

In Child C patients, a significant association between NSBB<br />

use and AKI was observed (69% vs 49%; p = 0.006). The<br />

variables age, gender, previous episode of gastrointestinal<br />

bleeding, previous episode of spontaneous bacterial peritonitis,<br />

history of diabetes mellitus, history of arterial hypertension,<br />

use of diuretics, use of NSBB, Child-Pugh score at admission<br />

and serum sodium at admission were included in a logistic<br />

regression model. In the final model, variables independently<br />

associated to AKI in the first 48 hours of admission were: age<br />

(OR: 1.03; 95% CI: 1.00 – 1.05; p=0.004), serum sodium at<br />

admission (OR: 0.95; 95% CI: 0.92 – 0.99; p=0.014) and<br />

use of NSBB (OR: 1.91; 95% CI: 1.20 – 3.06; p=0.007). Conclusion<br />

In Chid C cirrhotic patients with ascites, use of NSBB<br />

is associated with increased risk of AKI at hospital admission.<br />

Use of these drugs in this population should be reassessed.<br />

Disclosures:<br />

The following authors have nothing to disclose: Juliana R. de Carvalho, Rodrigo<br />

Luz, Henrique Sergio M. Coelho, Cristiane Villela-Nogueira, Renata M. Perez<br />

274<br />

Slow, continuous low-dose albumin and furosemide<br />

infusion (SCLAFI) mobilizes large ascites safely in<br />

decompensated liver cirrhosis<br />

Gaurav Pande; gastroenterology, Sanjay gandhi post graduate<br />

institute of medical sciences, Lucknow, India<br />

Background: Graded increase of oral diuretics has been the<br />

standard therapy for mobilizing large ascites in decompensated<br />

liver cirrhosis. Large volume paracentesis (LVP) with or without<br />

albumin infusions has been the commonest rescue therapy.<br />

Both are sought with complications. Aim: To study the efficacy<br />

and safety of low-dose, continuous, infusionof furosemide with<br />

albumin, administered according to a response-guided protocol<br />

in patients with cirrhosis and large ascites. Methods: All cirrhotic<br />

patients with large ascites on maximally tolerated dirutics<br />

and creatinine


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 353A<br />

SC and UO criteria. We defined oliguria as transient if patient<br />

had UO stage 1 and persistent if patient had UO stage 2 or<br />

3. Of the 3458 patients with CLD included in the study, 2997<br />

(86.7%) had cirrhosis. AKI occurred in 2854 (82.5%) patients.<br />

Hospital mortality for patient with AKI vs no AKI was 21.2% vs<br />

5% (p


354A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 355A<br />

failure of baseline mean arterial pressure to increase by 10<br />

mm Hg. Central venous pressure was checked every four hours<br />

with a goal of 4-10 mm Hg. Additional intravenous albumin<br />

(25 g) was given if goals were not met and furosemide (120<br />

mg) was administered when goals were exceeded. Progressive<br />

improvement in serum Cr led to weaning of NE. Patients were<br />

considered treatment failures if serum Cr continued to increase<br />

on protocol or renal replacement therapy was initiated. Results:<br />

A total of 42 patients were included. Prior to NE infusion, the<br />

median creatinine was 3.1 mg/dL and the median MELD score<br />

was 33 points. The median albumin dose was 93.5 g prior to<br />

NE therapy and the median duration of therapy was 6.5 days.<br />

New cardiac ischemia or arrhythmia developed in 4/42 (9%)<br />

of treated patients resulting in cessation of NE therapy in 2 of<br />

those patients. Improvement in renal function without need for<br />

renal replacement therapy was seen in 20/42 (48%) patients.<br />

The median Cr in responders went from 3.45 to 2.4 mg/dL. Of<br />

the responders, 15/20 (75%) were discharged from the hospital<br />

(N=8) or successfully bridged to liver transplantation (N=7),<br />

while 5/20 (25%) died or were transitioned to hospice for<br />

reasons not related to worsening renal function. Conclusions:<br />

NE infusion plus intravenous albumin appears to be a relatively<br />

safe and effective therapy for type 1 HRS. In this analysis, we<br />

were able to improve or stabilize renal function to the point of<br />

liver transplantation or discharge from the hospital in 48% of<br />

patients meeting strict criteria for type 1 HRS, a disorder historically<br />

associated with a median survival of about 2 weeks.<br />

Disclosures:<br />

The following authors have nothing to disclose: Crit T. Richardson, Michael K.<br />

Porayko<br />

280<br />

Clinical Utilization of Serum Albumin to Ascitic Fluid<br />

Albumin Gradient and Ascitic Fluid Protein Concentration<br />

in Pediatric Ascites<br />

Saachi Nangia 1 , Thammasin Ingviya 3 , Pavis Laengvejkal 2 , Ann<br />

O. Scheimann 1 , Wikrom Karnsakul 1 , Hejab Imteyaz 4 , Alexandra<br />

Vasilescu 4 ; 1 Pediatrics, Johns Hopkins University School of Medicine,<br />

Baltimore, MD; 2 Neurology, Texas Tech University Health<br />

Sciences Center, Lubbock, TX; 3 Environmental Health Sciences,<br />

Bloomberg School of Public Health, Baltimore, MD; 4 Pediatric Gastroenterology,<br />

Ft Myers, FL<br />

Background: Abdominal paracentesis and ascitic fluid (AF)<br />

analysis is performed in children with ascites to rule out infection<br />

or cancer. Serum albumin to ascitic fluid albumin gradient<br />

(SAAG) is frequently used in adults to distinguish portal<br />

hypertension (PHT) when SAAG is ≥1.1g/dL. High AF protein<br />

(>2.5 g/dL) is reported in cases with congestive heart failure<br />

or congestive hepatopathy (CH), hepatic vein outflow obstruction<br />

(HVOO) while low AF protein ( 2.5<br />

g/dL; uniquely presenting with intrahepatic mass and reactive<br />

mesothelial cells on AF cytopathology. Of 7 HVOO&CH<br />

cases only a BCS case with hepatic GVHD and liver fibrosis<br />

had AF protein < 2.5 g/dL. An idiopathic case with SAAG of<br />

0.6 and AF protein 2.3 g/dL led to suspect cirrhosis (APRI ><br />

1). None of 21 cases had definite SBP. Conclusion: Our data<br />

showed consistency with existing literature that cases with PHT<br />

either from cirrhosis, HVOO, or CH had SAAG ≥1.1 g/dL<br />

except for one case with peritoneal dialysis. SAAG < 1.1 g/dL<br />

was observed in cases with NS, serositis, PLE and peritoneal<br />

metastasis. AF protein has its value to reconfirm PHT types and<br />

potential causes of ascites along with SAAG and APRI values.<br />

Disclosures:<br />

The following authors have nothing to disclose: Saachi Nangia, Thammasin<br />

Ingviya, Pavis Laengvejkal, Ann O. Scheimann, Wikrom Karnsakul, Hejab<br />

Imteyaz, Alexandra Vasilescu<br />

281<br />

Multicenter retrospective analysis of the Model for Endstage<br />

Liver Disease (MELD) and proton pump inhibitors<br />

to predict in the first episode of spontaneous bacterial<br />

peritonitis<br />

Sakkarin Chirapongsathorn 1,2 , Rashid Khan 3 , Ashwani K. Singal 4 ,<br />

Patrick S. Kamath 1 ; 1 Gastroenterology and Hepatology, Mayo<br />

Clinic, Rochester, Rochester, MN; 2 Gastroenterology, Phramongkutklao<br />

Hospital and College of Medicine, Royal Thai Army, Bangkok,<br />

Thailand; 3 Gastroenterology and Hepatology, The University<br />

of Texas Medical Branch (UTMB) in Galveston, Galveston, TX;<br />

4 Gastroenterology and Hepatology, University of Alabama, Birmingham,<br />

AL<br />

Background: Spontaneous bacterial peritonitis (SBP) is a common<br />

infection in patients with cirrhosis. The Model for Endstage<br />

Liver Disease (MELD) is a prospectively developed and<br />

validated chronic liver disease severity scoring system to predict<br />

survival. At present, there are no data on its predictive ability<br />

on the first episode of SBP occurrence. Recent <strong>studies</strong> show<br />

controversial results regarding an increased risk of developing<br />

SBP in cirrhotic patients on proton pump inhibitors (PPI). The<br />

aim of the study was to investigate the relationship between<br />

MELD score and PPI therapy and the development of the first<br />

episode of SBP in cirrhotic patients in a multicenter study. Methods:<br />

A retrospective case-control study was performed. Seven<br />

hundred fifteen patients with a diagnosis of cirrhosis were identified<br />

from 3 tertiary referral centers. All medical records were<br />

manually reviewed by investigators to confirm the diagnosis<br />

of cirrhosis and SBP. Cirrhotic patients with a first episode<br />

of SBP (n=247 case) were matched to cirrhotics with ascites<br />

but without SBP (n=468 control). Every case was matched to<br />

2 controls for age, gender and race within each institution.<br />

Baseline MELD score was calculated within a month prior<br />

to SBP diagnosis for cases and within a month prior to any<br />

decompensation for controls. The time point for PPI usage was<br />

90 days before SBP diagnosis. Results: Proportion of patients<br />

with SBP increased with increasing MELD scores. For each unit<br />

increase in MELD, the odds ratio (OR; 95% CI) for development<br />

of SBP was 1.18 (1.14-1.21). Among patients with MELD score


356A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

greater than 10, 20 and 30, OR for development of SBP was<br />

4.62 (2.86-7.82), 7.9 (5.18-12.27) and 11.03 (4.92-28.1)<br />

respectively. MELD score cut-off level ≥ 17 had an area under<br />

the receiver operating characteristic curves (AUROC) 0.76 with<br />

sensitivity 63% and specificity 79%. PPI use was independently<br />

associated with development of SBP (OR 2.87, CI 1.78-4.69).<br />

There was an interaction between MELD score and PPI use, but<br />

MELD score was a stronger risk. Conclusion: MELD score and<br />

PPI therapy are associated with development of the first episode<br />

of SBP. Prospective <strong>studies</strong> are needed to validate these<br />

data and to determine the optimal MELD score above which<br />

patients would benefit from primary prophylaxis of SBP.<br />

18 and 24 months. Overall the included <strong>studies</strong> had a low risk<br />

of bias, except for two observational <strong>studies</strong> in which the risk<br />

was judged as moderate. Conclusions: Based on moderate<br />

quality evidence, the use of NSBBs was not associated with a<br />

significant increase in the risk of all-cause mortality in cirrhotic<br />

patients with ascites, including refractory ascites.<br />

Subgroup analyses of 12 <strong>studies</strong> included in meta-analysis<br />

Multivariate Logistic regression analyses of predictors the first<br />

episode of SBP<br />

Disclosures:<br />

Patrick S. Kamath - Advisory Committees or Review Panels: Sequana Medical<br />

The following authors have nothing to disclose: Sakkarin Chirapongsathorn,<br />

Rashid Khan, Ashwani K. Singal<br />

282<br />

Effect of Beta-Blockers on Survival in Patients with Cirrhosis<br />

and Ascites: A Systematic Review and Meta-analysis<br />

Sakkarin Chirapongsathorn 1,2 , Nelson Valentin 1 , Fares Alahdab<br />

3,4 , Chayakrit Krittanawong 5 , Patricia J. Erwin 6 , Mohammad<br />

H. Murad 3,7 , Patrick S. Kamath 1 ; 1 Gastroenterology and Hepatology,<br />

Mayo Clinic, Rochester, Rochester, MN; 2 Gastroenterology,<br />

Phramongkutklao Hospital and College of Medicine, Royal Thai<br />

Army, Bangkok, Thailand; 3 Robert D and Patricia E Kern Center<br />

for the Science of Health Care Delivery, Mayo Clinic, Rochester,<br />

Rochester, MN; 4 Knowledge and Evaluation Research Unit, Mayo<br />

Clinic, Rochester, Rochester, MN; 5 Cardiovascular Disease, Mayo<br />

Clinic, Rochester, Rochester, MN; 6 Mayo Clinic Libraries, Mayo<br />

Clinic, Rochester, Rochester, MN; 7 Preventive, Occupational and<br />

Aerospace Medicine, Mayo Clinic, Rochester, Rochester, MN<br />

Backgrounds: Recent <strong>studies</strong> have suggested that non-selective<br />

beta-blockers (NSBBs) are associated with negative impact<br />

on survival in patients with cirrhosis and refractory ascites.<br />

However, other <strong>studies</strong> showed contradictory findings and suggested<br />

that NSBBs are beneficial in these patients. We performed<br />

a systematic review and meta-analysis to evaluate the<br />

effect of NSBBs on all-cause mortality. Methods: We conducted<br />

a comprehensive search of MEDLINE, Embase, Web of Science<br />

and Scopus databases through January 2015 and manually<br />

reviewed the literature. Attempts were made to contact<br />

the corresponding authors of the relevant <strong>studies</strong> for additional<br />

information when needed. Trial-specific risk ratios (RRs) were<br />

calculated and pooled using random-effect model meta-analysis.<br />

Between-study heterogeneity was assessed using the I 2<br />

statistic. The quality assessment of included <strong>studies</strong> and publication<br />

bias were assessed. Result: We included 4 randomized<br />

control trials and 8 observational <strong>studies</strong> in which 1,206 deaths<br />

were reported in 2,486 patients with ascites. NSBBs users did<br />

not have an increase in all-cause mortality compared to NSBBs<br />

nonusers with overall ascites (RR, 0.94; 95% CI, 0.6-1.46; P<br />

= 0.77); non-refractory ascites (RR, 0.97; 95% CI, 0.45-2.09)<br />

or refractory ascites (RR, 0.86; 95% CI: 0.47-1.57; P = 0.63).<br />

Results were similar in RCTs and observational <strong>studies</strong>. Use of<br />

NSBBs was not associated with increased mortality at 6, 12,<br />

Disclosures:<br />

Patrick S. Kamath - Advisory Committees or Review Panels: Sequana Medical<br />

The following authors have nothing to disclose: Sakkarin Chirapongsathorn,<br />

Nelson Valentin, Fares Alahdab, Chayakrit Krittanawong, Patricia J. Erwin,<br />

Mohammad H. Murad<br />

283<br />

Increasing incidence and cost, but decreasing mortality<br />

in patients with hepatorenal syndrome: a study of the<br />

National Inpatient Sample 2005-2011<br />

Albert Do 1 , Ghideon Ezaz 2 ; 1 Internal Medicine, Yale-New Haven<br />

Hospital, New Haven, CT; 2 Internal Medicine, Beth Israel Deaconess<br />

Medical Center, Boston, MA<br />

Background: Hepatorenal syndrome (HRS) is significant in<br />

patients with cirrhosis, carrying a high morbidity and mortality<br />

risk. However, its burden on inpatient hospital care presently<br />

is unclear. We report the incidence, costs, and mortality<br />

trends of HRS in the American medical inpatient setting. Materials<br />

and Methods: The Nationwide Inpatient Sample contains<br />

discharge-level inpatient hospitalizations comprising 20% of<br />

non-government hospitals in the United States. We determined<br />

incidence rates, mortality rates, hospital length of stay, and<br />

hospital costs (adjusted to 2011 dollar terms to adjust for inflation)<br />

from 2005 to 2011 for patients with HRS (ICD-9 code<br />

572.4). Results: 153,057 cases of HRS were identified during<br />

2005-2011, comprising a cohort with mean age 57.6 years,<br />

primarily males (64.1%) of white race (55.9%), with public<br />

insurance payers (58.1%), and hospitalized at a teaching hospital<br />

(54.9%). Table 1 summarizes HRS-associated trends in<br />

incidence, mortality and cost, showing increased disease incidence.<br />

Overall hospital mortality was 34.4%, with decreasing<br />

mortality (43.8% in 2005 to 31.2% in 2011). $16.3 billion<br />

was spent in this time, with an annual expenditure increase<br />

from $1.4 billion to $3.5 billion. However, hospital length<br />

of stay did not change during this time period (range 10.1 to<br />

11.9 days). 5,268 patients (3.4% of all patients) received a<br />

liver transplant during their hospital stay. Conclusions: From<br />

2005 to 2011, the incidence and costs of HRS have increased.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 357A<br />

However, associated mortality has decreased and hospital<br />

length-of-stay has not changed. Further <strong>studies</strong> are needed to<br />

assess the reasons for decreased mortality, as well as measures<br />

to decrease costs associated with hospital stay.<br />

Table 1. Incidence, mortality, and costs of patients with hepatorenal<br />

syndrome, NIS sample 2005-2011<br />

had the highest bilirubin (27mg/dl) compared to ATN alone<br />

(18 mg/dl) p


358A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

is similar to previous reports. Guidelines recommend a paracentesis<br />

to evaluate for SBP in patients admitted with ascites,<br />

but do not address evaluation for SBP in outpatients. In outpatients,<br />

the need for routine ascitic fluid analysis in asymptomatic<br />

patients is an area of discussion based on a reported low<br />

prevalence of SBP. This study suggests that similar to inpatients,<br />

all asymptomatic outpatients undergoing paracentesis should<br />

routinely have ascitic fluid analyzed to assess for SBP.<br />

Disclosures:<br />

Maria Sjogren - Speaking and Teaching: Gilead, Bristol Myers<br />

The following authors have nothing to disclose: Dustin Albert, Manish B. Singla,<br />

Frank Cheng, Dawn Torres<br />

286<br />

Is Midodrine, Octreotide, and Albumin still worthwhile<br />

therapy for Hepatorenal Syndrome Type 1 (HRS1)?<br />

Katherine Hahn 1 , Amen Javaid 1 , Pooja Singhal 2 , Franca B. Barton<br />

3 , James H. Lewis 2 ; 1 Internal Medicine, Georgetown University<br />

Hospital, Washington, DC; 2 Georgetown University, Washington,<br />

DC; 3 The EMMES Corporation, Rockville, MD<br />

Introduction: HRS1 is a fatal complication of advanced cirrhosis,<br />

with mortality rates as high as 80% within 2 weeks<br />

of diagnosis. Current treatment options in the United States<br />

include vasoconstrictors (e.g.: midodrine [M], octreotide [O],<br />

and vasopressin), often combined with albumin [A], transjugular<br />

intrahepatic portosystemic shunt (TIPS), renal replacement<br />

therapy (RRT), and liver transplantation (LT). This retrospective<br />

study analyzes the outcomes of 137 HRS1 patients treated<br />

with MOA as primary therapy or as a bridge to LT. Methods: A<br />

retrospective cohort study evaluated HRS1 patients treated with<br />

MOA at Georgetown University Hospital over a 6 year period<br />

(Jan 2009 – Dec 2014). The demographics, underlying liver<br />

disease, creatinine levels, MELD scores, RRT requirements, and<br />

LT rates of these MOA patients were collected. The primary end<br />

point was survival to hospital discharge. Secondary outcomes<br />

included survival after RRT and survival with/without LT. The<br />

effects of patient characteristics on survival were tested with a<br />

Cox proportional hazards model. The relationship between LT<br />

and survival was analyzed using logistic regression. Results:<br />

From 2009-2014, a total of 238 patients were treated with<br />

MOA. Of these, 101 patients did not meet the diagnostic<br />

criteria for HRS1 and were excluded. Of the remaining<br />

137 patients, 70 (51%) had SBP, 73 (52.9%) received RRT,<br />

but only 34 (24.8%) met criteria to undergo LT. The overall<br />

survival to discharge was 33%. Among those who survived,<br />

57.1% had received LT and 42.8% had not. Factors showing<br />

a significant effect on survival were LT (p=0.015), MELD score<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 359A<br />

288<br />

Acute kidney injury in cirrhotic patients undergoing contrast-enhanced<br />

computed tomography<br />

Roberto Filomia 1 , Sergio Maimone 1 , Carlo Saitta 1 , Luca Visconti 5 ,<br />

Simona Caloggero 4 , Antonio Bottari 4 , Alessia Pitrone 4 , Angela<br />

Alibrandi 3 , Irene Cacciola 1 , Gaia Caccamo 1 , Domenica Vadalà 1 ,<br />

Carmine G. Gambino 1 , Maria Stella Franzè 1 , Giovanni Raimondo<br />

1 , Giovanni Squadrito 1,2 ; 1 Clinical and Molecular Hepatology,<br />

University Messina, Messina, Italy; 2 Human Pathology,<br />

University of Messina, Messina, Italy; 3 Department of Economical,<br />

Business and Environmental Sciences and Quantitative Methods,<br />

Univeristy of Messina, Messina, Italy; 4 Department of Biomedical<br />

Sciences and Morphological and Functional Imaging, University of<br />

Messina, Messina, Italy; 5 Internal Medicine, Unit Of Nephrology,<br />

Messina, Italy<br />

Introduction: Contrast-induced acute kidney injury (CI-AKI)<br />

following intravenous contrast medium administration is one<br />

of the most common causes of hospital-acquired acute kidney<br />

damage. Aim: To investigate the incidence and possible<br />

predisposing factors of CI-AKI in cirrhotic patients undergoing<br />

contrast-enhanced computed tomography (CECT). Patients<br />

and Methods: We analysed 393 consecutively hospitalized<br />

cirrhotic patients none of whom had active bacterial infection,<br />

glomerular filtration rate (GFR)


360A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

pressure (mPAP)


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 361A<br />

was observed between the two groups of patients as far as<br />

demographic, clinical, hemodynamic, and laboratory characteristics.<br />

No treatment was provided for Grade 1 ascites in<br />

the study. During the follow up twenty-two patients developed<br />

Grade 2 or 3 ascites, 14 in Group A (14,7 %) and 8 in Group<br />

B (25.0%, p = N.S.). In only 4 out of 14 patients of Group A,<br />

Grade 1 ascites was detected before they developed Grade 2<br />

or 3 ascites. The mean time for the development of Grade 2<br />

or 3 ascites was shorter in Group B than in Group A, but the<br />

difference was not statistically significant (17,00 ± 11,06 versus<br />

32,29 ± 27,59 months, p = N.S.). In nineteen patients in<br />

Group B (59.3%), Grade 1 ascites disappeared during the follow<br />

up. At univariate analysis, the finding of Grade 1 ascites,<br />

MELD, spleen diameter and the ratio platelets/spleen diameter<br />

were found as potential predictors of Grade 2 or 3 (Table). At<br />

multivariate analysis only the ratio platelets/spleen diameter<br />

was found to be the only independent predictor of Grade 2 or<br />

3 ascites (HR = 0,997, 95%, CI : 0,994; 0,999, p = 0,0018<br />

per el/mm 3 /mm of increase). In conclusion, the presence of<br />

Grade 1 ascites is neither a precursor nor an independent<br />

predictor of Grade 2 or 3 ascites in patients with cirrhosis.<br />

Thus, Grade 1 ascites does not require any treatment, including<br />

sodium dietary restriction.<br />

Disclosures:<br />

Paolo Angeli - Advisory Committees or Review Panels: Sequana Medical<br />

The following authors have nothing to disclose: Marta Tonon, Salvatore Piano,<br />

Matjaz Grbec, Chiara Pilutti, Silvano Fasolato, Antonietta Romano, Sara Montagnese,<br />

Filippo Morando, Massimo Bolognesi, Marialuisa Stanco, Giancarlo<br />

Bombonato, Silvia Rosi, Elisabetta Gola, Antonietta Sticca<br />

293<br />

Validation of Ascites-specific Patient Reported Outcome<br />

Questionnaires for Patients with Cirrhosis<br />

Patrick S. Kamath 3 , Myrte Neijenhuis 3 , Tom J. Gevers 2 , Thomas<br />

Atwell 1 , Joost Drenth 2 , Tim Gunneson 3 ; 1 Division of Gastroenterology,<br />

Hepatology & Internal Medicine, Mayo Clinic College of<br />

Medicine, Rochester, MN; 2 Gastroenterlogy and Hepatology,<br />

Radboud University Nijmegen Medical Centre, Nijmeten, Netherlands;<br />

3 Division of Gastroenterology and Hepatology, Mayo<br />

Clinic, Rochester, MN<br />

Background and aim: Since no treatment improves long-term<br />

survival, treatment of ascites should focus on symptom relief<br />

and prevention of complications. No patient reported outcomes<br />

have been used as endpoints in clinical trials in cirrhotic ascites.<br />

We aimed to identify a sensitive tool to measure symptoms<br />

in cirrhotic ascites. Methods: We identified 12 ascites-specific<br />

symptoms (fullness, lack of appetite, early satiety, nausea,<br />

abdominal pain, back pain, dyspnea, limited mobility, fatigue,<br />

insomnia, dissatisfaction with size of abdomen and sexual intimacy<br />

problems) based on literature, patient interviews (n=8)<br />

and clinician survey (n=8) to develop an ascites-specific questionnaire<br />

(Ascites-Q). We administered Ascites-Q; Japanese<br />

Ascites Symptom Inventory-7 (ASI-7, developed for cirrhotic<br />

ascites); FACIT-ascites index (developed for malignant ascites);<br />

and quality of life (QoL) VAS-scale in cirrhotic patients undergoing<br />

large volume paracentesis (target sample size n=90).<br />

Scores were transformed to a 0-100 scale with higher scores<br />

indicating more symptoms. To identify ascites-specific symptoms,<br />

we compared using the Mann Whitney U test scores<br />

of patients with refractory ascites prior to paracentesis with<br />

scores in cirrhotic patients with no ascites recruited at the Mayo<br />

Clinic between January - May 2015. Correlations of QoL VASscale<br />

and ascites questionnaires were calculated (Spearman).<br />

Responsiveness of questionnaires to detect changes in symptoms<br />

was assessed by comparing scores at baseline and 7<br />

days post paracentesis (paired T-test). Results: We included 32<br />

patients with refractory ascites requiring large-volume paracentesis<br />

(59% male, mean age 61 yr, MELD-score 15 and median<br />

paracentesis volume 3600 mL; and 61 patients without refractory<br />

ascites (controls) (76% male, mean age 57 yr, MELD-score<br />

11). Patients with refractory ascites scored higher than controls<br />

on all symptoms of Ascites-Q and ASI-7, except for Ascites-Q<br />

symptoms back pain (p=0.096) and fatigue (p=0.228). There<br />

was no difference between patients and controls in 6/13 FAC-<br />

IT-ascites index symptoms (insomnia, limited mobility, nausea,<br />

vomiting, polyuria and emotional distress). QoL correlated with<br />

Ascites-Q and ASI-7 (r=.497, p=0.008 and 0.471, p=0.013),<br />

but not with FACIT-ascites index (r=0.332, p=0.090). Scores<br />

of Ascites-Q, (-11.53 ± 18.874, p=0.013), ASI-7 (-35.46, ±<br />

29.22, p


362A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2 BA, 2 HVOO and 1 CH and HE diagnosed in 10 BA and<br />

7 HVOO. HRS and HE therefore were not included in logistic<br />

regression analysis as a result of samll sample size. Conclusion:<br />

BA group was younger and had higher propability of developing<br />

significant ascites and splenomegaly. Variceal bleeding<br />

was a complication in BA group presenting in a younger age<br />

range. Longitudinal study is needed for more detailed analysis<br />

to understand the natural history of children with PHT from<br />

diffrernet groups. -<br />

Disclosures:<br />

The following authors have nothing to disclose: Thammasin Ingviya, Ann O.<br />

Scheimann, Wikrom Karnsakul<br />

295<br />

CA-125 significance in cirrhosis and its correlation with<br />

disease severity<br />

Raja G. Reddy Edula, Yucai Wang, Serban A. Moroianu, Vivek<br />

A. Lingiah, Phoenix Fung, Maliha Ahmad, Lizza Bojito-Marrero,<br />

Nneoma O. Okoronkwo, Nikolaos Pyrsopoulos; Medicine, Rutgers<br />

New Jersey Medical School, Newark, NJ<br />

Background: Serum cancer antigen 125(CA125) is a tumor<br />

marker traditionally used in the diagnosis and monitoring of<br />

ovarian cancer. However it is reported to be elevated in a<br />

variety of non-neoplastic conditions like cirrhosis and congestive<br />

heart failure. Its prevalence, degree of elevation, correlation<br />

with severity of disease based on objective markers like<br />

MELD score, Child’s-Pugh classification and degree of ascites<br />

in cirrhosis is unknown. Aim: To evaluate the prevalence<br />

and significance of elevated CA125 in patients with cirrhosis<br />

and its correlation with objective markers of disease severity.<br />

Methods: This is a prospective study of 172 adult patients with<br />

cirrhosis due to any etiology after obtaining CA125 serum<br />

analysis at the time of enrollment. Demographics, etiology of<br />

cirrhosis, MELD score, Child’s-Pugh classification, degree of<br />

ascites, serum CA125 and other variables were collected. Statistical<br />

analysis including descriptive statistics, linear regression,<br />

one-way ANOVA and Student’s t-test was performed.<br />

Results: Etiology of cirrhosis is outlined in Table-1. Elevated<br />

CA125 levels were noted in 147(85%) of the study population<br />

with a mean value of 379. No difference was observed in<br />

elevated mean CA125 between male(411) and female(324)<br />

patients(P=0.207). Patients with higher MELD had higher<br />

CA125 level (P= 0.001). Subjects without ascites (N=35) had<br />

a mean CA125 level of 36, mild ascites(N=29): 218 and moderate<br />

ascites(N=108): 534(P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 363A<br />

297<br />

Impact of SNP rs2187668 in HLA-DQA1 on clinical<br />

presentation and treatment response in patients with<br />

type-1 autoimmune hepatitis<br />

Albert Stättermayer 1 , Michael Eder 1 , Sandra Beinhardt 1 , Karin<br />

Kozbial 1 , Clarissa Freissmuth 1 , Friedrich Wrba 2 , Michael Trauner 1 ,<br />

Peter Ferenci 1 , Harald Hofer 1 ; 1 Gastroenterology & Hepatology,<br />

Medical University of Vienna, Vienna, Austria; 2 Clinical Pathology,<br />

Medical University of Vienna, Vienna, Austria<br />

Background/aim: Autoimmune hepatitis (AIH) is a chronic<br />

inflammatory liver disease with limited understanding of etiopathogenesis.<br />

Recently, a genome-wide association study<br />

(GWAS) identified a single nucleotide polymorphism (SNP,<br />

rs2187668) in the HLA-DQA1 gene, which is strongly associated<br />

with AIH type-1. We evaluated the impact of this SNP on<br />

clinical presentation and treatment response in patients with<br />

type-1 AIH. Methods: One-hundred and seven patients with<br />

type-1 AIH (female: 83 [77.6%], mean age: 43.2 [CI95%:<br />

39.9-46.6] years), who reached a probable or definite score<br />

according to the simplified or revised scoring system were evaluated.<br />

SNP rs2187668 was investigated by real-time-PCR in<br />

all patients with AIH and in a control group of 97 healthy<br />

subjects. Results: Sixty-one (57.0%) patients with AIH had<br />

rs2187668 genotype GG, 39 (36.4%) were heterozygous<br />

and 7 (6.5%) had genotype AA. Minor allele frequency (MAF)<br />

of the risk allele (A) was 24.8% and was significantly higher<br />

than in the control group (MAF: 7.2%; p150 vs.


364A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

minimal and discordant, the aims of our study were to: 1)<br />

investigate HPSE expression in HSCs during their activation;<br />

2) evaluate HPSE activity in the plasma of patients with liver<br />

fibrosis. Methods. HSCs were isolated from rat liver and cultured<br />

on plastic to induce activation. HPSE expression, together<br />

with markers of HSCs transdifferentiation, were evaluated by<br />

real-time PCR at different days after isolation. HPSE protein<br />

amount was evaluated by immunofluorescence (IF). Utilizing<br />

an ELISA method, HPSE activity was measured in the plasma<br />

from 15 healthy subjects and 45 patients diagnosed with autoimmune<br />

chronic liver diseases (15 AIH, 17 PBC, 13 PSC);<br />

HPSE activity was correlated with fibrosis staging assessed by<br />

elastography. Results. After 15 days on plastic, HSCs resulted<br />

activated with the increase of alpha-SMA, fibronectin and TGFbeta<br />

expression with respect to 7 days cultured cells. Upon activation,<br />

HPSE mRNA was significantly 1.94-fold overexpressed<br />

in 15 days vs 7 days cultured cells. HPSE increase was confirmed<br />

by IF. HPSE plasma activity was significantly higher in<br />

patients with mild, significant and severe fibrosis compared<br />

to the control group (0.32±0.04, 0.25±0.03, 0.21±0.03<br />

respectively vs 0.06±0.01, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 365A<br />

(0.05 ng/mL -0.16 ng/mL), (p=0.003 and p=0.02, respectively)).<br />

Following in vitro activation, PD-1 was up-regulated<br />

3.3 fold (p=0.0019) on T cells from AIH patients compared<br />

with HC (1.5 fold). Conclusion: Present study supports an<br />

increased PD-1 mediated immune-regulation in AIH patients.<br />

Furthermore our results support the hypothesis about the presence<br />

of exhausted T cells in autoimmune diseases.<br />

Disclosures:<br />

Henning Gronbaek - Grant/Research Support: Novartis, Ipsen<br />

The following authors have nothing to disclose: Kristian Aarslev, Anders Dige,<br />

Stinne Greisen, Martin Kreutzfeltd, Niels Jessen, Hendrik V. Vilstrup, Bent Deleuran<br />

302<br />

Autoimmune hepatitis type 1 in Dutch elderly versus<br />

younger patients<br />

Martine A. Baven-Pronk 1 , Joanne J. van Silfhout 1 , Aad P. van<br />

den Berg 2 , Henk R. van Buuren 3 , Carin M. Van Nieuwkerk 4 , Bart<br />

van Hoek 1 ; 1 Gastroenterology and Hepatology, Leiden University<br />

Medical Center, Leiden, Netherlands; 2 Gastroenterology and<br />

Hepatology, University Medical Center Groningen, Groningen,<br />

Netherlands; 3 Gastroenterology and Hepatology, Erasmus Medical<br />

Center, Rotterdam, Netherlands; 4 Gastroenterology and Hepatology,<br />

Free University Medical Center, Amsterdam, Netherlands<br />

Background and Aims: Studies on elderly patients with autoimmune<br />

hepatitis (AIH) are limited, of small sample size and<br />

conflicting when it comes to mode of presentation, treatment<br />

and treatment outcome. The aim of this study was to investigate<br />

differences in baseline characteristics, laboratory features,<br />

histology, concomitant autoimmune disease, treatment,<br />

treatment adverse effects and outcome in elderly patients with<br />

AIH type one (age > 60 years at presentation) compared to<br />

younger patients with AIH type one (age < 60 years at presentation)<br />

and compare them to existing data to provide more<br />

insight into AIH in the elderly patient. Patients and Methods:<br />

All patients from 4 academic centres with probable or definite<br />

AIH type I according to the International AIH Group criteria<br />

were included. As much data as possible was retrospectively<br />

retrieved by chart review. Primary endpoints were defined as<br />

remission, liver transplantation or liver related death. Secondary<br />

endpoints were defined as differences in biochemistry and<br />

serology, symptoms, mode of presentation, concurrent autoimmune<br />

diseases, initial and maintenance treatment regimens,<br />

number of switches of therapy, adverse effects of treatment,<br />

achievement of remission, episodes of loss of remission, number<br />

of relapses, cirrhosis at presentation and progression of cirrhosis.<br />

Results: A total of 359 patients were included, 286 (80%)<br />

younger patients, 73 (20%) elderly patients. The young group<br />

presented with significantly higher serum alanine aminotransaminase<br />

(p


366A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

304<br />

Carcinogenesis in autoimmune hepatitis type 1 patients<br />

with a long-term follow-up in Japan<br />

Teruko Arinaga-Hino 1 , Tatsuya Ide 1 , Ichiro Miyajima 1 , Kei Ogata 1 ,<br />

Reiichiro Kuwahara 1 , Keisuke Amano 1 , Kazunori Noguchi 2 , Kunihide<br />

Ishii 3 , Kunitaka Fukuizumi 4 , Koji Yonemoto 5 , Takuji Torimura<br />

1 ; 1 Division of Gastroenterology, Department of Medicine,<br />

Kurume Univ.School of Medicine, Kurume, Japan; 2 Omuta City<br />

Hospital, Omuta, Japan; 3 Asakura Medical Association Hospital,<br />

Asakura, Japan; 4 National Kyushu Medical Center, Fukuoka,<br />

Japan; 5 The Biostatistics Center, Kurume University School of Medicine,<br />

Kurume, Japan<br />

Backgrounds/Aims: The risk of carcinogenesis in autoimmune<br />

diseases is high and has been attributed to immunological<br />

abnormalities, the use of immunosuppressive agents, and/or<br />

chronic inflammation. The aim of the present study was to determine<br />

the incidence and risk of carcinogenesis in patients with<br />

autoimmune hepatitis (AIH) type 1 in a large-scale population<br />

with a long-term follow-up in Japan. Methods: Two hundred<br />

and fifty-six patients diagnosed with AIH between Apr. 1978<br />

and Mar. 2014 were enrolled (F/M = 226:30; mean age<br />

at diagnosis AIH, 57.7 yrs). They had no malignancy before<br />

or within 1 yr of being diagnosed with AIH. A person-year<br />

calculation was performed for AIH patients, and the numbers<br />

of expected events were clarified using data from “A Study<br />

of 28 Population-based Cancer Registries for the Monitoring<br />

of Cancer Incidence in Japan (MCIJ) Project (2010)” in order<br />

to determine the standard incident rate (SIR) of each type of<br />

malignancy. Biochemical data regarding carcinogenesis and<br />

its background factors were also examined using a proportional<br />

hazards model, adopting the Kaplan-Meier method. The<br />

present study was approved by the proper Ethics Committee.<br />

Results: The mean observation period was 8.2 yrs (The person<br />

years at risk were 2,107; Females; 1,878, Males; 229).<br />

Twenty-seven patients developed malignancy (F:M = 24:3).<br />

The mean observation period of carcinogenesis from AIH diagnosis<br />

was 10.3 yrs. The number of patients with hepatobiliary<br />

cancer (Ca) was 11 (hepatocellular carcinoma; 10, cholangiocarcinoma;<br />

1), gastric Ca; 3, pancreatic Ca; 2, oral/pharyngeal<br />

Ca; 2, colorectal Ca; 2, breast Ca; 1, lung Ca; 1, acute<br />

myeloid leukemia; 1, malignant lymphoma 1, and malignant<br />

melanoma; 1. The overall SIR for malignancies in AIH was<br />

significantly high at 2.04 (95% CI 1.34-2.96), and was particularly<br />

high among females at 2.49 (95% CI 1.60-3.71). SIR for<br />

hepatobiliary Ca was 14.14 (95% CI 7.05-25.30), and was<br />

markedly high for females at 21.83 (95% CI 10.45-40.16).<br />

SIR for oral/pharyngeal Ca was significantly high for females<br />

at 14.61 (95% CI 1.64-52.77). No significant differences<br />

were observed in SIR among the other types of malignancies<br />

examined. The risk factors for overall carcinogenesis at the<br />

diagnosis of AIH were identified as low levels of alanine aminotransferase<br />

(p=0.0053), a low platelet count (p=0.0087),<br />

and liver cirrhosis (p=0.0067). No significant differences were<br />

observed among the treatment regimens, such as prednisolone,<br />

azathioprine, and ursodeoxycholic acid. Conclusion: The<br />

risk of malignancies was generally high among AIH patients.<br />

The risk of carcinogenesis needs to be carefully considered in<br />

patients with AIH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Teruko Arinaga-Hino, Tatsuya<br />

Ide, Ichiro Miyajima, Kei Ogata, Reiichiro Kuwahara, Keisuke Amano, Kazunori<br />

Noguchi, Kunihide Ishii, Kunitaka Fukuizumi, Koji Yonemoto, Takuji Torimura<br />

305<br />

Immunity in HLA-DR4 expressing healthy and autoimmune<br />

hepatitis induced NOD mice<br />

Muhammed Yuksel 1,2 , Xiaoyan Xiao 1,3 , Kathie Béland 4 , Pascal<br />

Lapierre 4 , Fernando Alvarez 4 , Isabelle Colle 2 , Yun Ma 5 , Li Wen 1 ;<br />

1 Endocrinology, Yale university, New haven, CT; 2 Internal medicine,<br />

Ghent university hospital, Ghent, Belgium; 3 Department of<br />

Nephrology, Shangdong University, Shangdong,., China; 4 Division<br />

of Gastroenterology, Hepatology and Nutrition, Sainte-Justine<br />

University Hospital, Montreal, QC, Canada; 5 Institute of Liver <strong>studies</strong>,<br />

Kings College London, London, United Kingdom<br />

Autoimmune hepatitis (AIH) is a chronic progressive inflammatory<br />

liver disease, characterized by interface hepatitis, positivity<br />

for autoantibodies and hyper-gammaglobulinemia. There are<br />

two types of AIH, type-1 (AIH-1) being characterized by seropositivity<br />

for anti-smooth muscle and/or anti-nuclear (SMA/<br />

ANA), and type-2 (AIH-2) by positivity for anti-liver kidney microsomal<br />

type 1 (anti-LKM1) and/or anti-liver cytosol type 1<br />

(anti-LC1) autoantibodies. Cytochrome P4502D6 (CYP2D6)<br />

has been defined as the antigenic target of anti-LKM1 and<br />

formiminotransferase cyclodeaminase (FTCD) as the target<br />

for anti-LC1. The development of AIH is linked to the Human<br />

Leukocyte Antigen (HLA) alleles, -DR3 for AIH-1 and -DR7 for<br />

AIH-2. However, the role of HLA-DR4 has been controversial,<br />

being the second disease predisposition gene in adults with<br />

AIH-1 and protective gene in pediatric patients. In this study,<br />

we investigated the role of HLA-DR4 transgene, expressed on<br />

non-obese diabetic (NOD) mice (DR4), healthy (baseline) mice<br />

and also after the AIH induction by immunization with the DNA<br />

coding for human CYP2D6/FTCD fusion protein. Immunization<br />

of wild type (WT) NOD and DR4 mice leads to a sustained<br />

mild elevation of alanine aminotransferase (ALT), development<br />

of anti-LKM1/anti-LC1, and chronic immune cell infiltration,<br />

mild parenchymal fibrosis and some plasma cell infiltration.<br />

Although these parameters were more slightly pronounced in<br />

DR4 mice than in WT control mice, the pathogenic role of<br />

HLA-DR4 remains uncertain. However, livers from DR4 mice<br />

contained less regulatory T cell (Tregs) which had decreased<br />

PD-1 expression and a dysfunction to suppress alloreaction.<br />

We also showed that, T cells and antigen presenting cells,<br />

such as macrophages and dendritic cells, were already activated<br />

in non-immunized naïve (healthy) DR4 mice compared to<br />

naïve WT NOD. These results show that HLA-DR4 molecule is<br />

involved in the spontaneous activation of the immune system,<br />

facilitating the induction of AIH.<br />

Disclosures:<br />

Isabelle Colle - Advisory Committees or Review Panels: Janssen; Grant/Research<br />

Support: Bayer; Patent Held/Filed: Trombogenics; Speaking and Teaching: BMS<br />

The following authors have nothing to disclose: Muhammed Yuksel, Xiaoyan<br />

Xiao, Kathie Béland, Pascal Lapierre, Fernando Alvarez, Yun Ma, Li Wen<br />

306<br />

Autoimmune Hepatitis Disproportionately Affects People<br />

of Color<br />

Michele M. Tana 1,2 , Edward W. Holt 3 , Robert J. Wong 4 , Justin L.<br />

Sewell 1 , Mandana Khalili 1 , Jacquelyn J. Maher 1,2 ; 1 Department of<br />

Medicine, Division of Gastroenterology, University of California,<br />

San Francisco, San Francisco, CA; 2 UCSF Liver Center, University<br />

of California, San Francisco, San Francisco, CA; 3 Department of<br />

Transplantation, Division of Hepatology, California Pacific Medical<br />

Center, San Francisco, CA; 4 Department of Gastroenterology<br />

and Hepatology, Alameda Health System, Oakland, CA<br />

Autoimmune hepatitis (AIH) is an uncommon disease that<br />

affects patients of all ethnicities, and there have been reports<br />

that its natural history varies in patients of different ethnic back-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 367A<br />

grounds. There is insufficient evidence about the prevalence<br />

of AIH among people of color.We sought to determine the<br />

racial/ethnic distribution of AIH cases in an ethnically diverse,<br />

urban, county hospital system. We performed an observational<br />

cohort study. A consecutive sample of patients with a diagnosis<br />

of AIH or AIH-PBC overlap syndrome per AASLD guidelines<br />

from September 2013 to April 2015 was included in<br />

the study. Demographic, clinical, laboratory, radiographic,<br />

and histologic data were collected. The characteristics of the<br />

cohort were compared with the published data on the patient<br />

population served by the integrated hospital system using the<br />

chi-squared test. The characteristics of different ethnic groups<br />

were compared using univariate and multivariate analysis. 23<br />

patients with AIH (6 had overlap syndrome) were seen in the<br />

hepatology clinic over the 20-month period. AIH occurred in<br />

significantly fewer white patients than would be expected if the<br />

prevalence were equal among ethnic groups, given the demographics<br />

of this integrated hospital system (Table, p=0.03).<br />

The median age at diagnosis was 49 years (range 18-91) and<br />

18 (78%) were female. At diagnosis, the median total bilirubin<br />

was 3.3 mg/dL, ALT was 483 U/L, IgG was 1995 mg/dL, INR<br />

was 1.2, albumin was 3.8 g/dL, and platelet count was 177<br />

k/uL. 19 patients had complete liver biopsy results, and ten<br />

(53%) had fibrosis stage ≥2, with four (21%) having advanced<br />

(stage 3-4) fibrosis. When stratified by Latino ethnicity, Latinos<br />

did not differ significantly from non-Latinos in terms of age at<br />

diagnosis, baseline labs, or severity of liver fibrosis. A two-predictor<br />

model that adjusted for age found no significant association<br />

between Latino ethnicity and fibrosis stage. AIH was<br />

significantly more prevalent among people of color (especially<br />

Latinos) suggesting that AIH may not affect all ethnic groups<br />

equally. Although the disease characteristics did not appear<br />

to differ by ethnicity, this disparity may be related to genetic,<br />

socioeconomic, environmental, or comorbid medical factors.<br />

Disclosures:<br />

Mandana Khalili - Consulting: Gilead Sciences Inc; Grant/Research Support:<br />

Gilead Sciences INc, Bristal Myer Squibb<br />

Jacquelyn J. Maher - Consulting: Durect<br />

The following authors have nothing to disclose: Michele M. Tana, Edward W.<br />

Holt, Robert J. Wong, Justin L. Sewell<br />

307<br />

Outcomes and mortality of de novo autoimmune hepatitis<br />

after liver transplantation<br />

Aishwarya Kuchipudi 1 , Ryan Morton 3 , Sarah Y. Chan 3 , Dilip<br />

Moonka 2 , Kimberly Ann Brown 2 , Syed M. Jafri 2 ; 1 Internal Medicine,<br />

Henry Ford Hospital, Detroit, MI; 2 Gastroenterology, Henry<br />

Ford Hospital, Detroit, MI; 3 Wayne State University School of Medicine,<br />

Detroit, MI<br />

Aim: To evaluate the incidence and outcomes of de novo Autoimmune<br />

Hepatitis (AIH) in patients with history of Orthotopic<br />

Liver Transplantation (OLT). Methods: We performed a retrospective<br />

chart review of 1343 OLT patients from 2000-2015.<br />

We identified patients with histopathological evidence of de<br />

novo AIH. Patients with biopsy features consistent with AIH<br />

pre-OLT and those with Hepatitis C were excluded from analysis.<br />

For each patient with de novo AIH, we identified 2 controls<br />

wherever possible, matched for etiology of liver disease<br />

and date of transplant (+/- 18 months). We then compared<br />

the two groups on factors including intra-operative or post-operative<br />

complications, immunosuppression levels at specific<br />

time points, development of fibrosis/cirrhosis, progression<br />

to re-transplant and mortality at 1, 3 & 5 years respectively.<br />

Results: Of 1343 patients who received an OLT, 6 developed<br />

de novo AIH. The etiology of initial liver failure was alcohol in<br />

3 of these patients; and steatohepatitis, cryptogenic hepatocellular<br />

carcinoma, primary sclerosing cholangitis in one patient<br />

each. We identified a total of 11 matched controls for these<br />

patients. Mean donor age was 45 years in the de novo group<br />

and 40 years in the control group (p=0.62). 2 patients in the<br />

de novo group and 4 in the control group developed biliary<br />

stricture post-operatively. One of the control patients developed<br />

post-op cholestasis. Mean Tacrolimus levels at week 1; week<br />

12, 6 months and 1 year were not statistically different. Mean<br />

Tacrolimus levels at week 4 was 10 in the de novo AIH group<br />

and 5.41 in the control group (p= 0.02). At 1 year, levels<br />

were 8.53 in the de novo group and 7.0 in the control group<br />

(p=0.61). The average time to development of de novo AIH<br />

was 1096 days (range 31-2822 days). Mean total prednisone<br />

used was 2697 mg in the de novo group and 1711 mg in the<br />

control group (p=0.60). Three de novo patients were noted<br />

to have fibrosis at the time of diagnosis (2 with severe fibrosis<br />

and 1 with mild fibrosis) and 1 control patient developed<br />

mild fibrosis (p=0.09). 1 de novo patient developed cirrhosis<br />

(788 days), none of the controls did. None of the patients<br />

received a re-transplantation. Survival at 1 year was 100% in<br />

the de novo group and 82% in the control group. Survival at 3<br />

years was 80% and 82% respectively. Survival at 5 years was<br />

40% and 82% respectively (p=0.34). Two deaths in the control<br />

group and one death in the de novo group were attributed to<br />

infection. Conclusion. De novo AIH is relatively rare post-OLT.<br />

It might be associated with a higher potential for significant<br />

fibrosis and increased mortality.<br />

Disclosures:<br />

Dilip Moonka - Advisory Committees or Review Panels: Gilead; Grant/Research<br />

Support: Bristol-Myers Squibb; Speaking and Teaching: AbbVie, Gilead<br />

Kimberly Ann Brown - Advisory Committees or Review Panels: CLDF, Merck,<br />

BMS, ABBVIE, Gilead, Gilead, Janssen, Salix; Consulting: Blue Cross Transplant<br />

Centers; Grant/Research Support: CLDF, Gilead, Exalenz, CDC, BMS, Hyperion,<br />

Merck; Speaking and Teaching: ABBVIE, Gilead, CLDF<br />

The following authors have nothing to disclose: Aishwarya Kuchipudi, Ryan<br />

Morton, Sarah Y. Chan, Syed M. Jafri


368A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

308<br />

Tacrolimus or Mycophenylate mofetil as a second-line<br />

therapeutic options in patients with Autoimmune Hepatitis:<br />

An international multicentre observational study.<br />

Cumali Efe 5 , Hannes Hagström 3 , Rahima A. Bhanji 2 , Niklas F.<br />

Müller 1 , Qixia Wang 4 , Tugrul Purnak 5 , Luigi Muratori 6 , Mårten<br />

Werner 7 , Hanns-Ulrich Marschall 8 , Paolo Muratori 6 , Fulya Gunsar<br />

9 , Daniel Klintman 10 , Albert Parés 11 , Alexandra Heurgué-Berlot<br />

12 , Thomas D. Schiano 13 , Xiong Ma 4 , Aldo J. Montano-Loza 2 ,<br />

Thomas Berg 1 , Ersan Ozaslan 14 , Eric M. Yoshida 15 , Staffan Wahlin<br />

3 ; 1 Universitätsklinikum Sektion Hepatologie, Klinik für Gastroenterologie<br />

und Rheumatologie., Leipzig,, Germany; 2 University of<br />

Alberta Division of Gastroenterology and Liver Unit., Alberta, AB,<br />

Canada; 3 Gastroenterology and Hepatology, Karolinska Institutet,<br />

Karolinska University Hospital., Stockholm, Sweden; 4 Department<br />

of Hepatology, Shanghai University of Traditional Chinese Medicine,<br />

Longhua Hospital, Shanghai,, China; 5 Department of Gastroenterology,<br />

Hacettepe University,, Ankara, Turkey; 6 Department of<br />

Clinical Medicine, Alma Mater Studiorum, University of Bologna,,<br />

Bologna, Italy; 7 Departments of Medicine, Sections for Hepatology<br />

and Gastroenterology, University Hospital of Skåne,, Malmö,<br />

Sweden; 8 Department of Molecular and Clinical Medicine, Sahlgrenska<br />

Academy, University of Gothenburg,, Gothenburg, Sweden;<br />

9 Department of Gastroenterology, Ege University, Bornova,,<br />

Izmir, Turkey; 10 Department of Molecular and Clinical Medicine,<br />

Lund University of Hospital,, Lund, Sweden; 11 Liver Unit, Hospital<br />

Clínic University of Barcelona C/ Villarroel 170 08036, Barcelona,<br />

Spain; 12 Department of Hepato-Gastroenterology, CHU<br />

Reims,, Reims, France; 13 Division of Liver Diseases, The Mount<br />

Sinai Medical Center, New York,, NY; 14 Department of Gastroenterology,<br />

Numune Research and Education Hospital,, Ankara, Turkey;<br />

15 Division of Gastroenterology University of British Columbia<br />

and Vancouver General Hospital., Vancouver, BC, Canada<br />

Background: Autoimmune hepatitis (AIH) is a well-recognized<br />

immune mediated liver disease. Eighty percent (80 %) of<br />

patients respond to standard therapy, including corticosteroids<br />

alone or combination with azathioprine (AZA). However, this<br />

therapy is not tolerated or insufficient to control disease activity<br />

in 20 % of patients. Second-line treatment with Tacrolimus (Tac)<br />

or Mycophenylate mofetil (MMF) has been described in only<br />

small case series. The aim of this study was to evaluate the outcome<br />

of Tac and/or MMF therapy in patients with AIH who are<br />

resistant or intolerant to standard immunosuppression. Patients<br />

and methods: This multi-centre retrospective chart review study<br />

evaluated data of AIH, including overlap with PBC or PSC<br />

patients, treated with Tac and/or MMF. Initial treatment was<br />

corticosteroids ±AZA. The patients were categorized as complete<br />

responders to initial therapy (Group-I), partial responders<br />

(Group-II) and lack of responders (Group-III). The IAIHG criteria<br />

were used for the diagnosis of AIH. Results: 163 patients (121<br />

female), mean age of 37 (9-76) years were included. Patients<br />

were treated with Tac (n=71) and MMF (n=92) a median of<br />

16 months (1-252 months) after standard immunosuppression.<br />

MMF and Tac resulted in complete or partial response in all<br />

of group-I patients. In group-II and group-III, Tac induced complete<br />

or partial response in 75% (36/45) vs 50 % (23/46) of<br />

MMF patients (p=0.003). Median follow up duration was 42.5<br />

(1-2430) months after MMF and/or Tac therapy. Five and 10<br />

years follow up rates were 37 % (34/92) and 19.6 % (18/92)<br />

for MMF patients and 51% (36/71) and 14% (10/71) for Tac<br />

patients. Combination of Tac and MMF was given to only 6<br />

patients, with 4 responding to therapy. Eleven patients were<br />

eventually transplanted. MMF and/or Tac were discontinued<br />

due to side effects in 7 patients including, 4 Tac and 3 MMF.<br />

Conclusions: MMF and Tac are generally effective and well tolerated<br />

in standard treatment resistant or intolerant AIH patients.<br />

There is no significant differences in outcomes between MMF<br />

and Tac treatments in AIH patients who responded to conventional<br />

therapy. However, Tac may be superior in AIH patients<br />

who showed incomplete or non-response to standard immunosupression.<br />

Disclosures:<br />

Hanns-Ulrich Marschall - Consulting: Albireo<br />

Alexandra Heurgué-Berlot - Consulting: Abbvie; Speaking and Teaching: Gilead<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

The following authors have nothing to disclose: Cumali Efe, Hannes Hagström,<br />

Rahima A. Bhanji, Niklas F. Müller, Qixia Wang, Tugrul Purnak, Luigi Muratori,<br />

Mårten Werner, Paolo Muratori, Fulya Gunsar, Daniel Klintman, Albert Parés,<br />

Thomas D. Schiano, Xiong Ma, Aldo J. Montano-Loza, Ersan Ozaslan, Eric M.<br />

Yoshida, Staffan Wahlin<br />

309<br />

Killer cell immunoglobulin-like receptors and their HLA<br />

class I ligands in autoimmune hepatitis: a possible<br />

pathogenetic role ?<br />

Simona Onali 2 , Roberto Littera 1 , Carlo Carcassi 1 , Luca Secci 2 ,<br />

Sara Lai 1 , Rita Porcella 1 , Sara Cappellini 2 , Claudia Salustro 2 ,<br />

Cinzia Balestrieri 3 , Giancarlo Serra 3 , Maria Conti 3 , Teresa<br />

Zolfino 4 , Francesco Figorilli 2 , Michele Casale 2 , Stefania Casu 2 ,<br />

Maria Cristina Pasetto 2 , Laura Matta 2 , Rosetta Scioscia 2 , Lucia<br />

Barca 2 , Luchino Chessa 2,3 ; 1 Medical Genetics, Department of<br />

Medical Sciences “M.Aresu”, University of Cagliari, Cagliari,<br />

Italy; 2 Center for the Study of Liver Diseases, Department of Medical<br />

Sciences “M.Aresu”, University of Cagliari, Cagliari, Italy;<br />

3 Department of Internal Medicine, Azienda Ospedaliero-Universitaria,<br />

Cagliari, Cagliari, Italy; 4 S.S.D. Coordinamento Trapianti di<br />

fegato, Azienda Ospedaliera Brotzu, Cagliari, Italy<br />

Background: Natural killer (NK) cells are widely distributed in<br />

hepatic tissue, but their role in the pathogenesis of autoimmune<br />

hepatitis (AIH) has not been fully estabilished yet. Killer cell<br />

immunoglobulin receptors (KIRs) represent the major family of<br />

cell surface receptors that inhibit and activate NK cells. Aim:<br />

To assess whether different type of KIRs could be involved in<br />

pathogenesis of AIH. Methods: We analyzed all consecutive<br />

Sardinian outpatients with type I AIH, referred to the Liver Unit<br />

of University of Cagliari (Italy) between January and October<br />

2014. AIH was diagnosed according to the International Autoimmune<br />

Hepatitis Group Revised Original Scoring System and<br />

AASLD guidelines. Patients with type II AIH, overlap syndrome<br />

and other chronic liver diseases were excluded. The immunogenetic<br />

characteristics of AIH patients were compared to those<br />

of 221 healthy individuals extracted from the Sardinian bone<br />

marrow donor registry. High resolution (4 digits) typing of HLA-<br />

A, B, C and DR loci and 15 KIRs gene loci was performed in<br />

both patients and controls. Subjects were divided into 2 groups<br />

according to homozygosity for KIR haplotype A (KIR genotype<br />

AA), heterozygosity or homozygosity for KIR haplotype B (KIR<br />

genotype Bx). The study was approved by local ethics committee<br />

and informed consent was obtained from all participants.<br />

Results. 114 subjects were included: median age 55 years<br />

(range 17-80), 87.7% females, median duration of disease 6<br />

years (range 0.6-23). The frequency of activating KIR gene,<br />

KIR2DS1, was significantly higher in AIH patients compared<br />

to controls (57% vs 43%, p=0.03, HR=1.7, 95% CI=1.0-2.7),<br />

as well as KIR2DL1+/KIR2DS1+/HLA-C2+ (47.3% vs 35.7%,<br />

p=0.04, HR=1.6, 95% CI=1.0-2.6). When we considered the<br />

age of onset of the autoimmune hepatitis, we found that activating<br />

KIR gene KIR2DS1 was more frequent in patients diag-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 369A<br />

nosed at younger age. Conclusions: Based on our findings, the<br />

activating KIR gene KIR2DS1 is highly expressed in patients<br />

with autoimmune hepatitis, suggesting its potential involvement<br />

in pathogenesis of type I AIH. However further <strong>studies</strong> are<br />

needed to fully understand the role of KIR2DS1 as marker for<br />

AIH.<br />

Disclosures:<br />

Teresa Zolfino - Board Membership: Abvie; Grant/Research Support: Bayer;<br />

Speaking and Teaching: Bristol<br />

Luchino Chessa - Board Membership: Abbvie; Speaking and Teaching: BMS,<br />

Jannsen<br />

The following authors have nothing to disclose: Simona Onali, Roberto Littera,<br />

Carlo Carcassi, Luca Secci, Sara Lai, Rita Porcella, Sara Cappellini, Claudia<br />

Salustro, Cinzia Balestrieri, Giancarlo Serra, Maria Conti, Francesco Figorilli,<br />

Michele Casale, Stefania Casu, Maria Cristina Pasetto, Laura Matta, Rosetta<br />

Scioscia, Lucia Barca<br />

310<br />

Possible involvement of activating follicular helper T<br />

cells in autoimmune hepatitis<br />

Naruhiro Kimura, Satoshi Yamagiwa, Hiroki Honda, Toru Setsu,<br />

Kentaro Tominaga, Hiroteru Kamimura, Masaaki Takamura,<br />

Shuji Terai; gastroenterology and hepatology, Niigata university,<br />

Niigata, Japan<br />

BACKGROUND: There is an increasing interest in the role of<br />

follicular helper T (Tfh) cells in autoimmunity from the perspective<br />

of both their role in breaking tolerance and their effects on<br />

disease progression. Dysregulated Tfh cells have been shown<br />

to be responsible for the induction of fatal autoimmune hepatitis<br />

(AIH) in mice that are unable to produce natural regulatory T<br />

cells after neonatal thymectomy and that are genetically devoid<br />

of the programmed cell death 1 (PD-1)-mediated signaling.<br />

However, the involvement of Tfh cells in the immune pathogenesis<br />

of AIH has not been fully characterized. AIM: To evaluate<br />

the numbers of different subsets of circulating Tfh and B cells<br />

and their potential role in the pathogenesis of AIH. METHODS:<br />

Heparinized peripheral blood was collected from 18 patients<br />

with AIH, 22 patients with chronic hepatitis B (CHB) and 21<br />

healthy volunteers (HC). Mononuclear cells were separated<br />

using the Ficoll gradient, and various surface markers were<br />

then investigated using flow cytometry. Cytokine production<br />

was investigated using peripheral blood Tfh cells after stimulation<br />

with phorbol myristate acetate (PMA) and ionomycin. We<br />

also investigated the distribution of CXCR5 + CD4 + cells in the<br />

liver using immunohistochemical staining. RESULTS: CXCR5 +<br />

CD4 + Tfh cells comprised 8.3% (median) (range 2.7-18.3),<br />

7.8% (4.1-13.5), and 8.1% (3.1-13.5) of the total T cells in the<br />

blood of patients with AIH, CHB, and HC, respectively. The Tfh<br />

cells were then classified into several subsets according to their<br />

expression of PD-1, inducible co-stimulator (ICOS) and CXCR3.<br />

We found a significant increase in the proportion of activated<br />

ICOS + Tfh cells in the blood (19.4%) and livers (24.3%) of<br />

patients with AIH compared with the proportion in the blood<br />

of the same patients after prednisolone (PSL) treatment (1.6%).<br />

The frequency of ICOS + Tfh cells was significantly decreased<br />

in the PSL-treated patients (median 22.6%, range 15.2-34.5)<br />

compared with the patients who were not treated with PSL<br />

(median 35.7%, range 26.0-56.2), although the levels of alanine<br />

aminotransferase (ALT) were decreased after treatments<br />

in both groups (17 ± 30.2 IU/l in PSL-treated and 15 ± 14.7<br />

IU/l in non-PSL-treated patients). CXCR5 + cells were detected<br />

among infiltrated lymphocytes in the portal area of the liver<br />

of patients with AIH. Furthermore, we confirmed that Tfh cells<br />

secreted IL-21 and IL-17 and expressed PD-1 after stimulation<br />

with PMA and ionomycin. CONCLUSIONS: Previous <strong>studies</strong><br />

have reported that an increase in Tfh cells is associated with<br />

the activity of autoimmune diseases. Our results suggest that<br />

ICOS + Tfh cells may be associated with the disease activity of<br />

AIH.<br />

Disclosures:<br />

Shuji Terai - Speaking and Teaching: Otsuka Pharma.<br />

The following authors have nothing to disclose: Naruhiro Kimura, Satoshi<br />

Yamagiwa, Hiroki Honda, Toru Setsu, Kentaro Tominaga, Hiroteru Kamimura,<br />

Masaaki Takamura<br />

311<br />

Methotrexate therapy for Autoimmune Hepatitis<br />

James Haridy 1 , Amanda J. Nicoll 1,2 , Siddharth Sood 1 ; 1 Department<br />

of Gastroenterology and Hepatology, Royal Melbourne<br />

Hospital, Melbourne, VIC, Australia; 2 Department of Gastroenterology,<br />

Eastern Health, Melbourne, VIC, Australia<br />

Background: Autoimmune hepatitis (AIH) is characterised by<br />

typically diverse phenotypical manifestation and response to<br />

treatment. Corticosteroids and azathioprine are standard firstline<br />

therapies. There are limited options in patients who display<br />

poor response or intolerance to these medications. Methotrexate<br />

(MTX) is readily available, often employed in other autoimmune<br />

conditions, and has therefore been postulated as a<br />

possible second-line therapy. Evidence of efficacy has been<br />

limited to a small number of case-reports, despite more widespread<br />

anecdotal use. Aim: To assess the response to MTX in<br />

AIH. Method: A retrospective case-series was conducted using<br />

subjects located through the gastroenterology liver database of<br />

a tertiary referral centre. Subjects with a diagnosis of AIH were<br />

identified and diagnosis confirmed using the International Autoimmune<br />

Hepatitis Group criteria. Medical records were examined<br />

to determine medication history. Patients were included<br />

if they had prior treatment exposure to methotrexate (MTX) for<br />

their AIH. Baseline and follow-up data was collected to determine<br />

response to MTX over the initial 12-months of therapy. Follow-up<br />

was collected up to 3-years following initiation of MTX<br />

if available. Results: 11 patients (age 24 – 72 years; mean,<br />

56 years) were identified with confirmed AIH and MTX use.<br />

10/11 (90.9%) were female and 4/11 (36.4%) had biopsy<br />

or imaging evidence of cirrhosis prior to initiation of MTX. The<br />

most common indication for MTX was incomplete response<br />

to thiopurines (5/11, 45.5%). The median dose of prednisolone<br />

prior to initiating MTX was 8.0mg (range 0-15mg). 5/11<br />

(45.5%) of subjects ceased MTX within twelve months due to<br />

an adverse event, most commonly deterioration in liver function<br />

(range 1-12 months). 5/11 (45.5%) achieved or maintained<br />

complete remission at 12-months on MTX. These patients experienced<br />

a median decrease in prednisolone dosage of 5.3 mg,<br />

with a median dose of 3 mg (range 0 – 5mg) at 12-months.<br />

1/11 (9%) achieved a partial response at 12-months but<br />

later achieved complete remission on MTX monotherapy by<br />

36-months. 1/6 (16.7%) patients continuing MTX at 12-months<br />

were steroid free. Conclusion: Prior use of MTX in AIH has been<br />

limited to three individual case reports, however it is mentioned<br />

in guidelines as a possible alternative second-line therapy. In<br />

our case series, we show that MTX is safe, and that over 50%<br />

of patients are able to be maintained on it with good effect on<br />

liver function and minimal steroid use. MTX may have a role in<br />

treatment of AIH in a subset of patients who are refractory or<br />

intolerant of first-line therapies.<br />

Disclosures:<br />

Amanda J. Nicoll - Grant/Research Support: MSD; Speaking and Teaching:<br />

Bayer<br />

Siddharth Sood - Grant/Research Support: Qiagen USA<br />

The following authors have nothing to disclose: James Haridy


370A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

312<br />

Features and patterns of concurrent extrahepatic autoimmune<br />

diseases in patients with Autoimmune Hepatitis<br />

and Overlap/Variant Syndromes: Implications for diagnosis<br />

and management<br />

Guan Wee Wong, Andrew D. Yeoman, Michael A. Heneghan;<br />

Institute of Liver Studies, King’s College Hospital, London, London,<br />

United Kingdom<br />

Aim: The presence of concurrent extrahepatic autoimmune disease<br />

(CEAID) represents an important diagnostic criterion in<br />

the International Autoimmune Hepatitis Group (IAIHG) scoring<br />

system. We aim to describe the clinical patterns of CEAID in<br />

patients with AIH and Overlap Syndromes (OS). Methods: The<br />

medical records and prospectively accrued electronic database<br />

of 300 patients with definite AIH as defined by the revised<br />

IAIHG criteria were reviewed. We also interrogated the medical<br />

records of 28 patients defined as OS, 10 AIH/PBC and<br />

18 AIH/PSC. Patients with Autoimmune Sclerosing Cholangitis<br />

(AISC) were excluded. Results: A total of 34 CEAID diagnosis<br />

were identified in our 328 patients. Among them, 81 patients<br />

(24.7%) had one, 19 (5.8%) had two, 2 (0.6%) had three<br />

diagnosed CEAIDs. In total, 88/300 (29%) of AIH patients<br />

had CEAID. Autoimmune thyroid disease (AITD) was identified<br />

in 36 patients (12%) with hypothyroidism being more common<br />

(21/300, 7%) followed by symptomatic goitre (8/300, 2.7%)<br />

and hyperthyroidism (5/300, 1.6%). Prevalence of a-priori diabetes<br />

(non-steroid induced) was found in 13 patients (4.3%),<br />

with 12 patients (92%) being Type 1. Inflammatory bowel disease<br />

(IBD) was present in 8 patients (2.7%) with ulcerative colitis<br />

(UC) representing the underlying IBD diagnosis. Connective<br />

tissue disease (CTD) was identified in 28 patients (9%), with<br />

rheumatoid arthritis reported in 15 patients (5%) and other<br />

rarer CTDs including non-specific arthritis (5 patients, 1.6%),<br />

SLE/ Lupus and Sjogren’s syndrome (3 patients each, 1%), Raynaud’s<br />

and antiphospholipid syndrome (1 patient each, 0.3%).<br />

Autoimmune skin lesions were identified in 13 patients (4.3%)<br />

ranging from vitiligo (4 patients, 1.3%), alopecia (3 patients,<br />

1%), leucocytoclastic vasculitis (2 patients, 0.6%), to erythema<br />

nodosum, lichen planus and erythema multiforme (1 patient<br />

each, 0.3%). Miscellaneous conditions (n=13) were reported<br />

including idiopathic thrombocytopenia purpura, multiple sclerosis<br />

and myasthenia gravis. For OS, the pattern of AIH/PBC<br />

associated CEAID differed from pure AIH, with Sjogren’s syndrome<br />

being more representative (30%). Similarly, UC was<br />

more commonly found in patients with AIH/PSC overlap (33%).<br />

Moreover, positive family history of CEAID was identified in<br />

52/300 (17.3%) patients with AIH. Of these, 23/52 (44 %)<br />

had CEAID in their personal history. Conclusions: CEAIDs are<br />

common in AIH and OS. ATID, Sjogren’s syndrome and UC<br />

are the most common CEAIDs identified in our patients with<br />

AIH, AIH/PBC and AIH/PSC respectively. Early recognition of<br />

these features may be the first clue leading to subsequent early<br />

diagnosis and effective management of AIH and OS.<br />

Disclosures:<br />

The following authors have nothing to disclose: Guan Wee Wong, Andrew D.<br />

Yeoman, Michael A. Heneghan<br />

313<br />

The Need for Improved Liver Literacy in the US Population<br />

Tracy J. Mayne, Herbert Swanson; Intercept Pharmaceuticals, San<br />

Diego, CA<br />

Objective: To assess the level of awareness, knowledge, attitudes<br />

and behaviors surrounding liver health and liver disease<br />

in the general US population. Methods: We conducted online<br />

surveys from 1/6/15 – 1/12/15 in the GfK Knowledge Panel,<br />

a stratified probability sample representative of the population<br />

of US households. To correct for technology & income bias,<br />

households without a computer/internet were provided hardware/internet<br />

access to participate. We selected 511 respondents<br />

using random probability address-based sampling. Final<br />

data were weighted by age, region, race/ethnicity, education,<br />

and income according to the 3/14 US Census Current<br />

Population Survey. Margin of error was ±3.4%. Results: Most<br />

respondents do not perceive themselves to be at risk for liver<br />

disease. They do not think about or discuss it with friends, family<br />

or their physician. Almost half had some belief that a person<br />

can live without a liver. Most participants were unaware that<br />

liver function testing is part of routine bloodwork, and few<br />

discussed liver test results with their physician. These findings<br />

do not reflect general lack of health awareness or knowledge:<br />

91% of participants were aware of tests and values for blood<br />

pressure (91%), blood sugar (81%), cholesterol (79%), and<br />

BMI (69%), and respondents were much more likely to discuss<br />

these with their physician. Respondents reported greater likelihood<br />

of thinking and worrying about other diseases (weight,<br />

heart, breast, mental, prostate, colon and kidney) than liver.<br />

Patients reported more stigma associated with liver cirrhosis<br />

than with kidney disease, heart disease, cancer (colon, breast,<br />

prostate, or lung), diabetes, or reproductive health problems.<br />

Conclusions: In a representative US sample, awareness and<br />

concern about liver disease ranked low compared to other<br />

diseases. Participants report low levels of interaction with physicians<br />

about liver disease compared to other diseases, and<br />

were much less familiar with measures of liver health than (for<br />

example) cardiovascular. Massive public health campaigns in<br />

cardiovascular health have raised awareness and increased<br />

monitoring, screening & treatment. With increasing non-viral<br />

liver diseases, such as non-alcoholic fatty liver and steatohepatitis,<br />

these data point to a need for broad public health educational<br />

campaigns in liver disease.<br />

Disclosures:<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

Herbert Swanson - Management Position: Intercept<br />

314<br />

Longitudinal alteration in health-related quality of<br />

life and its impact on clinical course in patients with<br />

advanced hepatocellular carcinoma who received<br />

sorafenib treatment<br />

Masako Shomura 1 , Tatehiro Kagawa 2 , Koichi Shiraishi 2 , Shunji<br />

Hirose 2 , Yoshitaka Arase 2 , Kota Tsuruya 2 , Sachiko Takahira 3 ,<br />

Haruka Okabe 1 , Tetsuya Mine 2 ; 1 Department of Nursing, Tokai<br />

University School of Health Sciences, Isehara city, Japan; 2 Department<br />

of Gastoroenterology, Tokai University School of Medicine,<br />

Isehara, Japan; 3 Department of Nursing, University of Nagasaki,<br />

Nagasaki, Japan<br />

Purpose: This study aims to clarify longitudinal alteration in<br />

health related quality of life (HRQOL) in patients with advanced<br />

hepatocellular carcinoma (HCC) receiving sorafenib until<br />

death or treatment discontinuation, and to seek the HRQOL


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 371A<br />

domains associated with prognosis. Methods: We assessed<br />

HRQOL using SF-36 v2 TM every three months and calculated<br />

eight norm-based domain scores defined 50 points as national<br />

standards in consecutive patients with advanced HCC who<br />

received sorafenib. We prospectively analyzed the relationship<br />

among HRQOL, baseline characteristics, overall survival<br />

(OS) and treatment duration by Cox’s proportional hazards<br />

model. All subjects were given educational instructions and<br />

telephone follow-up. Results: Fifty-four patients participated in<br />

the study, comprising 42 males (78%) with median age of 71<br />

years, 23 (43%) with hepatitis C virus infection, 33 (61%) with<br />

Child-Pugh score 5, and 30 (56%) with TNM stage IV. Median<br />

OS and treatment duration were 9 and 5 months, respectively.<br />

Forty patients (74%) died. Thirteen patients receiving sorafenib<br />

over one year maintained all domain scores above 40 without<br />

significant decline throughout the treatment period. In contrast,<br />

physical functioning (PF), role physical, and vitality scores<br />

declined continuously and significantly, one year before death<br />

in 40 deceased patients. Child-Pugh score 5 and PF score<br />

^40 at baseline were significantly associated with longer OS<br />

by multivariate analysis. Social functioning (SF) score ^40<br />

was the only significant predictor for longer treatment duration<br />

(Figure). Conclusions: HRQOL was not impaired significantly<br />

in patients who received sorafenib treatment over one year.<br />

PF score ^40 and SF score ^40 at baseline was significantly<br />

associated with longer OS and longer treatment duration,<br />

respectively. Thus, HRQOL could be a valuable marker to predict<br />

the clinical course of HCC patients receiving sorafenib.<br />

Disclosures:<br />

The following authors have nothing to disclose: Masako Shomura, Tatehiro<br />

Kagawa, Koichi Shiraishi, Shunji Hirose, Yoshitaka Arase, Kota Tsuruya, Sachiko<br />

Takahira, Haruka Okabe, Tetsuya Mine<br />

315<br />

Interferon (IFN), Ribavirin (RBV), Duration of Treatment<br />

and Baseline Patient-Reported Outcomes (PROs) Predict<br />

Adherence to the New Regimens for Treatment of<br />

Chronic Hepatitis C (CH-C)<br />

Zobair M. Younossi 1,2 , Maria Stepanova 3 , Linda Henry 3 , Fatema<br />

Nader 1 , Sharon L. Hunt 3 ; 1 Center For Liver Disease, Department<br />

of Medicine, Inova Fairfax Hospital, Falls Church, VA; 2 Betty and<br />

Guy Beatty Center for Integrated Research, Inova Health System,<br />

Falls Church, VA; 3 Center for Outcomes Research in Liver Disease,<br />

Washington, DC<br />

BACKGROUND AND AIM: Patients’ experience (PROs) and<br />

treatment regimens may affect treatment adherence. Our aim<br />

was to assess the impact of clinical and PRO factors on adherence<br />

to different anti-HCV regimens. METHODS: PRO data (SF-<br />

36, CLDQ-HCV, FACIT-F, WPAI:SHP) from 13 multinational<br />

clinical trials of anti-HCV treatment were available. Treatment<br />

non-adherence was defined as


372A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

316<br />

TNF-alpha is the Most Consistent Predictor of Impairment<br />

of Mental Health-Related Patient-Reported Outcomes<br />

(PROs) in Patients with Chronic Hepatitis C (CH-C)<br />

Treated with Ledipasvir (LDV) and Sofosbuvir (SOF) with<br />

or without Ribavirin (RBV)<br />

Zobair M. Younossi 1,2 , Maria Stepanova 2 , James M. Estep 2 , Azza<br />

Karrar 2 , Bibiana Oe 2 , Elzafir Elsheikh 2 , Siddharth Hariharan 2 ,<br />

Kazi Ahmed 2 , Patricia Tran 2 , Fatema Nader 2 , Linda Henry 2 , Ali<br />

A. Weinstein 2 , Naomi Lynn Gerber 1,2 ; 1 Center For Liver Disease,<br />

Department of Medicine, Inova Fairfax Hospital, Falls Church,<br />

VA; 2 Betty and Guy Beatty Center for Integrated Research, Inova<br />

Health System, Falls Church, VA<br />

Background and aim: HCV infection has been associated with<br />

changes in brain metabolites that may result directly from the<br />

virus or mediated by exposure to proinflammatory cytokines<br />

in the central nervous system. Our aim was to evaluate the<br />

association of cytokines and mental health PROs in HCV+<br />

patients treated with LDV/SOF±RBV. Methods: We included<br />

100 CH-C GT1 patients (age 53.1±9.6, 57% male, 86%<br />

white, BMI of 28.5±5.6) treated with LDV/SOF (N=50) or<br />

LDV/SOF+RBV (N=50). All patients completed several validated<br />

PRO instruments. Mental health PROs were measured<br />

by the Role Emotional (RE), Mental Health (MH) and Mental<br />

Component Summary (MCS) of SF-36, Emotional Well-being<br />

(EWB) of FACIT-F, and Emotional domain (EMM) of CLDQ-<br />

HCV. Frozen serum was analyzed for interleukins (IL) 1, 1ra,<br />

8, 10, tumor necrosis factor-α (TNF-α), interferon-γ, monocyte<br />

chemotactic protein 1/chemokine ligand 2, granulocyte colony<br />

stimulating factor, vaso-endothelial and platelet-derived<br />

growth factors using Bio-Plex, ELISA, and enzymatic assays<br />

baseline, treatment week 12 and 4 weeks post-treatment. Relationships<br />

between fatigue and clinical and laboratory data<br />

were assessed by multiple regression analyses. Results: Patients<br />

treated with LDV/SOF+RBV had more significant anemia (10%<br />

vs 0%; p=0.02), more psychiatric (26% vs 10%; p=0.04) and<br />

skin (22% vs 6%, p=0.02) adverse events during treatment.<br />

After adjustment for pre-treatment history of anxiety, the most<br />

consistent independent biomarker predictor of lower mental<br />

health PROs during treatment was the level of serum TNF-α<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 373A<br />

318<br />

Uncertainty and Hepatitis C: A Correlational Study<br />

Examining Mishel’s Uncertainty in Illness Theory<br />

Humberto Reinoso; Nursing, Mercer University, Atlanta, GA<br />

Background: Hepatitis C is the most common blood-borne infection<br />

worldwide. It is estimated that some 5.3 million Americans<br />

are living with hepatitis C. Some <strong>studies</strong> predict up to<br />

7.1 million individuals affected including high-risk populations<br />

(Chak, Talal, Sherman, Schiff, & Saab, 2011). Individuals born<br />

between 1945 and 1965, those from the baby boomers generation,<br />

account for approximately three-fourths of all hepatitis C<br />

cases in the U.S. (Centers for Disease Control and Prevention<br />

[CDC], 2012). Purpose: The purpose of this study was to examine<br />

Mishel’s Uncertainty in Illness Theory among individuals<br />

with hepatitis C. The study focused on the relationship a set<br />

of variables prescribed by the model had on individual’s perception<br />

of uncertainty. Methods: A cross-sectional design was<br />

used to study the relationship between uncertainty and a set<br />

of variables prescribed by Michel’s Uncertainty in Illness Theory<br />

(i.e. educational attainment, healthcare authority figures,<br />

social network, years since diagnosis, treatment attempts, and<br />

treatment experience). Results: Data were collected over an<br />

8-week period from a convenience sample of participants born<br />

between 1945 and 1965, who self-identified as having hepatitis<br />

C (n = 146). The relationship between perceived uncertainty<br />

and a set of variables prescribed by the theory was investigated<br />

using Pearson product-moment correlation coefficients.<br />

Preliminary analyses were preformed to ensure no violation of<br />

the assumptions of normality, linearity, and homoscedasticity.<br />

Educational attainment (r = .2, p = < .05) as well as treatment<br />

experience (r = .27, p = < .05) shared a positive relationship<br />

with the individual’s perception of uncertainty. Social network<br />

(r = -.51, p = < .01) shared a strong negative or inverse relationship<br />

with participant’s perception of uncertainty. Furthermore,<br />

there was a strong positive relationship between the<br />

perception of uncertainty experienced by those individuals with<br />

hepatitis C and information provided by healthcare authority<br />

figures (r = .73, p = < .01). Conclusion: The results of this study<br />

suggest that the perception of uncertainty experienced by those<br />

individuals with hepatitis C may be influenced positively or<br />

negatively by information provided by their healthcare authority<br />

figures as well as their social network. Further research is<br />

recommended, in the form of a multiple regression analysis, of<br />

those variables identified of having strong relationships with an<br />

individual’s perception of uncertainty in order to identify their<br />

predictive ability.<br />

Disclosures:<br />

The following authors have nothing to disclose: Humberto Reinoso<br />

319<br />

Small Anti-Warburg Molecules Kill Hepatocarcinoma<br />

Cells by Opening Voltage Dependent Anion Channels<br />

and Promoting Mitochondrial Oxidative Stress<br />

David N. DeHart 1 , Monika Gooz 1 , John J. Lemasters 1,2 , Eduardo<br />

N. Maldonado 1 ; 1 Drug Discovery & Biomedical Sciences, Medical<br />

University of South Carolina, Charleston, SC; 2 Biochemistry &<br />

Molecular Biology, Medical University of South Carolina, Charleston,<br />

SC<br />

BACKGROUND: Enhanced aerobic glycolysis and suppressed<br />

mitochondrial metabolism characterize the Warburg metabolism<br />

of tumors. Flux of metabolites across mitochondrial outer<br />

membranes occurs through voltage dependent anion channels<br />

(VDAC). Dimeric α,β-tubulin closes VDAC in planar lipid<br />

bilayers (PNAS 105:18746) and limits ingress of respiratory<br />

substrates and ATP through VDAC that support mitochondrial<br />

membrane potential (ΔΨ) formation in cancer cells (Cancer<br />

Res. 70:10192). The small molecule erastin opens VDAC by<br />

antagonizing the inhibitory effect of tubulin on VDAC (JBC<br />

288:11920). Here, we hypothesized that small molecules that<br />

antagonize VDAC-tubulin interaction increase mitochondrial<br />

formation of reactive oxygen species (ROS), decrease glycolysis,<br />

and activate c-jun N-terminal kinase (JNK), leading to<br />

mitochondrial dysfunction, cell death and tumor growth inhibition<br />

in hepatocarcinoma (HCC). Our AIM was to evaluate<br />

the effects of VDAC-tubulin antagonists on mitochondrial ΔΨ,<br />

ROS and lactate generation, JNK activation and cell killing<br />

in HepG2 and Huh7 HCC cells and in a xenograft model<br />

of Huh7 cells. METHODS: Confocal/multiphoton fluorescence<br />

microscopy assessed ΔΨ (tetramethylrhodamine methylester)<br />

and ROS (chloromethyldichlorofluorescein [cmDCF], MitoSOX<br />

Red). JNK was assessed by Western blotting and cell killing<br />

by propidium iodide fluorometry. Subcutaneous xenografts of<br />

Huh7 cells were developed in nude mice. RESULTS: Erastin<br />

and small molecules X1 and X2 identified in a high-throughput<br />

screen increased ΔΨ and prevented mitochondrial depolarization<br />

caused by cytosolic high free tubulin. Increased ΔΨ<br />

was followed by mitochondrial depolarization occurring 1-2<br />

h after X1 and X2 and 3-4 h after erastin. Lactate generation<br />

after X1 decreased by 60%. ROS increased 30 min after X1<br />

and 60 min after X2 and erastin, as assessed with cmDCF<br />

and MitoSOX. The mitochondrially targeted antioxidant MitoQ<br />

blocked this ROS increase. Additionally, erastin caused JNK<br />

phosphorylation within 1 h. X1 and X2-dependent cell killing<br />

of HCC cells was blocked by the antioxidant N-acetylcysteine.<br />

By contrast, X1 and X2 caused


374A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of CHOP and downstream target ATG5, an E3-like enzyme<br />

required for lipidation of LC3b, through short hairpin RNA<br />

repressed induction of autophagy and enhanced cell viability<br />

following saturated FFA exposure. Conversely, addition of<br />

chloroquine or bafilomycin A1, inhibitors of lysosome activity,<br />

further enhanced LC3b and P62 accumulation in combination<br />

with saturated FFA-induced cell death. Analysis of<br />

ultrastructural features and immunocytochemistry revealed that<br />

saturated FFA exposure perturbed proper fusion of autophagosomes<br />

with lysosomes. Together, these results suggest that<br />

exposure of hepatocytes to saturated FFA induce the UPR, and<br />

this stress response pathway enhances initiation of autophagy.<br />

Furthermore, saturated FFA block the flux of autophagy, which<br />

contributes to hepatoxicity. Co-treatment of saturated FFA with<br />

caspase inhibitor ZVAD-FMK, along with imaged localization<br />

of cleaved caspase 3, indicated that apoptosis was not playing<br />

a major role in hepatocyte cell death in this model system.<br />

However, treatment with saturated FFA or other inhibitors of<br />

autophagic flux led to an increase in proteasome activity in a<br />

CHOP dependent manner. It has been suggested that regulation<br />

of autophagy and the ubiquitin-proteasome pathway are<br />

closely linked, and often aberrant in diseases. This supports<br />

the idea that saturated FFA induces proteasome activity in a<br />

CHOP-dependent manner following inhibition of autophagic<br />

flux. To understand the relationship between the UPR and inhibition<br />

of autophagic flux in NASH patients, we assayed liver<br />

biopsy samples and found up-regulation of CHOP coincident<br />

with accumulation of LC3b and P62 in NASH patients. While<br />

previous work has suggested an association between the UPR<br />

and autophagy, our study provides a mechanistic understanding<br />

of the roles the UPR and CHOP play in regulating saturated<br />

FFA induced hepatotoxicity at the cellular level.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jeffrey A. Willy, Howard C.<br />

Masuoka, Ronald C. Wek<br />

321<br />

Hepatocyte surface localization of rat oatp1a4 requires<br />

cotrafficking in vesicles that also contain oatp1a1 and<br />

PDZK1: a new paradigm for subcellular trafficking<br />

of oatps that do not have a PDZ consensus binding<br />

domain<br />

Wen-Jun Wang, John W. Murray, Allan W. Wolkoff; Liver<br />

Research Center, Albert Einstein College of Medicine, Bronx, NY<br />

Organic anion transport proteins (oatp) 1a1 and 1a4 mediate<br />

uptake of endogenous and xenobiotic compounds by hepatocytes.<br />

In previous <strong>studies</strong> we showed that oatp1a1 and other<br />

oatps have a PDZ consensus binding sequence at their C-termini.<br />

Binding of oatp1a1 to PDZK1 is required for its microtubule-based<br />

vesicular trafficking to the plasma membrane via<br />

selective recruitment of specific motor proteins. Rat oatp1a4<br />

also traffics to the plasma membrane although it, like several<br />

other oatps, lacks a PDZ consensus binding sequence. In the<br />

present study, we tested the hypothesis that oatp1a4 traffics<br />

together with oatp1a1 in common intracellular vesicles, utilizing<br />

PDZK1 association with oatp1a1 for motor recruitment and<br />

plasma membrane targeting. Methods: Oatp1a1 and oatp1a4<br />

were immunolocalized in endocytic vesicles prepared from rat<br />

as well as wild type and PDZK1 knockout mouse livers. Immunofluorescence<br />

co-localization of proteins in vesicles was quantified.<br />

Plasma membrane distribution of oatp1a1 and oatp1a4<br />

was determined in HEK 293 cells (do not express endogenous<br />

PDZK1) transiently transfected with oatp1a1, oatp1a4,<br />

or oatp1a1 plus oatp1a4. Each of these transfections was performed<br />

with and without cotransfection of PDZK1. Results: 74%<br />

of vesicles prepared from rat liver that were associated with<br />

oatp1a4 (n=12,044) were also associated with oatp1a1. 51%<br />

of oatp1a1 associated vesicles (n=3608) were also associated<br />

with PDZK1. Similarly, 80% and 64%, respectively, of Oatp1a4-associated<br />

vesicles that were prepared from wild-type<br />

and PDZK1 knockout mouse liver were also associated with<br />

Oatp1a1 (n=5059 and 3415, respectively). Single transfection<br />

of HEK 293 cells with rat oatp1a1 or oatp1a4 revealed<br />

predominant cytoplasmic localization with little on the plasma<br />

membrane. Cotransfection of these cells with PDZK1 revealed<br />

that oatp1a1 localized to the cell surface while oatp1a4<br />

remained cytoplasmic. Double transfection with both transporters<br />

in the absence of PDZK1 revealed that they localized to the<br />

cytoplasm. In contrast, both oatp1a1 and oatp1a4 localized to<br />

the cell surface following triple transfection with PDZK1. Conclusions:<br />

(1) Interaction of rat oatp1a1 with PDZK1 is necessary<br />

for its trafficking to the cell surface. (2) Although rat oatp1a4<br />

does not interact with PDZK1, it can co-traffic (“hitchhike”) to<br />

the cell surface with oatp1a1 by virtue of their colocalization<br />

in intracellular vesicles containing PDZK1. (3) This provides a<br />

new paradigm to explain how oatps that lack a PDZ consensus<br />

binding domain can traffic to the cell surface and suggests<br />

heterooligomerization of oatps, a mechanism that is currently<br />

under study.<br />

Disclosures:<br />

Allan W. Wolkoff - Consulting: Synageva; Grant/Research Support: Merck<br />

The following authors have nothing to disclose: Wen-Jun Wang, John W. Murray<br />

322<br />

Tissue-Specific Subcellular Localization of Thioesterase<br />

Superfamily Member 1 (Them1) in the Pathogenesis of<br />

Nonalcoholic Fatty Liver Disease (NAFLD)<br />

Yue Li 1 , Susan J. Hagen 2 , David E. Cohen 1 ; 1 Medicine, Brigham<br />

and Women’s Hospital, Boston, MA; 2 Surgery, Beth Israel Deaconess<br />

Medical Center, Boston, MA<br />

Background: Altered regulation of lipid and glucose homeostasis,<br />

most often in the setting of insulin resistance and obesity,<br />

is central to the pathogenesis of NAFLD. Brown adipose tissue<br />

(BAT) is rich in mitochondria and mediates non-shivering<br />

thermogenesis in mice. BAT also plays important roles in promoting<br />

energy expenditure and its suboptimal function may<br />

contribute to the pathogenesis of obesity in humans. Them1<br />

is a long chain fatty acyl-CoA thioesterase that is enriched<br />

in BAT and is upregulated in mice in response to cold stress.<br />

Them1 -/- mice exhibit marked increases in energy expenditure,<br />

as well as resistance to diet-induced obesity and NAFLD.<br />

Under conditions of sustained overnutrition, Them1 appears<br />

to play a central role in the development of NAFLD both by<br />

limiting energy expenditure in BAT and by directly promoting<br />

steatosis in the liver, where it is also upregulated in response<br />

to high fat feeding. Notwithstanding these striking phenotypes<br />

and recent reports describing the enzymatic properties of the<br />

protein, very little is known about the cell biology of Them1.<br />

Aims: The objective of this study was to elucidate the subcellular<br />

localization of Them1 and its determinants in BAT and<br />

liver. Methods: Them1 -/- and wild type mice were subjected<br />

to cold stress by housing at 4 °C for 24 h or to overnutrition<br />

by high fat feeding for 16 w. Them1 in liver and BAT was<br />

visualized by immunofluorescence and confocal microscopy.<br />

Putative kinases of Them1 phosphorylation sites in BAT identified<br />

by phosphoproteomics (Cell 2010;143:1174) were predicted<br />

by sequence analysis (Nat Methods 2014;11:603).<br />

Them1-EGFP and a deletion mutant were studied using stably<br />

transfected 3T3-L1 cells, which had been differentiated into<br />

adipocytes using standard techniques. Results: In BAT of mice


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 375A<br />

subjected to cold stress, Them1 was strongly localized to lipid<br />

droplets. By contrast, Them1 in hepatocytes was visualized in<br />

cytosol, and this was not influenced by high fat feeding. PKCβ<br />

was the predicted kinase for phosphorylated sites (S15, S18,<br />

S25) near the N-terminus of Them1. Activation of PKCβ using<br />

phorbol 12-myristate 13-acetate led to the rapid redistribution<br />

of Them1-EGFP in a pattern consistent with lipid droplet association.<br />

The subcellular localization of Them1-EGFP was profoundly<br />

disrupted by the deletion of a short peptide fragment<br />

containing the phosphorylation sites. Conclusions: Our data<br />

demonstrate that cellular localization of Them1 is tissue specific<br />

and is regulated at least in part by phosphorylation. The tissue<br />

specific control of Them1 localization may play a key role in<br />

the pathogenesis of NAFLD.<br />

Disclosures:<br />

David E. Cohen - Advisory Committees or Review Panels: Merck, Aegerion,<br />

Genzyme; Consulting: Intercept<br />

The following authors have nothing to disclose: Yue Li, Susan J. Hagen<br />

323<br />

Longitudinal monitoring of hepatic endoplasmic reticulum<br />

calcium changes in rats<br />

Emily Wires, Mark Henderson, Kathleen Trychta, Brandon K. Harvey;<br />

National Institute on Drug Abuse, Baltimore, MD<br />

Obesity is a major health concern in developed countries; an<br />

estimated 1.5 billion individuals are considered overweight<br />

or obese worldwide. In stressed conditions, such as chronic<br />

obesity, hepatocytes accumulate excess long chain fatty acids,<br />

leading to the development of nonalcoholic fatty liver disease<br />

(NAFLD), representing a spectrum of liver disorders. Moreover,<br />

obese patients typically exhibit metabolic disorders, cardiovascular<br />

disease and obesity-related mortality. While the exact<br />

mechanisms of obesity-associated liver diseases are poorly<br />

understood, disruptions to endoplasmic reticulum (ER) homeostasis<br />

have been implicated in disease pathogenesis. The ER<br />

contains the highest level of intracellular calcium, an estimated<br />

1,000-10,000 fold greater than cytosolic levels. Perturbations<br />

to ER calcium levels are associated with a variety of pathologies,<br />

and have been predicted as a contributing factor to<br />

the development and progression of obesity-associated disorders.<br />

Recently, our lab developed a novel secreted ER calcium<br />

monitoring protein (SERCaMP), to longitudinally monitor ER<br />

calcium levels in vivo by measuring small volumes of blood;<br />

a notable enhancement from calcium-specific fluorescent dyes<br />

that become easily saturated in high levels of calcium and<br />

are unable to be utilized for in vivo <strong>studies</strong>. In the present<br />

study, rats expressing SERCaMP appended to Gaussia luciferase<br />

(GLuc-SERCaMP) in the liver were fed highly palatable<br />

foods, colloquially referred to as a cafeteria diet (e.g. cookies,<br />

pepperoni, potato chips), to mimic dietary intake observed<br />

in obese individuals. Additional groups included ad libitum<br />

access to chow and restricted access. Rats exposed to cafeteria<br />

diet items and ad libitum chow showed significant weight gain<br />

when compared to restricted rats. Blood samples indicated<br />

increased levels of circulating GLuc-SERCaMP in both cafeteria<br />

diet and ad libitum chow, suggesting that a dietary intake, particularly<br />

over eating, alters ER calcium homeostasis in the liver.<br />

The half-life of GLuc-SERCaMP is estimated to be less than 10<br />

minutes; eliminating the possible confound of sensor accumulation<br />

in the blood. Our study is the first to use a secreted ER calcium<br />

probe to demonstrate that dietary intake alters ER calcium<br />

homeostasis in the liver. Furthermore, parallel work has been<br />

done to monitor hepatic ER stress in order to establish a temporal<br />

profile between ER calcium dysregualtion and ER stress.<br />

Disclosures:<br />

The following authors have nothing to disclose: Emily Wires, Mark Henderson,<br />

Kathleen Trychta, Brandon K. Harvey<br />

324<br />

Mouse CD11b + Kupffer cells recruited from bone marrow<br />

accelerate liver regeneration after partial hepatectomy<br />

Hiroyuki Nakashima 1 , Kiyoshi Nishiyama 2 , Masahiro Nakashima 1 ,<br />

Manabu Kinoshita 1 , Shuhji Seki 1 ; 1 Immunology and Microbiology,<br />

National Defense Medical College, Saitama, Japan; 2 Surgery,<br />

National Defense Medical College, Saitama, Japan<br />

Background and Aims: Innate immune cells are known to<br />

contribute to the liver regeneration after partial hepatectomy<br />

(PHx). Although TNF and Fas/FasL are the vital components<br />

for hepatocyte regeneration, the role of Kupffer cells is still in<br />

controversy. Liver F4/80 + Kupffer cells are classified into two<br />

subsets; resident radio-resistant CD68 + cells with phagocytic<br />

and bactericidal activity, and recruited radio-sensitive CD11b +-<br />

cells with cytokine-producing capacity. The aim of this study is<br />

to investigate the role of these Kupffer cells in the liver regeneration<br />

after PHx in mice. Material and Methods: C57BL/6 mice<br />

were subjected to PHx (70%) or Sham operation. The proportion<br />

of Kupffer cell subsets in the remnant liver was examined<br />

by flow cytometry. To examine the role of CD11b + Kupffer cells,<br />

mice were depleted of these cells before PHx by non-lethal 5 Gy<br />

irradiation with or without bone marrow transplantation (BMT).<br />

To investigate the mechanism of macrophage (Mφ) accumulation<br />

into liver, mice were injected with CCR2 (MCP-1 receptor)<br />

antagonist and liver regeneration was evaluated. Results: The<br />

proportion of CD11b + Kupffer cells was greatly increased at<br />

three days after PHx when the hepatocytes vigorously proliferate,<br />

whereas the proportion of CD68 + Kupffer cells did not<br />

significantly change after PHx. Serum TNF levels peaked one<br />

day after PHx, and intracellular TNF staining revealed that<br />

CD11b + Kupffer cells are the main source of TNF. Additionally,<br />

the CD11b + Kupffer cells expressed surface FasL at three days<br />

after PHx. Serum MCP-1 levels (produced by CD68 + resident<br />

Kupffer cells) peaked at twelve hours after PHx. Irradiation eliminated<br />

the CD11b + Kupffer cells for approximately two weeks<br />

in the liver, while CD68 + Kupffer cells, NK cells and NKT cells<br />

remained, and hepatocyte regeneration was retarded. However,<br />

BMT partially restored CD11b + Kupffer cells and recovered<br />

the liver regeneration. Furthermore, CCR2 antagonist<br />

treatment decreased the CD11b + Kupffer cells and significantly<br />

retarded liver regeneration. Conclusion: The CD11b + Kupffer<br />

cells recruited from bone marrow via the MCP-1/CCR2 and<br />

TNF and FasL produced by these cells play a pivotal role in the<br />

liver regeneration after PHx. The possibility is raised that BMT<br />

can be applied to patients with hepatocellular carcinoma after<br />

partial hepatectomy for the acceleration of liver regeneration<br />

and prevention of hepatic failure.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hiroyuki Nakashima, Kiyoshi<br />

Nishiyama, Masahiro Nakashima, Manabu Kinoshita, Shuhji Seki


376A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

325<br />

CD45+ fraction in murine adipose tissue-derived stromal<br />

cells is therapeutical to Concanavalin A-induced<br />

murine acute hepatitis model.<br />

Akihiro Seki 1 , Yoshio Sakai 2 , Masatoshi Yamato 1,2 , Kosuke<br />

Ishida 1 , Hisashi Takabatake 1 , Alessandro Nasti 1 , Masao Honda 2 ,<br />

Shuichi Kaneko 1,2 ; 1 Disease control and homeostasis, Kanazawa<br />

university, Kanazawa, Japan; 2 Department of Gastroenterology,<br />

Kanazawa university hospital, Kanazawa, Japan<br />

Adipose tissue-derived stromal cells (ADSCs) are expected to<br />

be useful in regeneration therapy for liver disease because of<br />

their pluripotency and immunomodulatory capability. However,<br />

ADSCs are composed of heterogeneous cell fractions, such as<br />

mesenchymal stem cells, and the regenerative or therapeutic<br />

effects of all ADSC cell fractions have not been fully elucidated,<br />

particularly for liver diseases. In this study, we assessed<br />

whether the substantial ADSC CD45+ leukocyte-lineage fraction<br />

ameliorates the hepatic inflammatory disease condition<br />

induced by injecting mice with Concanavalin A (ConA). [Materials<br />

and Methods] ADSCs were obtained from inguinal subcutaneous<br />

adipose tissue of C57Bl/6 male mice by collagenase<br />

digestion. C57Bl/6 female mice were injected intravenously<br />

with 200 mg ConA to generate the murine hepatitis model.<br />

The CD45+ fraction, which was about 15% of all ADSCs, was<br />

isolated with a cell sorter (FACSAriaTMII), and 5 × 10 4 CD45+<br />

cells, 1 × 10 6 ADSCs, or PBS were injected intravenously into<br />

C57BL/6 mice 3 hours after the ConA injection (n = 5 each).<br />

Sera and liver tissues were obtained from the ConA-injected<br />

mice after 16 hours to measure serum alanine aminotransferase<br />

(ALT) and lactate dehydrogenase activities (LDH), and to conduct<br />

immunohistochemical and gene expression analyses using<br />

the real-time polymerase chain reaction. CD45+ cells (1 × 10 4<br />

cells) were also co-cultured with 1 × 10 5 splenocytes stimulated<br />

with ConA for 2 hours, followed by a gene expression analysis.<br />

[Results] Injecting 2 × 10 4 CD45+ cells into the ConA-stimulated<br />

hepatitis mice improved liver congestion compared with<br />

that of the PBS control. Correspondingly, injecting CD45+ cells<br />

into ConA-stimulated hepatitis mice suppressed serum ALT and<br />

LDH activities (P = 0.02 each) but increased hepatic alpha<br />

fetoprotein and albumin RNA expression (P = 0.01 and 0.02).<br />

These therapeutic effects of the CD45+ cells injection on the<br />

ConA-stimulated hepatitis mice were similar to those produced<br />

after ADSCs were injected. The immunohistochemical analysis<br />

showed that administrating CD45+ cells ameliorated the<br />

numbers of CD4+, CD11b+, and Gr-1+ cells infiltrating the<br />

livers of the ConA-stimulated hepatitis mice. Enhanced interferon-gamma,<br />

tumor necrosis factor-alpha, and chemokine (C-C<br />

motif) ligand 3 gene expression by ConA-stimulated splenocytes<br />

was suppressed when they were co-cultured with CD45+<br />

cells in vivo. Thus, the ADSC CD45+ fraction therapeutically<br />

suppressed ConA-induced mouse hepatitis. [Conclusion] The<br />

CD45+ leukocyte-lineage fraction in ADSCs, which is uniquely<br />

immunosuppressive and beneficial, may have important therapeutic<br />

effects in an acute hepatitis mouse model.<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Akihiro Seki, Yoshio Sakai,<br />

Masatoshi Yamato, Kosuke Ishida, Hisashi Takabatake, Alessandro Nasti,<br />

Masao Honda<br />

326<br />

Cross-talk between autophagy and the transcription<br />

factor KLF2 determines endothelial cell phenotype, liver<br />

microvascular function and hepatic damage after acute<br />

liver injury.<br />

Sergi Guixé-Muntet 1 , Fernanda C. de Mesquita 1,3 , Sergi Vila 1 ,<br />

Carmen Peralta 2 , Juan Carlos Garcia-Pagan 1 , Jaime Bosch 1 , Jordi<br />

Gracia-Sancho 1 ; 1 Barcelona Hepatic Hemodynamic Lab, Hospital<br />

Clinic de Barcelona - IDIBAPS - CIBEREHD, Barcelona, Spain;<br />

2 IDIBAPS - CIBEREHD, Barcelona, Spain; 3 Laboratório de Biofísica<br />

Celular e Inflamação, PUCRS, Porto Alegre-RS, Brazil<br />

Background & aims: The transcription factor Kruppel-like Factor<br />

2 (KLF2) can be induced by different drugs, including simvastatin,<br />

and its expression confers a vasoprotective phenotype to<br />

the endothelium. Considering recent data suggesting activation<br />

of autophagy by statins, we aimed at: 1) characterizing the<br />

molecular relationship between autophagy and KLF2 in the<br />

endothelium, 2) assessing this relationship in acute liver injury<br />

due to cold ischemia/warm reperfusion (I/R), and 3) studying<br />

the effects of modulating KLF2-autophagy in in vitro and ex<br />

vivo models of acute liver injury. Methods: 1) In vitro: Autophagy<br />

and KLF2 were modulated and characterized in vascular<br />

endothelial (HUVEC) and sinusoidal endothelial cells (LSEC)<br />

cultured in standard conditions or under I/R. KLF2 was induced<br />

by simvastatin and resveratrol or genetically regulated with<br />

siRNA or adenovirus in the presence or absence of autophagy<br />

modulators. Cell viability was evaluated by double-staining<br />

with acridine orange and propidium iodide, and by trypan<br />

blue exclusion. Autophagic flux was assessed by western blot<br />

of LC3B, immunofluorescence of accumulation of autophagosomes,<br />

and colocalization of autophagosomes & lysosomes. 2)<br />

Ex vivo: Wistar rats received a) vehicle, b) simvastatin (1mg/<br />

kg, i.v.) or c) chloroquine (inhibitor of autophagy; 60mg/kg,<br />

i.p.) + simvastatin. Livers were explanted, cold stored in Wisconsin<br />

solution (16h) and warm-reperfused (2h). Microvascular<br />

function was assessed by liver vasorelaxation to acetylcholine.<br />

Results: A positive feed-back between autophagy and KLF2<br />

was observed: KLF2-inducers simvastatin and resveratrol, but<br />

not the autophagy activator Rapamycin, caused endothelial<br />

KLF2 overexpression through an autophagy-Rac1-rab7-dependent<br />

mechanism. In turn, KLF2 induction promoted further<br />

activation of autophagy. Cold ischemia blunted autophagic<br />

flux in LSEC. Upon reperfusion, LSEC cold stored in Celsior<br />

solution showed proper reactivation of autophagy. However,<br />

LSEC stored in Wisconsin failed to reactivate autophagy, most<br />

probably due to impairment in autophagosome and lysosome<br />

fusion. Simvastatin re-activated autophagy by up-regulating<br />

rab7 (restoring autolysosome formation), which resulted in<br />

increased KLF2 levels, improved cell viability (in vitro), and<br />

ameliorated hepatic damage and microvascular function (ex<br />

vivo). Conclusions: We herein describe for the first time the<br />

complex link between autophagy and KLF2 modulating the<br />

phenotype and survival of the endothelium. These results help<br />

understanding the underlying molecular mechanisms of the<br />

protection conferred by KLF2-inducers, such as simvastatin, in<br />

hepatic and extra-hepatic vascular disorders.<br />

Disclosures:<br />

Juan Carlos Garcia-Pagan - Consulting: Novartis; Grant/Research Support:<br />

GORE<br />

Jaime Bosch - Consulting: Falk, Gilead Science, Intercept Therapeutics, Conatus<br />

Pharmaceuticals, Exalenz, Almirall, Chiasma<br />

The following authors have nothing to disclose: Sergi Guixé-Muntet, Fernanda C.<br />

de Mesquita, Sergi Vila, Carmen Peralta, Jordi Gracia-Sancho


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 377A<br />

327<br />

Histone 3 lysine 9 trimethylation (H3K9me3) and<br />

inflammasome as novel senescent associated markers<br />

of drug induced premature senescence in hepatocellular<br />

carcinoma cell lines<br />

Bijoya Sen 1 , Tarique Anwar 2 , Chhagan Bihari 3 , Archana Rastogi 3 ,<br />

Nirupma Trehanpati 1 , Sukriti Sukriti 1 , Shvetank Sharma 1 , Shiv K.<br />

Sarin 4 , Gayatri Ramakrishna 1 ; 1 Dept of Research, Institut of Liver<br />

and Biliary Scienc, Delhi, India; 2 Center for DNA fingerprinting<br />

and diagnostics, Hyderabad, India; 3 department of pathology,<br />

institute of liver and biliary sciences, New Delhi, India; 4 institute of<br />

liver and biliary sciences, New Delhi, India<br />

Abstract Chemotherapeutic drugs can induce premature<br />

senescence in cancerous cells. The p53-p21 and pRb-p16 axis<br />

are the most well studied mechanisms involved in induction<br />

of senescence. However, these pathways are also associated<br />

with cell cycle arrest conditions of quiescence, differentiation<br />

and also cell death. Purpose of this study was to evaluate<br />

novel markers specific to premature senescence. Hepatocellular<br />

carcinoma cell lines (Huh7, Hep3B and HepG2) when<br />

treated with low dose of doxorubicin, a well known chemotherapeutic<br />

agent, showed permanent cell cycle arrest in G2/M<br />

phase. Compared to proliferating cells, the senescent cells<br />

showed enlarged cytomorphology, increased expression of<br />

cyclin dependent kinase (p21) and positivity for SA-β-galactosidase,<br />

all indicative of premature senescence. To analyze<br />

a novel structural marker associated with senescence, transmission<br />

electron microscopy (TEM) was done which showed<br />

accumulation of autophagic vacuoles, condensed chromatin<br />

and prominent mitochondrial-endoplasmic reticulum (ER)<br />

coupling which were conspicuosly absent in non-senescent<br />

hepatocytes. The changes in chromatin was further evaluated<br />

by immunofluorescence microscopy which showed presence of<br />

heterochromatic foci only in the senescent cells. Hetrochromatin<br />

formation in turn involves histone modification and senescent<br />

cells showed a significant increase in inactive chromatin mark<br />

viz., H3K9 trimethylation compared to non-senescent cells. Further,<br />

microarray based gene expression analysis was done to<br />

identify novel pathways uniquely associated with senescence<br />

which showed a differential increase in cytokine-chemokine<br />

response. This in turn was supported by the fact that the secretome<br />

of senescent cell was highly effective in closure of wound<br />

as assessed by the wound healing assay, unlike the control<br />

cell secretome. The active senescent secretome was associated<br />

with increased transcript level of PyCard/Asc, and IL-1β indicative<br />

of inflammasome activation. In conclusion we have now<br />

identified inflammasomes and H3K9 trimethylation as specific<br />

senescence associated markers. This project was funded by<br />

Dept of Biotechnology, India.<br />

328<br />

Netrin-1 Promotes Liver Cancer Cells Collective Invasion<br />

in a 3D Cell Culture Model<br />

Ping Han 1 , Yu Fu 2 , Jingmei Liu 1 , Dongxiao Li 1 , Yunwu Wang 1 ,<br />

Dean Tian 1 , Wei Yan 1 ; 1 Gastroenterology, Tongji Hospital, Tongji<br />

Medical College, Huazhong University of Science and Technology,Wuhan,China,<br />

Wuhan, China; 2 Gastroenterology, Union Hospital,<br />

Tongji Medical College, Huazhong University of Science and<br />

Technology,Wuhan,China, Wuhan, China<br />

Collective invasion is the new understanding of tumor metastasis,<br />

while the mechanism needs to be further studied. Our<br />

previous <strong>studies</strong> showed that Netrin-1 promoted the invasion<br />

ability of liver cancer cells(LCCs). Immunohistochemical <strong>studies</strong><br />

showed that LCCs invaded surrounding tissue as collective<br />

invasion. The expression of Netrin-1 in these LCCs was<br />

increased. In this study, we performed a 3D cell culture model<br />

to further confirm the role of Netrin-1 in LCCs collective invasion.<br />

We designed a 3D cell culture model, after centrifuge,<br />

5×10 6 cells were resuspended by 120μl 10% sucrose solutions,<br />

and quickly mixed with equal volume of 0.5% hydrogel<br />

solution( BeaverNano ), the cells were then seeded in a Class<br />

Bottom Cell Culture Dish(NEST), and allowed to grow for six<br />

days. Then, cells were observed under a confocal microscopy.<br />

The number and distance of collective invasion in LCCs were<br />

evaluated after enhanced or silenced expression of Netrin-1.<br />

The N-cadherin shRNAs were used to verify the role of N-cadherin-based<br />

junctions in Netrin-1 promoted collective invasion.<br />

We found the number and distance of invaded SK-Hep1 cells,<br />

a high metastasis potential cell line, were significantly more<br />

than Huh7 cells with low metastasis potential, the normal liver<br />

cell line LO2 cells nearly had no cell invaded. Meanwhile,<br />

the invaded cells had significant more expression of Netrin-1.<br />

Overexpression of Netrin-1 increased collective invasion in the<br />

3D cell culture model and elevated N-cadherin expression in<br />

Huh7 cells, while stable knockdown of Netrin-1 remarkably<br />

inhibited collective invasion and decreased N-cadherin expression<br />

in SK-Hep1 cells. Interestingly, knockdown N-cadherin in<br />

Huh7 cells significantly diminished Netrin-1 promoted liver cancer<br />

cell collective invasion. These results suggest that Netrin-1<br />

enhances N-cadherin junctions to promote LCCs collective invasion,<br />

subsequently may increase LCCs metastasis.(This study<br />

is supported by the National Natural Science Foundation of<br />

China (81472311), the Fundamental Research Funds for the<br />

Central Universities (2014ZHYX020,2014TS077), the Project<br />

sponsored by SRF for ROCS, SEM (2014-1685) and Hubei<br />

Province health and family planning scientic research project<br />

(WJ2015Q006))<br />

Collective invasion of SK-Hep1 cells in 3D cell culture<br />

Disclosures:<br />

The following authors have nothing to disclose: Ping Han, Yu Fu, Jingmei Liu,<br />

Dongxiao Li, Yunwu Wang, Dean Tian, Wei Yan<br />

Disclosures:<br />

The following authors have nothing to disclose: Bijoya Sen, Tarique Anwar,<br />

Chhagan Bihari, Archana Rastogi, Nirupma Trehanpati, Sukriti Sukriti, Shvetank<br />

Sharma, Shiv K. Sarin, Gayatri Ramakrishna


378A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

329<br />

Interaction with collagen matrix modulates macrophage<br />

expression of MMPs in part by controlling cell shape.<br />

Toshihiko Matsumoto 1 , Stephen Jenkins 3 , Koichi Fujisawa 2 , Taro<br />

Takami 2 , Naoki Yamamoto 2 , John P. Iredale 3 , Isao Sakaida 2 ;<br />

1 Department of Gastroenterology & Hepatology, Department of<br />

Oncology and Laboratory Medicine, Yamaguchi University Graduate<br />

School of Medicine, Ube, Japan; 2 Department of Gastroenterology<br />

& Hepatology, Yamaguchi University Graduate School of<br />

Medicine, Ube, Japan; 3 The Queen’s Medical Research Institute,<br />

University of Edinburgh, Edinburgh, United Kingdom<br />

Background and Objectives: Macrophages have a pivotal<br />

role in resolution of liver fibrosis, during which, pro-fibrotic<br />

macrophages undergo a switch in phenotype and up-regulate<br />

expression of MMP9, MMP12 and MMP13, which promote<br />

myofibroblast apoptosis and degradation of the scar extracellular<br />

matrix (ECM). However, the signals that regulate this balance<br />

of opposing functions remain to be fully determined. It was<br />

recently reported that cell shape had an important role in regulating<br />

macrophage function. We hypothesized that alterations<br />

in cell shape associated with changes in ECM architecture may<br />

provide cues to modulate the phenotype and MMPs expression<br />

of macrophages in fibrotic liver. We therefore examined the<br />

effects of collagen on macrophage shape and MMP expression<br />

in vitro, and investigated the underlying mechanisms linking<br />

these processes with the aim of developing future anti-fibrotic<br />

therapies. Method: Mouse bone marrow derived macrophages<br />

were treated with GM-CSF (GM-BMDM) to model inflammatory<br />

macrophages and subsequently cultured on a monolayer of<br />

collagen I, plastic alone or BSA control and the shape, size<br />

and MMP expression determined. To clarify the relationship<br />

between effects on cell shape and MMPs expression, actin<br />

polymerization of GM-BMDM cultured on Collagen I or plastic<br />

was assessed by phalloidin staining. GM-BMDM were then<br />

treated with Cytochalasin D and Rho family GTPase inhibitors,<br />

such as NSC23766 and ML141, to inhibit actin polymerization<br />

and determine the effect on MMP expression. Result: Culture<br />

on collagen I inhibited expression of MMP9, 12 and 13<br />

by GM-BMDM (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 379A<br />

331<br />

Genetic Induction of Metabolically Functional, Polarized<br />

Cultures of Proliferating Human Hepatocytes<br />

Gahl Levy 1 , Stefan Heinz 3 , Ella H. Sklan 2 , Oren Shibolet 2 , Joris<br />

Braspenning 3 , Yaakov Nahmias 1 ; 1 Bioengineering, The Hebrew<br />

University of Jerusalem, Jerusalem, Israel; 2 Tel Aviv University, Tel<br />

Aviv, Israel; 3 Medicyte, Heidelberg, Germany<br />

Primary hepatocytes are responsible for the modification, clearance,<br />

and transformational toxicity of most xenobiotics. Regretfully,<br />

like other primary cells, hepatocytes do not proliferate<br />

in vitro, while their transformation causes loss of metabolic<br />

function. Here, we report a method to conditionally induce<br />

proliferation in primary human hepatocytes by genetic modification<br />

(upcyte ® process). Proliferating human hepatocytes can<br />

undergo up to 40 population doublings, producing 10 13 to<br />

10 16 cells, and differentiate when contact-inhibited into metabolically<br />

functional, polarized cells with functional bile canaliculi.<br />

The cells lack fetal and tumor markers and show high levels<br />

of CYP3A4 expression and function. Differentiated cells show<br />

gene expression, CYP450 activity, induction, and TC 50<br />

toxicity<br />

profile comparable to primary human hepatocytes. Importantly,<br />

proliferating hepatocytes can be readily infected with Hepatitis<br />

C Virus (HCV) showing replication and infectivity profile comparable<br />

to Huh7.5.1 cells. Our results offer a promising new<br />

technology to expand human hepatocytes from varying genetic<br />

backgrounds for scientific research, clinical applications, and<br />

pharmaceutical development.<br />

Disclosures:<br />

Oren Shibolet - Advisory Committees or Review Panels: MSD, Abbvie, Novartis;<br />

Grant/Research Support: MSD<br />

The following authors have nothing to disclose: Gahl Levy, Stefan Heinz, Ella H.<br />

Sklan, Joris Braspenning, Yaakov Nahmias<br />

332<br />

Hepatocyte Isolation from Liver Fine Needle Aspiration<br />

and Core Needle Biopsy<br />

Andrew H. Talal 1,2 , Rosemary Furlage 3 , Yvonne J. Woolwine-Cunningham<br />

1 , Jun Li 4 , Robin Difrancesco 4,5 , Gene D. Morse 4,5 , Georg<br />

Lauer 6 , Jun Qu 4 , Paul K. Wallace 3 ; 1 Center for Clinical Care and<br />

Research in Liver Disease, SUNY-Buffalo, Buffalo, NY; 2 Division of<br />

Gastroenterology, Hepatology and Nutrition, Department of Medicine,<br />

SUNY-Buffalo, Buffalo, NY; 3 Department of Flow & Image<br />

Cytometry, Roswell Park Cancer Institute, Buffalo, NY; 4 School of<br />

Pharmacy and Pharmaceutical Sciences, SUNY-Buffalo, Buffalo,<br />

NY; 5 NYS Center of Excellence in Bioinformatics and Life Sciences,<br />

University at Buffalo, Buffalo, NY; 6 Gastrointestinal Unit, Massachusetts<br />

General Hospital, Boston, MA<br />

Background: Although the liver is the major site of drug metabolism,<br />

little is known about the determinants of intrahepatic drug<br />

concentration. We sought to develop techniques for hepatocyte<br />

isolation for measurement of intrahepatic drug concentration.<br />

Methods: Thirteen Sprague Dawley rats had liver specimens<br />

collected via FNA (21G needle), core needle biopsy (CNB)<br />

(16G needle), and surgical resection. Blood was collected by<br />

cardiac puncture at the same time. The optimized procedure<br />

initially requires the addition of chilled washing solution to<br />

the sample and centrifugation at 150G for 3 min at 4 o C. Subsequently,<br />

the sample is incubated with collagenase at 37 o C<br />

while rocking for 30 minutes, pushed through a 100m filter,<br />

centrifuged and washed twice more. A sample aliquot is subsequently<br />

counted (Countess, Invitrogen) and analyzed (Fortessa,<br />

Becton Dickinson). On one animal, proteins were initially<br />

extracted with a Polytron sonication procedure, separated using<br />

a nano-LC system coupled to an Orbitrap Fusion Mass Spect on<br />

a 100-cm-long column. Results: Average length and weight of<br />

the CNB samples was 2.0 cm and 10 mg, respectively, while<br />

FNA samples ranged from 3-12 mg. The total number of cells<br />

isolated, the percentage of hepatocytes in the sample, and the<br />

number of sorted hepatocytes is shown (Table). Combining<br />

three FNA samples together increased cell yield to 222,308<br />

with 46% hepatocytes resulting in 89,358 sorted hepatocytes.<br />

Combining five FNA samples together resulted in a very similar<br />

number of sorted hepatocytes. Samples weighing 8.0 and<br />

20.6 mgs yielding 427.2 and 2670.6 mg, respectively, were<br />

used for proteomic analysis. A total of 879 and 1299 proteins<br />

were isolated from the FNA and CNB samples, respectively, of<br />

which those involved in oxidative reduction and responses to<br />

organic substances had the highest percentages (FNA: 34%,<br />

CNB: 28%). Mitochondrion and cytosol protein analytes were<br />

the most frequently identified cellular components (FNA: 60%;<br />

CNB: 52%). Conclusions: One FNA pass results in sufficient<br />

hepatocyte yield that should be suitable for measurement of<br />

intrahepatic drug concentration and for proteomic analysis.<br />

Combining three FNA passes into the same sample increases<br />

the cell yield proportionally. Isolation of nonparenchymal cells<br />

from the same sample should also be feasible.<br />

Cell Yields<br />

SD, standard deviation; Hep, hepatocytes<br />

Disclosures:<br />

Andrew H. Talal - Advisory Committees or Review Panels: Merck and Co, Abbvie,<br />

Gilead; Board Membership: Pfizer; Grant/Research Support: Merck and Co,<br />

Gilead, Abbott Molecular, Abbvie<br />

The following authors have nothing to disclose: Rosemary Furlage, Yvonne J.<br />

Woolwine-Cunningham, Jun Li, Robin Difrancesco, Gene D. Morse, Georg Lauer,<br />

Jun Qu, Paul K. Wallace<br />

333<br />

Characterization of Microtubule Motor Proteins in the<br />

Endocytic and Autophagic Compartments During Nutritional<br />

Stress and Aging<br />

Eloy Bejarano-Fernandez, John W. Murray, Xintao Wang, Ana<br />

Maria Cuervo, Allan W. Wolkoff; Albert Einstein College of Medicine,<br />

Bronx, NY<br />

Purpose: These <strong>studies</strong> were designed to survey and functionally<br />

analyze microtubule motor proteins present on autophagic<br />

and endocytic organelles from liver in response to nutritional<br />

stress and aging. Methods: Organelles were purified from livers<br />

of 3-month and 22-month old mice; motor proteins and<br />

organelle markers were analyzed by western blot. Primary<br />

fibroblasts generated from the same animals were subject to fluorescent<br />

microscope imaging and automated analyses. Results:<br />

A detailed analysis of motor proteins present on intracellular<br />

organelles revealed a high association of dynein with autophagosomes<br />

(Enrichment Versus Homogenate (EVH): 2.39 ± 0.28)<br />

compared to endosomes (EVH 1.01 ± 0.37), and lysosomes<br />

(EVH 0.60 ± 0.06). Kif5b (kinesin-1) showed more equal distribution<br />

between organelles (EVH: 0.57 ± 0.30, 0.27 ± 0.04,<br />

0.22 ± 0.09 for endosomes, autophagosomes and lysosomes,<br />

respectively). Kif3A (kinesin-2) was abundant on endosomes<br />

(EVH: 1.61 ± 1.57) and autophagosomes (EVH: 0.56 ± 0.15)<br />

but not lysosomes (EVH: 0.14 ± 0.03). Since dynein was relatively<br />

low in liver lysosomes, minus end directed kinesins were<br />

examined. Among these, KifC3 was strongly enriched in lysosomes<br />

(EVH: 12.08 ± 4.74) as compared to autophagosomes<br />

(EVH: 0.56 ± 0.35) and endosomes (EVH: 1.12 ± 0.50). Aged<br />

mice were found to exhibit an 84.5% decrease in the level


380A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of dynein in autophagosomes (p=0.002). Interestingly, the<br />

age-dependent differential recruitment is lost under acute starvation.<br />

A similar behavior was also found for the KifC3 motor<br />

in lysosomes. Live cell vesicle tracking revealed differences in<br />

intracellular motility that correlated with these findings. The<br />

motility of autophagosomes was reduced by 16% in cells from<br />

aged mice (p=0.03, p>0.05). In addition, a dynein inhibitor<br />

(EHNA) inhibited autophagosome motility (-24.6%, p=0.02)<br />

exclusively in young mice. Upon serum starvation, no differences<br />

were found in autophagosomal motility pointing out the<br />

recruitment of motors during nutrient deprivation. In agreement<br />

with these results, in vitro assays showed that on average 22%<br />

of the purified organelles moved along microtubules whereas<br />

30% of autophagolysosomes from starved mice were motile<br />

(p=0.04). Conclusions: Our analysis provides for first time<br />

quantitative biochemical and imaging evidence for the impact<br />

of aging in the association of motors to degradative compartments<br />

under basal and nutritional stress-induced autophagy.<br />

Disclosures:<br />

Allan W. Wolkoff - Consulting: Synageva; Grant/Research Support: Merck<br />

The following authors have nothing to disclose: Eloy Bejarano-Fernandez, John<br />

W. Murray, Xintao Wang, Ana Maria Cuervo<br />

334<br />

Acute alcohol binge causes skeletal muscle mitochondrial<br />

intermediate metabolite deficiency<br />

Allawy Allawy 2,3 , Gangarao Davuluri 2 , Avinash Kumar 2 , Bin Gao 4 ,<br />

Ming-Jiang Xu 4 , Megan R. McMullen 2 , Rebecca L. McCullough 2 ,<br />

Srinivasan Dasarathy 1,2 ; 1 Department Of Gastroenterology and<br />

Hepatology, Cleveland Clinic, Cleveland, OH; 2 Department Of<br />

Pathobiology, Cleveland Clinic, Cleveland, OH; 3 Department Of<br />

Internal medicine, Cleveland Clinic, Cleveland, OH; 4 Laboratory<br />

of Liver Diseases. NIAAA, NIH, Bethesda, MD<br />

Introduction Binge drinking is believed to result in more severe<br />

tissue injury than chronic ethanol exposure. There are no data<br />

on the comparing binge drinking and chronic ethanol exposure<br />

on skeletal muscle mitochondrial function. We used multiple<br />

models of in vivo and an in vitro model of skeletal muscle<br />

ethanol exposure to determine its impact on tricarboxylic acid<br />

cycle intermediates. Materials and Methods. Murine C2C12<br />

myotubes were exposed to 100mM ethanol for 6 h and 24 h<br />

and lysates used to quantify TCA cycle intermediates. In vivo<br />

<strong>studies</strong> were performed in mice exposed to different patterns of<br />

ethanol exposure and included chronic alcohol feeding for 25<br />

or 56 days, ethanol feeding for 10 days followed by a single<br />

binge, twice weekly ethanol feeding for 8 weeks, pair fed<br />

or maltose fed controls. Gastrocnemius muscle was harvested<br />

and TCA cycle intermediates extracted using ethyl acetate,<br />

tertbutyldimethylsilyl derivatives generated and quantified by<br />

gas chromatography-mass spectrometry using protocols established<br />

in our laboratory. All experiments were performed in<br />

3-4 mice in each group and 6 independent experiments in<br />

myotubes. All data are expressed as mean±SD of percentage<br />

change after normalization against the concentrations in the<br />

appropriate controls. Results. In the chronic ethanol exposure<br />

for either 25 or 56 days, no significant changes were observed<br />

in the skeletal muscle concentration of citrate, αketoglutarate,<br />

succinate, fumarate or malate. Interestingly, compared to pair<br />

fed controls (100%), in the single binge model, concentration<br />

of citrate was 49.7±8.9% (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 381A<br />

hepatic injury, increase macrophage recruitment, and increase<br />

hepatic expression of CXCL1 following IRI. The dramatic reduction<br />

in tissue injury by MSC-EV support the therapeutic use of<br />

MSC-EV to reduce hepatic IRI following hepatic surgery or in<br />

hepatic transplantation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hiroaki Haga, David D. Lee,<br />

Sarah Nix, Irene K. Yan, Tushar Patel<br />

336<br />

Liver Fabrication Using Decellularized Whole-Organ<br />

Scaffold for Transplantable Graft in large animal model<br />

Hiroshi Yagi, Kenta Inomata, kazuki tajima, Taizo Hibi, Yuta Abe,<br />

Minoru Kitago, Masahiro Shinoda, Osamu Itano, Yuko Kitagawa;<br />

Surgery, Keio University School of Medicine, Shinjuku-ku, Japan<br />

Background: Whole-organ scaffold generated by tissue decellularization<br />

technology can be one of the solutions to fabricate<br />

transplantable organ graft, which contains functional cell<br />

types including progenitor cells. However, the feasibility of this<br />

technology for human-scale application, and the biological<br />

alterations of the scaffold after implantation are still unclear.<br />

Methods: To determine if the decellularization methodology<br />

could be applied in large animal model and it could be sufficiently<br />

recellularized with primary cells as well as human iPS<br />

cell derived hepatic progenitor cells, the decellularized scaffold<br />

was first generated using SDS, TrytonX-100 and CHAPS.<br />

Next, endothelial cells and primary porcine hepatocytes were<br />

infused into the scaffold under the monitoring of intravascular<br />

pressure. Finally, the fabricated liver was implanted into the<br />

porcine body and evaluated by angiography after vascular<br />

anastomoses. Histological study was performed to evaluate cell<br />

infiltration, degree of coagulation and adhesion of the scaffold.<br />

In addition, the liver scaffold containing human iPS cell derived<br />

liver progenitor cells was transplanted into porcine body under<br />

immune-suppression. Results: The endothelial cells could cover<br />

most of the vessel lumen of the scaffold and the primary hepatocytes<br />

were well distributed into the parenchymal space. The<br />

fabricated liver could be successfully transplanted into porcine<br />

body connecting with portal vein and IVC. The graft was well<br />

perfused and preserved in the porcine abdominal cavity without<br />

bleeding or absorption. Interestingly, histological analysis<br />

of transplanted scaffold with human iPS derived progenitor<br />

cells revealed liver specific architecture with ALB positive cells<br />

in parenchymal space. Conclusions: We could scale-up and<br />

optimize the liver fabrication system to apply not only primary<br />

cells but also human iPS derived cells into clinically feasible<br />

model.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hiroshi Yagi, Kenta Inomata,<br />

kazuki tajima, Taizo Hibi, Yuta Abe, Minoru Kitago, Masahiro Shinoda, Osamu<br />

Itano, Yuko Kitagawa<br />

337<br />

Three-dimensional bioactive human liver acellular scaffold<br />

with preserved architecture and biomechanical<br />

properties<br />

Giuseppe Mazza 1 , Walid Al-Akkad 1 , Lisa Longato 1 , Andrea<br />

Telese 1 , Andrew R. Hall 1 , Luca Urbani 2 , Benjamin Robinson 3 ,<br />

Giusi Marrone 1 , Oliver Willacy 1 , Luca Frenguelli 1 , Marco Curti 1 ,<br />

Massimo Malago 1 , Tu Vinh Luong 1 , Kevin Moore 1 , Armando E.<br />

Del Rio Hernandez 3 , Paolo De Coppi 2 , Krista Rombouts 1 , Massimo<br />

Pinzani 1 ; 1 Institute for Liver and Digestive Helath, UCL, London,<br />

United Kingdom; 2 Institute for Child Health, UCL, London, United<br />

Kingdom; 3 Imperial College, London, United Kingdom<br />

Background: Human tissue engineering combines cells and<br />

extracellular matrix (ECM) saffolds for the development<br />

of 3D-structure in order to regenerate tissues/organs and<br />

to recapitulate disease in vitro. The key challenge in tissue<br />

engineering is the development of biomaterials that mimic<br />

the complexity of human tissue. Techniques for tissue decellularization<br />

have been introduced in order to obtain natural<br />

acellular scaffold. Optimal scaffolds should be characterised<br />

by preserved ECM integrity, bioactivity and three-dimensional<br />

organisation. Aim: The aim of this study was to develop a rapid<br />

protocol for the decellularisation of small samples of healthy<br />

human liver and demonstrate preservation of ECM composition,<br />

3D-architecture, bioactivity and biomechanical properties.<br />

In addition, repopulation with cultured human liver cell lines,<br />

namely hepatoblastoma (HepG2), hepatic stellate (LX2) cells,<br />

and human umbilical vein endothelial cells (HUVEC). Methods:<br />

Decellularisation of healthy human liver cubes (i.e. 125mm 3 )<br />

was completed within 3 hours of agitation in solutions with<br />

detergents and enzymes. The decellularisation efficiency was<br />

determined by immunohistochemistry for ECM components and<br />

residual DNA, scanning electron and second harmonic generation<br />

microscopy, proteomic, raman spectroscopy, atomic force<br />

microscopy and chorioallantoic membrane assay. RESULTS:<br />

This innovative protocol resulted in liver cube scaffolds with<br />

a preserved 3D structure and ECM composition, while DNA<br />

and cellular residues were successfully removed. Tissue stiffness<br />

was preserved in the decellularized tissues (2KPa) and<br />

the acellular liver maintained the capability to induce neo-angiogenesis<br />

after 7 days. Interestingly, HUVEC repopulated the<br />

decellularised vessels within the liver scaffold expressing functional<br />

marker such as FVIII. Acellular human liver repopulated<br />

with LX2 and HepG2 cells showed remarkable difference in<br />

gene and protein expression when compared with 2D-system.<br />

COL1A1 gene was less expressed in LX2 cells when cultured<br />

in 3D while TGFB1 and LOX were increased. Notably, albumin<br />

mRNA expression was strongly upregulated in HepG2<br />

cells after 14 days in 3D-culture. CONCLUSION: This is the first<br />

report describing a protocol for fast human liver tissue cubes<br />

decellularization. The decellularization protocol maintained the<br />

natural 3D structure and ECM composition and organisation of<br />

human liver tissue. This is a key advance in the development<br />

of 3D technologies for the study of liver disease as well as for<br />

the development of 3D-carrier for hepatocyte transplantation.<br />

Disclosures:<br />

Kevin Moore - Advisory Committees or Review Panels: Servier<br />

Massimo Pinzani - Advisory Committees or Review Panels: Intercept Pharmaceutical,<br />

Silence Therapeutic, Abbot; Consulting: UCB; Speaking and Teaching:<br />

Gilead, BMS


382A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Giuseppe Mazza, Walid Al-Akkad,<br />

Lisa Longato, Andrea Telese, Andrew R. Hall, Luca Urbani, Benjamin Robinson,<br />

Giusi Marrone, Oliver Willacy, Luca Frenguelli, Marco Curti, Massimo<br />

Malago, Tu Vinh Luong, Armando E. Del Rio Hernandez, Paolo De Coppi, Krista<br />

Rombouts<br />

338<br />

Promotion of liver transplant tolerance by hepatic plasmacytoid<br />

dendritic cells<br />

Osamu Yoshida 2,1 , Shoko Kimura 2 , Benjamin M. Matta 2 , Yusuke<br />

Imai 1 , Masanori Abe 1 , Yoichi Hiasa 1 , Angus W. Thomson 2,3 ;<br />

1 Department of Gastroenterology and Metabology, Ehime University<br />

Graduate School of Medicine, Ehime, Japan; 2 Starzl<br />

Translantation Institute, University of Pittsburgh School of Medicine,<br />

Pittsburgh, PA; 3 Department of Immunology, University of<br />

Pittsburgh School of Medicine, Pittsburgh, PA<br />

Background: Plasmacytoid dendritic cells (pDC) are an<br />

important source of interferon-alpha and serve as a first line<br />

of defense against viral infections. On the other hand, pDC,<br />

especially liver pDC, display tolerogenic properties and play<br />

an important role in induction of oral tolerance. The liver is<br />

both an immune and tolerogenic organ that is accepted without<br />

immunosuppressive therapy when transplanted across MHC<br />

barriers in mice. We have reported previously that myeloid<br />

DC can promote the induction of transplant tolerance in mouse<br />

models. Here we hypothesized that liver pDC might play an<br />

important role in immune regulation and the promotion of liver<br />

transplant tolerance. Methods: Liver pDC were depleted from<br />

10-12 week-old male C57BL/6 (B6; H2 b ) mice by intravenous<br />

injection of mPDCA (120G8) mAb. Profound pDC depletion<br />

was confirmed by flow cytometry. Untreated or pDC-depleted<br />

B6 livers were transplanted orthotopically into normal 10-12<br />

week-old male C3H (H2 k ) mice. Graft rejection was assessed<br />

by recipient survival and serum ALT levels and histology on day<br />

4 after transplantation. We also assessed the impact of pDC<br />

on alloimmunity using a delayed-type hypersensitivity model.<br />

Normal C3H mice were sensitized with normal or pDC-depleted<br />

donor (B6) splenocytes (10 7 , sc) 2 weeks before ex<br />

vivo re-stimulation of host T cells with donor antigen. T cell<br />

proliferation and cytokine production were assessed by CFSE-<br />

MLR and ELISA, respectively. Results: More than 90% of pDC<br />

were depleted in both liver and spleen. Whereas mice transplanted<br />

with normal livers (n= 6) survived >100 days without<br />

any immunosuppression, 83% (n=5/6) mice that received<br />

pDC-depleted livers died within 55 days (MST: untreated:<br />

>100d vs pDC-depleted 25d; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 383A<br />

340<br />

Novel Vasodilator and Anti-inflammatory Drug-based<br />

Strategy for Restoring Ischemia/ reperfusion-Exposed<br />

Donor Liver to Healthy State<br />

Thirunavukkarasu Chinnasamy 1,2 , Fadi L. Jaber 1 , Sanjeev Gupta 3 ;<br />

1 Department of Medicine, Albert Einstein College of Medicine,<br />

Bronx, NY, USA, Bronx, NY; 2 Department of Biochemistry and<br />

Molecular Biology, Pondicherry University, Puducherry, India;<br />

3 Departments of Medicine and Pathology, Marion Bessin Liver<br />

Research Center, Diabetes Center, Cancer Center, Ruth L. and<br />

David S. Gottesman Institute for Stem Cell and Regenerative Medicine<br />

Research, Albert Einstein College of Medicine, Bronx, NY<br />

Maintaining donor organs in healthy state is critical for isolating<br />

viable cells and for liver transplantation. A major cause of<br />

donor organ deterioration is ischemia/reperfusion (IR) injury<br />

due to complex events, e.g., vasoconstriction or inflammation<br />

with release of ROS, cytokines/chemokines/receptors,<br />

etc. We hypothesized that donor organ preconditioning with<br />

early control of vasoconstriction and inflammation by NO and<br />

TNF-α blockade as master regulators of interest will be effective.<br />

Therefore, we used NO donor, nitroglycerin (NG), and<br />

TNF-α blocker, Etanercept (ETN), before nonlethal hepatic IR in<br />

F344 rats with clamping of portal and hepatic blood flow for<br />

15 min and reperfusion for 24h. In drug-untreated controls, IR<br />

produced much oxidative stress, including hepatic GGT expression,<br />

nitrotyrosine, 8-oxo-dG DNA adducts and liver necrosis,<br />

as expected. In hepatocytes isolated from donors with IR, cell<br />

viability and attachment in dishes was inferior to control donor<br />

cells without IR. However, cultured cells from IR donors better<br />

withstood secondary H 2<br />

O 2<br />

or CCl 4<br />

injury. By contrast, preconditioning<br />

with ETN 12-18h before I/R and/or NG starting<br />

5min before IR and continuing during and after IR for 20 min<br />

lessened hepatic injury. ETN+NG combination was most effective<br />

versus either drug alone in restoring IR liver to near-normal<br />

without necrosis or GGT and nitrotyrosine expression. Hepatocytes<br />

from ETN+NG-preconditioned rats with IR were more<br />

viable, attached better in dishes, and were also more sensitive<br />

to H 2<br />

O 2<br />

and CCl 4<br />

injury, which was similar to healthy control<br />

hepatocytes. FACS analysis showed that polyploid hepatocytes<br />

were depleted in cells from IR rats but not in cells from healthy<br />

control and ETN+NG preconditioned donors, again indicating<br />

less oxidative insult. Next, to further examine potential of<br />

isolated donor cells, we performed cell transplantation assays<br />

in DPPIV- rats. Hepatocytes from healthy donor rats engrafted<br />

and were normally distributed in periportal areas in zones<br />

1/2 of liver lobules. However, hepatocytes from IR donor rats<br />

engrafted less and were mainly in portal vein radicles, suggesting<br />

differences in their cell adhesion and matrix-interacting<br />

properties. By contrast, engraftment and intrahepatic distribution<br />

of transplanted hepatocytes isolated from ETN+NG preconditioned<br />

livers was similar to cells from healthy control rats.<br />

Conclusions: The donor liver was successfully preconditioned<br />

and restored to normal health with a simple regimen of NO-donor<br />

and TNF-α blockade. This drug approach will be readily<br />

translated to clinical setting for improving outcomes in liver<br />

transplantation and use of donor livers for cell therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Thirunavukkarasu Chinnasamy,<br />

Fadi L. Jaber, Sanjeev Gupta<br />

341<br />

Identification of microRNAs Associated with Allograft<br />

Tolerance<br />

Matthew J. Vitalone 1 , Liang Wei 2 , Karine Piard-Ruster 1 , Carlos<br />

O. Esquivel 1 , Olivia M. Martinez 1 , Sheri M. Krams 1 ; 1 Surgery,<br />

Stanford University School of Medicine, Stanford, CA; 2 Sichuan<br />

Academy of Medical Sciences and Sichuan Provincial People’s<br />

Hospital, Chengdu, China<br />

Introduction: Although the liver is less immunogenic than<br />

other solid organs, liver transplant recipients receive lifelong<br />

immunosuppression. The aim of this study was to profile the<br />

expression of microRNAs (miRNAs) associated with the induction<br />

and maintenance of tolerance to an allograft. Methods:<br />

Previous <strong>studies</strong> showed that donor-specific tolerance can be<br />

induced in recipients of rat orthotopic liver transplants (OLT)<br />

after post-transplant total lymphoid irradiation (TLI). To identify<br />

miRNAs associated with tolerance we profiled liver grafts from<br />

syngeneic (DAàDA) and allogeneic OLT recipients (DAàLewis)<br />

that received post-transplant TLI, allograft recipients that were<br />

not treated post-transplant, and normal DA livers. Untreated<br />

allograft recipients reject their grafts within 10 days whereas<br />

TLI treated recipients have long-term graft survival (>100 days),<br />

thus miRNAs were examined at two time points, seven days<br />

(induction of tolerance) and 100 days (established tolerance)<br />

post-transplant Results: A supervised principal component analysis<br />

(PCA) using the top 15 differentially altered miRNA was<br />

able to robustly model the study groups with as little as three<br />

principal components describing 88.5% of the variance of<br />

the data. Indeed, three miRNA, miR142-3p, miR142-5p and<br />

miR181a were determined to be key miRNA associated with<br />

tolerance. Further, analysis of the full complement of miRNAs,<br />

by PCA, revealed that both the established tolerance and the<br />

syngeneic groups occupied the same relative three-dimensional<br />

space close to the normal liver group, indicating similar miRNA<br />

profiles. Conclusions: A small group of miRNA can distinguish<br />

graft status post-transplant. Further, our findings support a<br />

model whereby tolerant liver allografts express a profile consistent<br />

with that of healthy livers.<br />

Disclosures:<br />

The following authors have nothing to disclose: Matthew J. Vitalone, Liang Wei,<br />

Karine Piard-Ruster, Carlos O. Esquivel, Olivia M. Martinez, Sheri M. Krams<br />

342<br />

Previous ischemic preconditioning in experimental<br />

model of steatotic liver trasplantation with brain death<br />

Mónica B. Jiménez-Castro 1 , Mariana Mendes-Braz 1 , Jordi<br />

Gracia-Sancho 3 , Maria Eugenia Cornide-Petronio 1 , Araní<br />

Casillas-Ramírez 2 , Juan Rodes 1 , Carmen Peralta 1 ; 1 Institut D’Investigacions<br />

Biomèdiques August Pi i Sunyer, Barcelona, Spain; 2 Hospital<br />

Regional de Alta Especialidad de Ciudad Victoria, Ciudad<br />

Victoria, Mexico; 3 Barcelona Hepatic Hemodynamic Laboratory,<br />

IDIBAPS, CIBEREHD, Institut D’Investigacions Biomèdiques August<br />

Pi i Sunyer, Barcelona, Spain<br />

Background & Aims: Liver transplantation has evolved to<br />

become a standard therapy for certain end-stage liver diseases.<br />

Nowadays, high percent of organs come from donors who<br />

have suffered brain trauma-brain dead donors, which may also<br />

show hepatic steatosis, being both characteristics risk factors<br />

in liver transplantation. Nevertheless brain dead reduces the<br />

tolerance of liver grafts to the preservation/reperfusion injury<br />

and reduces graft survival. Ischemic preconditioning shows<br />

benefits when applied in the liver transplantation from nonbrain<br />

dead patients like hepatectomies, whereas it has been<br />

less promising in the transplantation from brain dead patients.


384A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

This study examined how brain death affects preconditioned<br />

steatotic liver grafts undergoing transplantation. Methods: Steatotic<br />

grafts from non-brain dead and brain dead-donors were<br />

cold stored for 6 hours and then transplanted. After 4 hours of<br />

reperfusion, hepatic damage was analyzed. Two strategies,<br />

acetylcholine pre-treatment and ischemic preconditioning or<br />

the combination of them were evaluated and their underlying<br />

mechanisms were characterized. Results: The presence<br />

of brain dead exacerbated hepatic damage in steatotic liver<br />

since increased transaminase levels and damage score in<br />

comparison with the liver transplantation group. Acetylcholine<br />

treatment reduced hepatic damage after liver transplantation<br />

in brain dead donors, through PKC, increased antioxidants<br />

and reduced lipid peroxidation, nitrotyrosines and neutrophil<br />

accumulation,. On the other hand, preconditioning benefits in<br />

non-brain dead donors were associated with nitric oxide and<br />

acetylcholine generation. In brain dead donors, preconditioning<br />

generated nitric oxide but did not promote acetylcholine<br />

up-regulation, and this resulted in inflammation and damage.<br />

Conclusions: We suggest that the combination of acetylcholine<br />

treatment and preconditioning conferred stronger protection<br />

against damage, oxidative stress and neutrophil accumulation<br />

than acetylcholine treatment alone. These superior beneficial<br />

effects were due to a selective preconditioning-mediated<br />

generation of nitric oxide and regulation of PPAR and TLR4<br />

pathways, which were not observed when acetylcholine was<br />

administered alone. We herein propose that the time frame<br />

between the declaration of brain dead and organ retrieval<br />

provides an important window for cytoprotective intervention,<br />

which may counteract the detrimental effects of brain dead.<br />

Consequently, our findings show that the combination of acetylcholine<br />

and preconditioning as a feasible and a protective<br />

strategy to reduce the adverse effects of brain dead and to<br />

improve the quality of liver grafts.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mónica B. Jiménez-Castro, Mariana<br />

Mendes-Braz, Jordi Gracia-Sancho, Maria Eugenia Cornide-Petronio, Araní<br />

Casillas-Ramírez, Juan Rodes, Carmen Peralta<br />

343<br />

WITHDRAWN<br />

344<br />

Survival After Intra-Arterial Treatment for Hepatocellular<br />

Carcinoma: Impact of Liver Dysfunction and Development<br />

of a Predictive Statistical Model: an International<br />

Collaborative Project<br />

Imam Waked 1 , Sarah Berhane 2 , Stephen L. Chan 3 , Asmaa<br />

Gomaa 4 , Mercedes Iñarrairaegui 5 , Zahraa Alkhatib 4 , Takashi<br />

Kumada 6 , Paul B. Lai 7 , Tim Meyer 8 , Frankie Mo 3 , Daniel Palmer 2 ,<br />

Bruno Sangro 5 , Nicholas Stern 9 , Toshifumi Tada 6 , Hidenori<br />

Toyoda 6 , Arndt Vogel 10 , Winnie Yeo 3 , Phillip Johnson 2 ; 1 Hepatology,<br />

National Liver Institute, Menoufiya, Egypt; 2 Molecular<br />

and Clinical Cancer Medicine;, University of Liverpool, Liverpool,<br />

United Kingdom; 3 Sir Y. K. Pao Centre for Cancer, Chinese University<br />

of Hong Kong, Hong Kong, China; 4 Oncology, National<br />

Liver Institute, Shebeen El Kom, Egypt; 5 Centro de Investigacion<br />

Biomedica en Red de Enfermedades Hepaticas y Digestivas, Clinica<br />

Universidad de Navarra, Pamplona, Spain; 6 Ogaki Municipal<br />

Hospital, Ogaki, Japan; 7 Surgery, The Chinese University of Hong<br />

Kong, Hong Komg, China; 8 Oncology, University College of London<br />

Cancer Institute, London, United Kingdom; 9 Gastroenterology,<br />

Aintree University Hospital, Liverpool, United Kingdom; 10 Hannover<br />

Medical School, Hannover, Germany<br />

Background: TransArterial Chemo-Embolisation (TACE) is recommended<br />

for patients with BCLC intermediate stage HCC<br />

(stage B) particularly in patients with excellent underlying liver<br />

function and minimal symptoms. A recently developed measure<br />

of liver function (the Albumin-Bilirubin (ALBI) grade) now permits<br />

detailed assessment of the impact of liver function on survival<br />

(1). Methods: We accrued patient level data from 2454<br />

patients undergoing TA(C)E – China (n= 242), Japan (n=655),<br />

Europe (UK, Germany and Spain combined, n=559 ) and<br />

Egypt (n=998). Overall, 59 % had Child-Pugh grade ‘A’ and<br />

17%, portal vein invasion. Liver function was graded according<br />

to their ALBI grade. A statistical model for survival after<br />

TA(C)E was built on a randomly selected half of a combined<br />

group of 1214 Japanese and European patients, n=602 (323<br />

Japanese and 279 European) and then validated the on the<br />

remaining 612 patients (332 Japanese and 280 European).<br />

The TACE model was also validated on an independent cohort<br />

from Egypt (n=998). Results: Our TACE model accurately predicted<br />

survival in the second half validation set (actual median<br />

survival = 20.4 months, predicted = 20.7 months) as well as<br />

in the Egyptian cohort (actual median survival = 18 months,<br />

predicted = 20 months). One year survival was 65.7 and<br />

71.3% for actual and predicted respectively). Classification of<br />

patients according their ALBI grade resulted in clear, non-overlapping<br />

survival curves in all cohorts (figure) Conclusion: ALBI<br />

categorised patients receiving TACE into three clear prognostic<br />

groups. Our TACE model accurately predicted survival on the<br />

basis of baseline clinical and laboratory features. References<br />

1. Johnson PJ et al., J Clin Oncol. JCO.2014.57.9151


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 385A<br />

Multivariate Cox proportional hazards model analysis for risk of<br />

overall mortality<br />

Disclosures:<br />

The following authors have nothing to disclose: Teng-Yu Lee, Yi-Ju Chen, Jaw-<br />

Town Lin, Hsiu-Jon Ho, Ming-Shiang Wu, Chun-Ying Wu<br />

Disclosures:<br />

Imam Waked - Advisory Committees or Review Panels: Janssen; Speaking and<br />

Teaching: Hoffman L Roche, Merck, BMS, Gilead, AbbVie<br />

Bruno Sangro - Speaking and Teaching: Sirtex Medical Europe, Bayer Schering<br />

Pharma<br />

Nicholas Stern - Speaking and Teaching: Bayer Pharmaceuticals<br />

The following authors have nothing to disclose: Sarah Berhane, Stephen L. Chan,<br />

Asmaa Gomaa, Mercedes Iñarrairaegui, Zahraa Alkhatib, Takashi Kumada,<br />

Paul B. Lai, Tim Meyer, Frankie Mo, Daniel Palmer, Toshifumi Tada, Hidenori<br />

Toyoda, Arndt Vogel, Winnie Yeo, Phillip Johnson<br />

345<br />

Nucleos(t)ide analogue therapy in relation to survival<br />

after transarterial chemoembolization for HBV-related<br />

hepatocellular carcinoma<br />

Teng-Yu Lee 1 , Yi-Ju Chen 1 , Jaw-Town Lin 2 , Hsiu-Jon Ho 2 , Ming-Shiang<br />

Wu 3 , Chun-Ying Wu 1 ; 1 Taichung Veterans General Hospital,<br />

Taichung, Taiwan; 2 Fu Jen Catholic University, New Taipei City,<br />

Taiwan; 3 National Taiwan University Hospital, Taipei, Taiwan<br />

Transarterial chemoembolization (TACE) is the most widely<br />

used primary treatment for unresectable hepatocellular carcinoma<br />

(HCC), but patient survival is unsatisfactory. We aimed<br />

to investigate the association between nucleos(t)ide analogue<br />

(NA) therapy and patient survival in hepatitis B virus (HBV)-related<br />

HCC following TACE. Using the Taiwan National Health<br />

Insurance Research Database between October 1, 2003 and<br />

December 31, 2011, we screened 57,566 patients with newly<br />

diagnosed HCC. Excluding patients with advanced HCC, we<br />

identified 5,182 patients who received TACE for HBV-related<br />

HCC (869 used NAs 90 days and 4,313 never used NA<br />

after TACE). Patients in the NA-treated cohort were randomly<br />

matched 1:4 with patients in the untreated cohort by age, sex,<br />

cirrhosis, coexisting diseases and the time period between<br />

TACE and initiation of NA therapy. Finally, 582 patients were<br />

recruited in the NA-treated group and 2,328 in the untreated<br />

group. Cumulative incidences of and hazard ratios (HRs) for<br />

patient mortality were analyzed. The mortality rates of the<br />

NA-treated group were significantly lower than those of the<br />

untreated group (1-year: 43.5%, 95% confidence interval [CI]:<br />

39.3-47.7% vs. 56.8%, 95%CI: 54.7-58.9%; 2-year: 58.4%,<br />

95%CI: 53.9-62.9% vs. 74.5%, 95%CI: 52.3-76.6%; 5-year:<br />

78.4%, 95%CI: 73.2-83.6% vs. 88.6%, 95%CI: 86.3-90.8%;<br />

P< 0.001). In multivariate regression analysis, NA therapy was<br />

independently associated with a decreased mortality risk (HR<br />

0.66, 95%CI: 0.59-0.75; P< 0.001). Multivariate stratified<br />

analyses verified the association of NA therapy and decreased<br />

mortality in all patient subgroups. Conclusion: NA therapy was<br />

associated with a decreased mortality among patients with<br />

HBV-related HCC following TACE.<br />

346<br />

Risk assessment of hepatocellular carcinoma in chronic<br />

hepatitis B patients with long-term antiviral therapy<br />

Kyu sik Chung, Do Young Kim, Beom Kyung Kim, Seung Up Kim,<br />

Jun Yong Park, Hye Jin Ku, Kwang-Hyub Han, Sang Hoon Ahn;<br />

Department of Internal Medicine, Yonsei University College of<br />

Medicine, Seoul, Korea (the Republic of)<br />

Background: In the era of potent antiviral therapy, the prognostic<br />

significance of a biological gradient of serum HBV-DNA levels<br />

measured at given time points has substantially diminished,<br />

another predictive factors, determining the long-term prognosis<br />

such as hepatocellular carcinoma (HCC) development, are<br />

required Methods: The aim of this study was to evaluate the<br />

incidence and risk factor of HCC occurrence in a large group<br />

of treatment-naïve chronic hepatitis B (CHB) patients, treated<br />

with entecavir (ETV) or tenofovir (TDF). Additionally, predictive<br />

accuracy of published HBV-HCC risk scores (CU-HCC<br />

score, GAG-HCC score, REACH-B score and LSM-HCC score)<br />

were evaluated. Results: A total of 762 CHB patients, who<br />

were treated with ETV (n=546) or TDF (n=216), were consecutively<br />

enrolled in this study. The median age (486 males and<br />

276females) was 51 years and 261 (34.3%) patients have<br />

liver cirrhosis. During the median follow-up of 45.8 months,<br />

HCC developed in 39 patients (5.1%) and the 1-, 3-, 5-, and<br />

7-year cumulative incidences of HCC were 1.6%, 4.0%, 7.2%<br />

and 9.5%, respectively. In multivariate analysis, old age (hazard<br />

ratio [HR] 1.091; 95% confidence interval [CI], 1.040-<br />

1.144;P0.05). In terms of HCC<br />

development at 3-/5-years, the AUROC of CU-HCC score were<br />

0.851/0.846, LSM-HCC score, 0.772/0.763, GAG-HCC<br />

score, 0.719/0.739, and REACH-B score, 0.710/0.691, indicating<br />

that liver fibrotic burden need to be incorporated properly<br />

into HBV-HCC prediction models to improve predictive<br />

performance. Conclusions: In the era of antiviral therapy, individual<br />

risks of HBV-HCC should be assessed effectively based<br />

on age and the fibrotic burden, not serum HBV-DNA level or<br />

virological response to antiviral therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kyu sik Chung, Do Young Kim,<br />

Beom Kyung Kim, Seung Up Kim, Jun Yong Park, Hye Jin Ku, Kwang-Hyub Han,<br />

Sang Hoon Ahn


386A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

347<br />

Mutations of p53, ARID1A and KRAS mutations and<br />

their association with “subtypes with stem-cell feature”<br />

in combined hepatocellular-cholangiocellular carcinoma<br />

Motoko Sasaki 1 , Yasunori Sato 1 , Yasuni Nakanuma 2,1 ; 1 Human<br />

Pathology, Kanazawa University Graduate School of Medicine,<br />

Kanazawa, Japan; 2 Shizuoka Cancer Center, Shizuoka, Japan<br />

Backgrounds/Aims. Combined hepatocellular-cholangiocarcinoma<br />

(cHC-CC) is composed of hepatocellular carcinoma<br />

(HCC), cholangiocarcinoma (CC) and diverse components with<br />

intermediate features between HCC and CC and is characterized<br />

by poor prognosis. Subtypes with stem cell features (SC<br />

subtype) including typical subtype (TS), intermediate cell subtype<br />

(INT) and cholangiolocellular type (CLC) were proposed<br />

in the WHO classification 2010. We have previously reported<br />

that each SC subtype may have different clinicopathological<br />

significance in cHC-CC. That is, the proportion of INT was significantly<br />

correlated to histological grading of coexistent HCC<br />

and tumor size, whereas the proportion of CLC was inversely<br />

correlated to histological grading of coexistent HCC and<br />

tumor size. In this study, we examined mutation status of p53,<br />

ARID1A and KRAS and their association with clinicopathological<br />

features, especially with SC subtypes, in cHC-CC. Methods.<br />

Fifty-three patients with cHC-CC (14 women and 39 men, age<br />

ranged 36-83 yrs, mean±SD, 65± 9.5yrs) were retrieved from<br />

our pathological files (1996-2014). The background diseases<br />

were hepatitis B (n=9), hepatitis C (21), chronic alcoholism or<br />

nonalcoholic fatty liver disease (NAFLD) (8) and cryptogenic<br />

(15). Mutations of p53, ARID1A and KRAS were also evaluated<br />

using immunohistochemistry (p53, ARID1A) and direct<br />

sequencing (KRAS). The prevalence of each component (HCC,<br />

TS, INT, CLC and CC) was histologically assessed with assistance<br />

of mucin staining and immunohistochemical staining<br />

for CK7, CK19, EMA, EpCAM, NCAM, AFP and HepPar1.<br />

Results: Mutations of p53, ARID1A and KRAS were detected<br />

in 24 cases (45%), 7 (13%) and 4 of 39 (10%) respectively.<br />

Mutations of both p53 and ARID1A were detected in 3 and<br />

mutations of both p53 and KRAS were detected in 2 cHC-<br />

CCs. SC subtypes were observed in all cHC-CCs in various<br />

amount and combination. The prevalence of each SC subtype<br />

in cHC-CC was as follows; TS, 8 (15%); INT, 33 (62%); and<br />

CLC, 35 (66%). ARID1A-mutated cHC-CCs were significantly<br />

smaller size (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 387A<br />

The following authors have nothing to disclose: Suraj Sharma, Matthew Kowgier,<br />

Bettina E. Hansen, Korosh Khalili, Willem Pieter Brouwer, Gideon Hirschfield<br />

The following authors have nothing to disclose: Yee-Kit Tse, Terry C. Yip<br />

349<br />

Oral nucleos(t)ide analogues (NAs) improve survival<br />

of patients with hepatocellular carcinoma after tumor<br />

resection but not after other treatment modalities – a<br />

study of 2,198 patients with chronic hepatitis B<br />

Grace LH Wong, Yee-Kit Tse, Vincent W. Wong, Terry C. Yip,<br />

Henry Lik-Yuen Chan; Institute of Digestive Disease, The Chinese<br />

University of Hong Kong, Shatin, Hong Kong<br />

Background: Oral nucleos(t)ide analogues (NAs) effectively<br />

suppress hepatitis B virus (HBV) and reduce tumor recurrence<br />

and death in patients with HBV-related hepatocellular carcinoma<br />

(HCC). We aimed to investigate the effect of NAs on<br />

the clinical outcomes after different HCC treatments (tumor<br />

resection, local ablative therapy [LAT], transartierial chemoembolization<br />

[TACE]) for HCC in these patients. Methods: We<br />

conducted a territory-wide cohort study using the database<br />

from Hospital Authority, which provides medical services at<br />

both in-patient and out-patient settings for 70-80% of the Hong<br />

Kong citizens. We identified chronic hepatitis B (CHB) patients<br />

with HCC by International Classification of Diseases, Ninth<br />

Revision, Clinical Modification (ICD-9-CM) diagnosis codes,<br />

diagnosed between 2000 and 2012. HCC treatments, NA<br />

use and laboratory parameters were retrieved and studied.<br />

The primary outcome was HCC recurrence and death. A<br />

3-month landmark analysis was used to evaluate the relative<br />

risk of primary outcome in patients with or without NA treatment.<br />

Cox proportional hazards models were used to estimate<br />

adjusted HRs with 95% CIs of HCC recurrence (taking death<br />

into account as a competing risk) and mortality associated<br />

with post-treatment use of NA. Results: 2,198 CHB patients<br />

(1,230 NA-untreated and 968 NA-treated) with HCC received<br />

at least one type of HCC treatment were included in the analysis.<br />

Tumor resection, LAT, TACE, resection-LAT and resection/<br />

LAT-TACE were the HCC treatments in 810 (36.9%), 1,015<br />

(46.2%), 222 (10.1%), 73 (3.3%) and 78 (3.5%) patients.<br />

At a median follow-up of 2.8 years (IQR 1.4 – 4.9 years)”,<br />

tumor recurrence and death occurred in 451 (36.7%) and 578<br />

(47.0%) of NA-untreated subjects; and 216 (22.3%) and 301<br />

(31.1%) of NA-treated subjects. NA therapy reduced the risk of<br />

overall HCC recurrence (adjusted sub-hazard ratio [SHR] 0.62,<br />

95% confidence interval [CI] 0.49–0.80; P 60<br />

years old). Results: Surveillance starts to be effective above 45<br />

years of age (gaining >100 quality-adjusted life days) though<br />

it is still relatively cost-ineffective with an incremental cost-effectiveness<br />

ratio (ICER) of $88,160/QALY. As the starting age<br />

of surveillance increases, the ICER for surveillance decreases.<br />

Beyond the age of 58, the ICER for surveillance is consistently<br />

lower than the standard willingness-to-pay threshold of<br />

$50,000/QALY. Conclusions: Due to the marked decrease in<br />

HCC incidence, universal screening in patients with cirrhosis<br />

post-SVR is very unlikely to be cost-effective; however, by targeting<br />

surveillance to those over 55 years of age (figure), HCC<br />

surveillance would represent better value-for-money. Data to<br />

identify high-risk populations may allow for further tailoring of<br />

recommendations according to a patient’s risk profile.


388A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Adriaan J. van der Meer - Consulting: Gilead; Speaking and Teaching: MSD,<br />

Gilead<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Jordan J. Feld - Advisory Committees or Review Panels: Merck, Janssen, Gilead,<br />

AbbVie, Theravance, Bristol Meiers Squibb; Grant/Research Support: AbbVie,<br />

Boehringer Ingelheim, Janssen, Gilead, Merck<br />

The following authors have nothing to disclose: Hooman Farhang Zangneh, William<br />

W. Wong, Beate Sander, Chaim M. Bell, Khalid Mumtaz, Matt Kowgier,<br />

Sean P. Cleary, Kelvin K. Chan<br />

351<br />

Whole blood coagulation tests are not equally able to<br />

detect haemostatic prothrombotic alterations in patients<br />

with liver cirrhosis and hepatocellular carcinoma (HCC):<br />

rotational thromboelastometry versus thrombingeneration<br />

test.<br />

Alberto Zanetto 1 , Alberto Ferrarese 1 , Alessandro Vitale 2 , Umberto<br />

Cillo 2 , Mariangela Fadin 3 , Sabrina Gavasso 3 , Claudia M. Radu 3 ,<br />

Luca Spiezia 3 , Fabio Farinati 1 , Patrizia Burra 1 , Paolo Simioni 3 ,<br />

Marco Senzolo 1 ; 1 Surgery, Oncology and Gastroenterology,<br />

University of Padua, Gastroenterology, Padova, Italy; 2 Surgery,<br />

Oncology and Gastroenterology, University of Padua, Liver Transplantation<br />

Unit, Padova, Italy; 3 Department of Medicine, Padua<br />

University Hospital, V Chair of Internal Medicine, Padova, Italy<br />

Background and aim: neoplasms are considered as one of the<br />

main important cause of acquired thrombophilia but <strong>studies</strong><br />

exploring coagulation imbalance induced by HCC in liver cirrhosis<br />

are lacking. The aim of this study study was to evaluate the<br />

thrombophilic role of HCC in patients with cirrhosis. Methods:<br />

cirrhotic patients with and without HCC were enrolled in the<br />

study and underwent rotational thromboelastometry (ROTEM),<br />

platelet count, determination of levels of pro and anticoagulation<br />

factors and thrombin generation test (TG), with and without<br />

trombomodulin (TM). Results: 76 cirrhotics, of whom 41 with<br />

HCC, were included. Volume of active HCC was >5 cm 3 in<br />

18 patients. Levels of pro and anticoagulation factors were<br />

similar in patients with and without HCC, but fibrinogen was<br />

increased in HCC patients with tumor volume >5cm 3 compared<br />

to those with 5cm 3 compared to those with


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 389A<br />

that mediates doxorubicin sensitivity through a process that<br />

involves its acetylation and S574 phosphorylation. SIRT6 can<br />

deacetylate FOXO3 preventing apoptosis, and TACE-resistant<br />

tumors have higher SIRT6 levels. Manipulation of FOXO3 modifications<br />

may thus prove useful in enhancing the chemotherapy<br />

sensitivity of HCC.<br />

Disclosures:<br />

Steven A. Weinman - Consulting: Cardax, Inc.<br />

The following authors have nothing to disclose: Josiah Cox, Zhuan Li, Anusha<br />

Vittal, Maura O’Neil, Brian Bridges, Sean Kumer, Timothy Schmitt<br />

353<br />

Long non-coding RNA integrator complex subunit 6<br />

pseudogene 1 (INTS6P1), a tumor suppressor and prognostic<br />

factor in hepatocellular carcinoma, induces apoptosis<br />

via intrinsic mitochondrial pathway<br />

Minqiang Lu 1 , Ka Yin LUI 1 , Haoran Peng 1,2 , Rongdang Fu 1 , Huan<br />

Chen 1 , Chunhui Qiu 1 ; 1 Hepatic Surgery, The Third Affiliated Hospital<br />

of Sun Yat-sen University, Guangzhou, China; 2 Division of<br />

Gastroenterology and Hepatology, The Johns Hopkins Hospital,<br />

Baltimore, MD<br />

Purpose: Long non-coding RNAs (lncRNAs) have been proved<br />

its close association with many diseases especially with tumor.<br />

We aimed to determine whether the lncRNA integrator complex<br />

subunit 6 pseudogene 1 (INTS6P1) could be as a tumor<br />

suppressor and its mechanism in hepatocellular carcinoma<br />

(HCC). Materials and methods: Firstly, the expression level of<br />

INTS6P1 mRNA were measured in a cohort of 60 HCC tissues<br />

and adjacent normal liver tissues by using the quantitative<br />

real time-polymerase chain reaction (qRT-PCR); Secondly, the<br />

functional <strong>studies</strong> of INTS6P1, include growth curves, migration<br />

assays and cell death, were detected in HCC cell lines.<br />

Finally, the mechanism experiment was investigated for tumor<br />

suppressive roles of INTS6P1. Results: We found that lncRNA<br />

INTS6P1 was remarkably down-regulated in HCC (71.4%),<br />

and its expression was significantly correlated with pathology<br />

grade (p


390A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

355<br />

Prognostic assessment of patients with intermediate<br />

stage, untreated hepatocellular carcinoma<br />

Edoardo G. Giannini 1 , Alessandro Moscatelli 1 , Gaia Pellegatta 1 ,<br />

Fabio Farinati 2 , Francesca Ciccarese 3 , Fabio Piscaglia 4 , Gian<br />

Ludovico Rapaccini 5 , Mariella Di Marco 6 , Eugenio Caturelli 7 ,<br />

Marco Zoli 8 , Franco Borzio 9 , Giuseppe Cabibbo 10 , Martina M.<br />

Felder 11 , Rodolfo Sacco 12 , Filomena Morisco 13 , Gabriele Missale<br />

14 , Francesco G. Foschi 15 , Antonio Gasbarrini 16 , Gianluca<br />

Svegliati Baroni 17 , Roberto Virdone 18 , Maria Chiaramonte 19 ,<br />

Franco Trevisani 20 ; 1 Dipartimento di Medicina Interna, Unità di<br />

Gastroenterologia, IRCCS-Azienda Ospedaliera Universitaria San<br />

Martino-IST, Università di Genova, Genova, Italy; 2 Dipartimento di<br />

Scienze Chirurgiche e Gastroenterologiche, Unità di Gastroenterologia,<br />

Università di Padova, Padova, Italy; 3 Divisione di Chirurgia,<br />

Policlinico San Marco, Zingonia, Italy; 4 Dipartimento di Scienze<br />

Mediche e Chirurgiche, Unità di Medicina, Alma Mater Studiorum<br />

– Università di Bologna, Bologna, Italy; 5 Unità di Medicina Interna<br />

e Gastroenterologia, Complesso Integrato Columbus, Università<br />

Cattolica di Roma, Roma, Italy; 6 Divisione di Medicina, Azienda<br />

Ospedaliera Bolognini, Seriate, Italy; 7 Unità Operativa di Gastroenterologia,<br />

Ospedale Belcolle, Viterbo, Italy; 8 Dipartimento<br />

di Scienze Mediche e Chirurgiche, Unità di Medicina Interna,<br />

Alma Mater Studiorum – Università di Bologna, Bologna, Italy;<br />

9 Dipartimento di Medicina, Unità di Radiologia, Ospedale Fatebenefratelli,<br />

Milano, Italy; 10 Dipartimento Biomedico di Medicina<br />

Interna e Specialistica, Unità di Medicina Gastroenterologia,<br />

Università di Palermo, Palermo, Italy; 11 Ospedale Regionale<br />

di Bolzano, Unità di Gastroenterologia, Bolzano, Italy; 12 Unità<br />

Operativa Gastroenterologia e Malattie del Ricambio, Azienda<br />

Ospedaliero-Universitaria Pisana, Pisa, Italy; 13 Dipartimento di<br />

Medicina Clinica e Chirurgia, Unità di Gastroenterologia, Università<br />

di Napoli “Federico II”, Napoli, Italy; 14 Unità di Malattie Infettive<br />

ed Epatologia, Azienda Ospedaliero-Universitaria di Parma,<br />

Parma, Italy; 15 Dipartimento di Medicina Interna, Ospedale per<br />

gli Infermi di Faenza, Faenza, Italy; 16 Unità di Medicina Interna<br />

e Gastroenterologia, Policlinico Gemelli, Università Cattolica di<br />

Roma, Roma, Italy; 17 Clinica di Gastroenterologia, Università<br />

Politecnica delle Marche, Ancona, Italy; 18 Dipartimento Biomedico<br />

di Medicina Interna e Specialistica, Unità di Medicina Interna<br />

2, Azienda Ospedaliera Ospedali Riuniti Villa Sofia-Cervello, Palermo,<br />

Italy; 19 Unità di Gastroenterologia, Ospedale Sacro Cuore<br />

Don Calabria, Negrar, Italy; 20 Dipartimento di Scienze Mediche<br />

Chirurgiche, Unità di Semeiotica Medica, Alma Mater Studiorum<br />

– Università di Bologna, Bologna, Italy<br />

Background: The Barcelona Clinic Liver Cancer (BCLC) intermediate<br />

stage (BCLC B) includes a heterogeneous population<br />

of patients with hepatocellular carcinoma (HCC), who often<br />

display different survival rates. Due to these reasons, a panel<br />

of experts recently proposed a sub-classification of the intermediate<br />

stage HCC in order to facilitate prognostic assessment<br />

and treatment decisions in clinical practice. Aim: In this study<br />

our aim was to assess the prognostic capability of the re-classification<br />

of the intermediate stage HCC (BCLC B) in a large<br />

cohort of patients with untreated HCC managed by the Italian<br />

Liver Cancer (ITA.LI.CA) Group. Methods: We assessed the<br />

prognosis of 269 untreated patients with HCC observed in the<br />

period 1987-2012, and who were classified according to the<br />

proposed sub-classification of the BCLC B stage from stage B1<br />

to stage B4 (Semin Liver Dis 2012;32:348-59). Survival was<br />

defined as the time, expressed in months, that elapsed from<br />

the date of HCC diagnosis and the date of death, or of the<br />

most recent follow-up information. Results: According to the<br />

proposed sub-classification of the intermediate stage HCC, 65<br />

patients were classified B1 (24.2%), 105 patients B2 (39.0%),<br />

22 patients B3 (8.2%), and 77 patients B4 (28.6%). Overall<br />

median survival was 13 months, and progressively decreased<br />

from stage B1 (25 months) through stages B2 (16 months)<br />

and B3 (9 months), to stage B4 (5 months, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 391A<br />

B3, C), alpha-fetoprotein, performance status and Child–Pugh<br />

score in predicting survival and guiding treatment allocation.<br />

An ITA.LI.CA prognostic score and treatment scheme were then<br />

constructed by integration of clinical judgments, and then compared<br />

with other systems (BCLC, HKLC, MESIAH, CLIP, JIS).<br />

Findings. Median survival was well stratified by the main ITA.<br />

LI.CA stages (p


392A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(r=0.68). exRNA aligned read counts were lower in cancer<br />

than in normal bile samples. Read counts between cancer and<br />

normal bile exRNA samples had a wide range of correlation (r:<br />

0.64 to 0.95). A set of 14 mature (sense and antisense) miRNA<br />

were differentially expressed > 3SD fold change in expression<br />

(based on reads per million) in bile from CCA compared with<br />

normal bile. Other differentially expressed exRNA identified<br />

included tRNA and piRNA. Summary and Conclusions: Biliary<br />

exRNA sequencing provides a novel strategy for biomarker<br />

discovery. exRNA isolation protocols have a major impact on<br />

yield and quality of exRNA for downstream gene expression<br />

<strong>studies</strong>. Differentially expressed candidate exRNA biomarkers<br />

have been identified that may be useful for the diagnosis of<br />

CCA.<br />

Disclosures:<br />

The following authors have nothing to disclose: Irene K. Yan, Waseem David,<br />

Swathi Mohankumar, Sarah Nix, Hiroaki Haga, Yan W. Asmann, Tushar Patel<br />

359<br />

Risk Differences of Hepatocellular Carcinoma (HCC)<br />

Among Patients with Chronic Hepatitis C (CHC) of Different<br />

Viral Genotypes: a Historical Cohort Study in the<br />

United States<br />

Mingjuan Jin 2,1 , Changqing Zhao 1,7 , Joseph K. Hoang 1 , Nghia<br />

H. Nguyen 1,3 , An K. Le 1 , Christine Y. Chang 1 , Richard H. Le 1 ,<br />

Alina Kutsenko 4 , Lee Ann Yasukawa 5 , Jian Q. Zhang 6 , Susan C.<br />

Weber 5 , Mindie H. Nguyen 1 ; 1 Division of Gastroenterology and<br />

Hepatology, Stanford University Medical Center, Palo Alto, CA;<br />

2 Department of Epidemiology and Biostatistics, School of Public<br />

Health, Zhejiang University, Hangzhou, China; 3 Department of<br />

Medicine, University of California, San Diego, San Diego, CA;<br />

4 Department of Medicine, Stanford University Medical Center,<br />

Palo Alto, CA; 5 Center for Clinical Informatics, Stanford University<br />

School of Medicine, Palo Alto, CA; 6 Chinese Hospital, San Francisco,<br />

CA; 7 Department of Cirrhosis, Institute of Liver Disease, Shuguang<br />

Hospital, Shanghai University of T.C.M., Shanghai, China<br />

Purpose: Whether there exists differences in the risk for HCC<br />

among patients with CHC of different genotypes is still controversial,<br />

which were explored in the present study in a historical<br />

cohort from 2 ethnically diverse US centers. Methods:<br />

The subject inclusion was as follows: CHC patients aged 20<br />

years or older, presented at study centers between 1/1/1999<br />

and 12/31/2013, followed up for at least 1 year, having<br />

HCV genotype information, and no HCC at baseline. The HCC<br />

cumulative incidence and incidence density were calculated.<br />

Cox proportional hazards models were used to testify the associations<br />

of HCV genotypes with HCC. Results: A total of 1990<br />

eligible CHC patients were included, among whom 1174<br />

(59.0%) were female and 1594 (80.1%) were ≤ 60 years old<br />

with mean age of 53.20±9.63 years. The ethnicity distribution<br />

was 7.9% for Asian, 40.6% for White, 18.3% for Hispanc and<br />

Black, and 33.2% for others and unknown. The frequencies of<br />

genotypes 1a/untyped, 1b, 2, 3, and others (genotype 4 and<br />

6) were 51.4%, 18.6%, 13.0%, 13.1%, and 3.9%, respectively.<br />

Overall, the 1-year, 3-year and 5-year HCC cumulative<br />

incidence were 448.53, 1341.88, and 2354.31 per 10,000<br />

persons, respectively, with the incidence density of 297.37<br />

per 10, 000 person-years in this cohort. Compared to patients<br />

with the most prevalent genotype 1a/untyped, the age-, gender-<br />

and ethnicity-adjusted hazard ratios (95%CIs) were 0.88<br />

(0.64-1.22), 0.58 (0.37-0.93), 1.42 (0.99-2.04), and 1.89<br />

(1.14-3.14) for those with genotype 1b, 2, 3, and 4 and 6,<br />

respectively. Conclusions: The results suggest that HCV genotype<br />

4 and 6 are associated with the highest risk of HCC, and<br />

genotype 2 has the lowest risk, and there were no significant<br />

differences between genotype 1b and 1a/untyped . Preventive<br />

measures such as antiviral therapy and HCC surveillance<br />

efforts should especially target the higher risk groups.<br />

Disclosures:<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Mingjuan Jin, Changqing Zhao,<br />

Joseph K. Hoang, Nghia H. Nguyen, An K. Le, Christine Y. Chang, Richard H.<br />

Le, Alina Kutsenko, Lee Ann Yasukawa, Jian Q. Zhang, Susan C. Weber<br />

360<br />

Surveillance for Hepatocellular Carcinoma (HCC) Using<br />

Ultrasonography Alone is Associated with Improved<br />

Survival: Results of Inverse Probability of Treatment<br />

Weighting (IPTW)-Based Survival Analysis<br />

Piyanat Chattieng 2 , Jonggi Choi 1 , Nopavut Geratikornsupuk 2 ,<br />

Kessarin Thanapirom 2 , Sombat Treeprasertsuk 2 , Piyawat Komolmit<br />

2 , Roongruedee Chaiteerakij 2,1 ; 1 Mayo Clinic, Rochester, MN;<br />

2 Medicine, Chulalongkorn University, Bangkok, Thailand<br />

Background/Aim: The current American Association for the<br />

Study of Liver Disease guideline recommend HCC surveillance<br />

program using ultrasound modality alone. The impact of the surveillance<br />

program on survival remains unclear due to unavoidable<br />

bias and confounders in observational <strong>studies</strong>. We aimed<br />

to determine the effect of HCC surveillance on survival using<br />

an IPTW analysis controlling for inherent bias and confounders<br />

existing in observational <strong>studies</strong>. Methods: We conducted a<br />

retrospective cohort study of 446 patients with cirrhosis and/<br />

or chronic hepatitis B infection who were diagnosed with HCC<br />

between 2007 and 2013 in a major referral center in Bangkok,<br />

Thailand. Surveillance was defined as having repeated<br />

abdominal ultrasonography within 1 year prior to HCC diagnosis.<br />

Survivals of patients with and without surveillance were<br />

estimated using the Kaplan-Meier method with lead-time bias<br />

adjustment and compared using the log-rank test. Hazard ratio<br />

(HR) and 95% confidence interval (CI) of surveillance was<br />

computed using conventional Cox and weighted Cox proportional<br />

hazards analysis with IPTW adjustment. Results: Of the<br />

446 HCC patients, 79% were male with a mean age of 58.1<br />

years. Hepatitis B (54%), hepatitis C (20%), alcohol (11%) and<br />

non-alcoholic fatty liver disease (9%) were the most common<br />

underlying chronic liver diseases. A hundred and three (23%)<br />

were diagnosed with HCC through surveillance. The surveillance<br />

group had a higher proportion of patients diagnosed<br />

with the Barcelona-Clinic Liver Cancer (BCLC) stage A (82% vs.<br />

34%, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 393A<br />

P=0.005. The estimated effect of surveillance remained similar<br />

using IPTW-adjusted Cox analysis, with HR (95% CI) of<br />

0.57 (0.43-0.76), P


394A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

363<br />

Prediction of Recurrence after Radiofrequency Ablation<br />

of Hepatocellular Carcinoma using Liver Stiffness Measurement<br />

(FibroScan)<br />

Yu Rim Lee 1 , Soo Young Park 1 , Se Young Jang 1 , Su Hyun Lee 1 ,<br />

Sun Kyung Jang 1 , Won Young Tak 1 , Young Oh Kweon 1 , Jung Gil<br />

Park 2 , Seung Up Kim 3 ; 1 Kyunopook National University Hospital,<br />

Daegu, Korea (the Republic of); 2 Internal Medicine, Gastroenterology<br />

and Hepatology, Gumi, Korea (the Republic of); 3 Internal<br />

Medicine, Gastroenterology, Seoul, Korea (the Republic of)<br />

Background. Hepatocellular carcinoma(HCC) is a third leading<br />

cause of cancer related death and incidence is increasing<br />

every year. As active surveillance programs develop, detection<br />

of early stage HCC are increasing. RFA as well as hepatic<br />

resection and liver transplantation have been considered<br />

standard treatments of HCC. However, long-term prognosis<br />

is unsatisfactory due to HCC recurrence associated with liver<br />

fibrosis. The purpose of this study is to predict recurrence after<br />

radiofrequency ablation of hepatocellular carcinoma using<br />

liver stiffness measurement(LSM), which is known related to be<br />

effective in evaluating the liver fibrosis noninvasively. Methods.<br />

A total of 131 patients who performed LSM before RFA<br />

between November, 2008 and December, 2014(training set)<br />

and 55 between November, 2006 and December, 2012(validation<br />

set) was enrolled. Multiple clinical variables including<br />

LSM were analysed for association with recurrence. The result<br />

was applied to validation set. Results. Baseline characteristics<br />

including alfa-fetoprotein, INR, Child-Pugh Class(A/B), underlying<br />

liver disease(B/C/nonB,nonC/Alcohol) and recurrence<br />

rate of training set and validation set were different(p 12.3 kPa had shorter disease free survival<br />

in training set. In validation set, using a cutoff value of LSM<br />

12.3 kPa, recurrence rate were also significantly greater at<br />

LSM values > 12.3 kPa patients(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 395A<br />

365<br />

When to perform surgical resection or radiofrequency<br />

ablation for early hepatocellular carcinoma? A nomogram-guided<br />

treatment strategy<br />

Po-Hong Liu 1,2 , Chia-Yang Hsu 2,3 , Yi-Hsiang Huang 1,4 , Chien-<br />

Wei Su 1,2 , Han-Chieh Lin 1,2 , Teh-Ia Huo 1,5 ; 1 Department of Medicine,<br />

Taipei Veterans General Hospital, Taipei, Taiwan; 2 Faculty<br />

of Medicine, National Yang-Ming University, Taipei, Taiwan;<br />

3 Department of Internal Medicine, University of Nevada School<br />

of Medicine, Reno, NV; 4 Institute of Clinical Medicine, National<br />

Yang-Ming University, Taipei, Taiwan; 5 institute of Pharmacology,<br />

National Yang-Ming University, Taipei, Taiwan<br />

Background&Aims: Radiofrequency ablation (RFA) is indicated<br />

for early hepatocellular carcinoma (HCC). The comparative<br />

efficacy between RFA and surgical resection (SR) is inconclusive.<br />

We aimd to develop a prognostic nomogram for recurrence-free<br />

survival (RFS) after RFA and to evaluate its ability<br />

in improving treatment algorithm. Methods: We enrolled 836<br />

patients with BCLC very early/early HCC receiving SR or RFA.<br />

Nomogram was constructed with Cox proportional hazards<br />

model. RFA patients were stratified into low-risk and high-risk<br />

groups. The RFS and overall survival (OS) of two risk groups<br />

were compared with SR patients by propensity score matching<br />

analysis. Results: A prognostic nomogram was built based<br />

on number and size of tumor, platelet count, albumin level,<br />

and model for end-stage liver disease score with a c-index of<br />

0.69. SR provided better RFS and OS compared with high-risk<br />

(nomogram score ≥9.8) RFA patients in the propensity model.<br />

The 5-year RFS rates were 36% vs. 11%, while 5-year OS rates<br />

were 74% vs. 60% for SR and high-risk RFA group, respectively<br />

(both p


396A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

367<br />

Meta-Analysis: Proportions of Hepatocellular Carcinoma<br />

(HCC) Diagnosed by Screening, Surveillance, or Without<br />

Symptoms is Dismally Low Across Different World<br />

Regions and Underlying Liver Diseases<br />

Michelle Q. Jin 1 , Richard H. Le 1 , Mingjuan Jin 1,2 , Changqing<br />

Zhao 1,3 , Michael H. Le 4 , Vincent L. Chen 1 , Grace Wong 5 , Vincent<br />

W. Wong 5 , Young-Suk Lim 6 , Wan-Long Chuang 7 , Ming-Lung Yu 7 ,<br />

Mindie H. Nguyen 1 ; 1 Division of Gastroenterology and Hepatology,<br />

Stanford University Medical Center, Palo Alto, CA; 2 Department<br />

of Epidemiology and Biostatistics, School of Public Health,<br />

Zhejiang University, Hangzhou, China; 3 Department of Cirrhosis,<br />

Institute of Liver Disease, Shuguang Hospital, Shanghai University<br />

of T.C.M, Shanghai, China; 4 University of California, Santa Cruz,<br />

Santa Cruz, CA; 5 Division of Gastroenterology and Hepatology,<br />

Department of Medicine and Therapeutics, The Chinese University<br />

of Hong Kong, Hong Kong, China; 6 Department of Gastroenterology,<br />

Asan Liver Center, Asan Medical Center, University of Ulsan<br />

College of Medicine, Seoul, Korea (the Republic of); 7 Department<br />

of Internal Medicine, Kaohsiung Medical University Hospital,<br />

Kaohsiung Medical University, Kaohsiung, Taiwan<br />

Background: Guidelines for HCC screening and surveillance<br />

are available worldwide, but poor adherence to guidelines<br />

may result in symptomatic diagnosis with late-stage disease.<br />

Our goal is to estimate the proportions of HCC patients diagnosed<br />

via screening, surveillance or at asymptomatic stage<br />

(asymptomatic/screening HCC). Methods: PubMed search<br />

with MEDLINE filter for “HCC,” “Screening/Surveillance” and<br />

associated terms was performed in March 2015. Abstracts<br />

from recent international liver meetings were also reviewed.<br />

Two investigators identified original <strong>studies</strong> published in or<br />

after 2000 that reported proportions of asymptomatic/screening<br />

HCC with ≥100 HCC patients (except for subanalysis).<br />

We performed analysis using a random-effects model. Results:<br />

A total of 1356 <strong>studies</strong> were initially identified and full paper<br />

review was performed for 128 <strong>studies</strong>, yielding 28 eligible<br />

<strong>studies</strong> with 15,244 HCC patients. The overall pooled proportion<br />

of asymptomatic/screening HCC patients was 34%<br />

(95% CI 27-40%), and it was significantly lower in Asia-Pacific<br />

(23%, 95% CI 16-31%) compared to Europe (39%, 95% CI<br />

28-50%) (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 397A<br />

compared to those with score


398A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

needle-aspiration biopsy specimens of 44 HCC patients (20<br />

in the HBV group, 19 in the HCV group and 5 in the NASH<br />

group) were subjected to real-time PCR to measure expression<br />

levels of TAA mRNAs in tumor tissues. Results IFN-g ELISpot<br />

assay revealed that immune responses to the p53-derived epitope<br />

(positive rates: 20.7% in the HBV group, 5.0% in the HCV<br />

group and 0% in the NASH group), and to the MRP3-derived<br />

epitope (positive rates: 20.7% in the HBV group, 2.5% in the<br />

HCV group and 5.6% in the NASH group) were significantly<br />

induced in the HBV group than in other groups (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 399A<br />

median survival (MS) significantly improved from 1990-2011<br />

in localized, regional and distant metastases with CDS. In the<br />

65-75 age population MS improved only in CDS localized<br />

disease; 7 months (1990-2000) to 9.8 months (2001-2011). In<br />

the 75+ age population, the MS from 1990-2000 and 2001-<br />

2011 did not show significant change regardless of extent of<br />

disease; localized CDS patients was 6.6 months (1990-2000)<br />

compared to 6 months (2001-2011). Conclusion: Over the last<br />

two decades, there has been an increased incidence in patients<br />

receiving CDS likely due to advancements in hepatic resection.<br />

The overall survival in the elderly population with CDS (75+)<br />

though remains poor throughout the last two decades regardless<br />

of the extent of the disease. Advanced age seems to be a<br />

poor prognostic factor even with CDS, suggesting future study<br />

in this age population.<br />

Disclosures:<br />

The following authors have nothing to disclose: George Cholankeril, Ponnandai<br />

Somasundar<br />

373<br />

Patients with alcohol compared to HCV-related hepatocellular<br />

carcinoma (HCC) have a reduced survival.<br />

Results of a Nationwide study.<br />

Charlotte E. Costentin 1 , Nathalie GOUTTE 2 , Philippe Sogni 3 ,<br />

Bruno Falissard 4 , Noelle Bendersky 5 , Olivier Farges 2 ; 1 Hepatology,<br />

Hôpital Henri Mondor, Creteil, France; 2 Hepatology – Hepato-biliary<br />

surgery, Beaujon Hospital, Clichy, France; 3 Hepatology,<br />

Cochin Hospital, Paris, France; 4 biostatistique, Paris-Sud University,<br />

INSERM, Paris, France; 5 Medical Informatics, Beaujon Hospital,<br />

Clichy, France<br />

Introduction: Data on alcohol-related HCC are scarce. The<br />

aim of this study was to describe the incidence, management<br />

and prognosis of alcohol compared to HCV-related HCC, at a<br />

national level. Patients and Methods: French healthcare databases<br />

were screened to identify all adult patients who were<br />

hospitalized between 2007 and 2012 with a diagnosis of<br />

HCC. Incident cases of HCC were selected by defining the<br />

entry point as a first admission in 2009. Demographic data,<br />

associated conditions, underlying liver disease and etiology,<br />

as well as the type (teaching, general, private, charitable),<br />

location, and annual HCC-caseload of the hospitals where<br />

patients were first managed (1st tertile 45), were retrieved. Treatments were divided into<br />

curative (liver transplantation, resection, radiofrequency) or<br />

not. Survival of incident-cases was computed from the time of<br />

diagnosis and adjusted for potential confounding variables.<br />

In order to compare two homogeneous populations of alcohol<br />

and HCV-related HCC, patients with unknown etiology of<br />

liver disease, mixed etiologies or HBV were excluded, Results:<br />

Between 2009 and 2012, the study population included<br />

13,942 (83.4%) incident-cases of alcohol and 2,774 (16.6%)<br />

HCV-related HCC. The incidence of alcohol- and HCV-related<br />

HCC varied up to 2 fold from one mainland region to another,<br />

and in some regions, alcohol accounted for up to 96% of HCC.<br />

Alcohol- compared to HCV-related HCC occurred more often<br />

in men (89.2% vs 63.7% p


400A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ing program were achieved as all of the patients completed<br />

the follow-up till dead. Forty patients were identified as HCC<br />

during the 10 years follow-up. The totally 1-, 3-, 5-, and 10-year<br />

cumulative HCC incidence rate were 1.6% (n=6), 4.6%(n=18),<br />

7.1%(n=27) and 10.5% (n=40) respectively. Among 40 HCC<br />

patients, thirty-eight (95%) were identified as asymptomatic<br />

subclinical early HCC with tumor size less than 3.0cm, and<br />

without portal vein infiltration or metastasis. Thirteen patients<br />

had the chance of surgical operation, and 12 (92.3%) patients<br />

survived with no recurrence. Twenty-five patients received<br />

radiofrequency ablation or other palliative therapy because of<br />

liver function reserve limitation. Though totally sixteen patients<br />

died, only 3 (18.8%) patients died of HCC(two were late stage<br />

HCC with metastasis, one was liver failure after operation),<br />

other 13(81.2%) patients died of concomitant liver decompensation<br />

complications. The median survival time is 3 years after<br />

identification of HCC. The totally 1-, 3-, and 5-year survival<br />

rate of HCC patients were 97.5%, 91.2%, 67.7% respectively.<br />

Conclusion: This 10 years regular integrated follow-up experience<br />

indicates that HCC surveillance using ultrasonography<br />

at six month intervals really performs good with early HCC<br />

detection rate of 95%.<br />

Disclosures:<br />

Jidong Jia - Consulting: BMS, MSD, Roche<br />

The following authors have nothing to disclose: Xiaoning Wu, Xiaoming Wang,<br />

yameng sun, Yuanyuan Kong, Yu Wang, Xinyan Zhao, Qianyi Wang, Weijia<br />

Duan, Hong You, Xiaojuan Ou<br />

375<br />

Integrator complex subunit 6 (INTS6) is a prognostic<br />

marker for hepatocellular carcinoma<br />

Minqiang Lu 1 , Ka Yin LUI 1 , Rongdang Fu 1 , Haoran Peng 1,2 , Huan<br />

Chen 1 ; 1 Hepatic Surgery, The Third Affiliated Hospital of Sun Yatsen<br />

University, Guangzhou, China; 2 Division of Gastroenterology<br />

and Hepatology, The Johns Hopkins Hospital, Baltimore, MD<br />

Purpose: Integrator complex subunit 6 (INTS6), previously<br />

known as the gene encoding deleted in cancer cells 1 (DICE1)<br />

was found to play a tumor suppressive role in certain types of<br />

solid tumors just like prostate cancer and cervical cancer. However,<br />

the clinical characteristic of INTS6 hasn’t been reported<br />

by far. In this study, we wanted to investigate the expression of<br />

INTS6 in hepatocellular carcinoma (HCC) and evaluated the<br />

clinical characteristic and prognosis in HCC patients. Material<br />

and methods: (1) Firstly, the expression level of INTS6 mRNA<br />

were measured in a cohort of 50 HCC tissues and adjacent<br />

normal liver tissues by using the qRT-PCR; (2) Secondly, the<br />

expression level of INTS6 protein were detected in 20 paired<br />

HCC and corresponding adjacent normal tissue by using western<br />

blot; (3) At last, Immunohistochemistry was performed on<br />

70 archived paraffin-embedded HCC samples. Results: qRT-<br />

PCR and western blot analyses showed the INTS6 mRNA and<br />

protein expression was significantly down-regulated in tumor<br />

tissues compared to adjacent normal liver tissues (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 401A<br />

min ratio and NLR demonstrated a prognostic value in HCC<br />

patients treated with sorafenib in a western collective. Inflammation<br />

based scores have the potential of clinical utility to identify<br />

patients who are likely not to derive significant benefit from<br />

a systemic therapy. References J Budczies, F Klauschen, BV<br />

Sinn, B Gyrffy, WD Schmitt, S Darb-Esfahani, C Denkert. Cutoff<br />

Finder: a comprehensive and straightforward Web application<br />

enabling rapid biomarker cutoff optimization.. PLoS One 7,<br />

e51862 ().<br />

Disclosures:<br />

Martin F. Sprinzl - Grant/Research Support: GILEAD<br />

The following authors have nothing to disclose: Arndt J. Weinmann, Sandra<br />

Koch, Peter R. Galle, Marcus Wörns<br />

377<br />

Development of a model predicting survival of patients<br />

with recurrent or progressive hepatocellular carcinoma:<br />

MOREPROH score<br />

Sang Il Choi 1 , Ami Yu 2 , Bo Hyun Kim 1 , Eun Jeong Ko 1 , Sun Seob<br />

Park 1 , Byung Ho Nam 2,3 , Joong-Won Park 1,3 ; 1 Center for Liver<br />

Cancer, National Cancer Center, Korea, Goyang-si, Korea (the<br />

Republic of); 2 Biometric Research Branch, National Cancer Center,<br />

Korea, Goyang-si, Korea (the Republic of); 3 Department of<br />

Cancer Control and Policy, Graduate School of Cancer Science<br />

and Policy, National Cancer Center, Korea, Goyang-si, Korea (the<br />

Republic of)<br />

Introduction: Most prognostic models developed for hepatocellular<br />

carcinoma (HCC) are based on data collected at the time<br />

of initial diagnosis. However, HCC frequently recur or progress<br />

after the initial treatment and the patient’s tumor burden,<br />

underlying liver function, and performance status are certain to<br />

change over time, affecting the prognosis of the patient. Therefore,<br />

we aimed to develop a model to predict the survival of<br />

patients at the time of disease recurrence or progression after<br />

the initial treatment. Methods: 1,972 patients in the cohort that<br />

consist of patients who were newly diagnosed as HCC at the<br />

National Cancer Center, Korea between January 2004 and<br />

December 2009, were followed till May 2014, and 1,301<br />

patients who had disease recurrence or progression after the<br />

initial treatment were identified. They were randomly assigned<br />

into the development cohort (75%, n = 976) and the validation<br />

cohort (25%, n = 325). Variables that significantly influence the<br />

survival of patients were sorted out and a survival prediction<br />

model for recurrent or progressive HCC patients was formulated<br />

using the multivariate Cox proportional hazards model.<br />

Performance of the model was evaluated using the c-statistics<br />

for discrimination ability and Hosmer-Lemeshow χ2 statistics<br />

for calibration ability on the both development cohort and validation<br />

cohort. Results: According to the multivariate analysis,<br />

age, serum albumin level, Model For End-Stage Liver Disease<br />

(MELD) score, tumor factors (such as number of nodules, size of<br />

the largest nodule, vascular invasion, and extrahepatic metastasis),<br />

serum alpha-fetoprotein (AFP) level, presence of ascites,<br />

initial treatment modality, and the best response following the<br />

initial treatment were independent prognostic factors for overall<br />

survival. Using these variables, survival prediction model was<br />

developed: MOREPROH score (Model predicting survival of<br />

patients with Recurrent or Progressive HCC) MOREPROH score<br />

was tested on the development cohort and the predefined validation<br />

cohort. The respective c-statistics and χ2 statistics of<br />

this novel model for the validation cohort were 0.836 (95%<br />

confidence interval [CI] 0.806-0.865) and 6.339(p = 0.706);<br />

0.808 (95% CI 0.781-0.834) and 4.408(p = 0.883); and<br />

0.803 (95% CI 0.777-0.829) and 10.960(p = 0.278) for 1-,<br />

3-, and 5-year survival, respectively. Conclusion: A new model,<br />

MOREPROH to predict survival for patients with recurrent or<br />

progressive HCC using only objective variables was developed<br />

and validated for the first time. This model may be useful for<br />

planning of subsequent treatment strategy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sang Il Choi, Ami Yu, Bo Hyun<br />

Kim, Eun Jeong Ko, Sun Seob Park, Byung Ho Nam, Joong-Won Park<br />

378<br />

Undetected asymptomatic alcoholic cirrhosis underlies<br />

the diagnosis of hepatocellular carcinoma (HCC) out of<br />

surveillance programs in Spain. Analysis of 720 cases<br />

in 74 centers.<br />

Carlos Rodriguez-Lope 1 , Maria Elisa Reig 2 , Ana Matilla 3 , Maria<br />

Teresa Ferrer 4 , Eva Dueñas 5 , Beatriz Minguez 6 , Javier F. Castroagudin<br />

7 , Inmaculada Ortiz 8 , Sonia Pascual 9 , Jose Luis Lledo 10 ,<br />

Adolfo Gallego 11 , Juan I. Arenas 12 , Carles Aracil 13 , Montserrat<br />

Forne 14 , Carolina Muñoz 15 , Fernando Pons 16 , Margarita<br />

Sala 17 , Mercedes Iñarrairaegui 18 , Marta Martin-llahi 19 , Victoria<br />

Andreu 20 , Carmen Garre 21 , Paloma Rendon 22 , Javier Fuentes 24 ,<br />

Javier Crespo 1 , Manuel Rodríguez 23 , Jordi Bruix 2 , Maria Varela 23 ;<br />

1 Liver Unit, Hospital Universitario Marqués de Valdecilla. IDIVAL,<br />

Santander, Spain; 2 Liver Unit. BCLC group., Hospital Clinic de<br />

Barcelona, Barcelona, Spain; 3 Gastroenterology and Hepatology,<br />

H.G.U. Gregorio Marañon, Madrid, Spain; 4 Gastroenterology<br />

and hepatology, H. Virgen del Rocío, Sevilla, Spain; 5 Gastroenterology<br />

and Hepatology, H. de Bellvitge, Hospitalet de Llobregat,<br />

Spain; 6 Liver Unit, H. Vall d’Hebrón, Barcelona, Spain; 7 H. Clínico<br />

Universitario de Santiago, Santiago de Compostela, Spain; 8 H.<br />

Universitario Doctor Peset, Valencia, Spain; 9 H.G.U. de Alicante,<br />

Alicante, Spain; 10 H. U. Ramón y Cajal, Madrid, Spain; 11 H. de la<br />

Santa Creu i Sant Pau, Barcelona, Spain; 12 H. Donostia, Donostia,<br />

Spain; 13 H. U. Arnau de Vilanova, Lleida, Spain; 14 H. U. Mutua<br />

de Terrassa, Terrassa, Spain; 15 H. U. 12 de Octubre, Madrid,<br />

Spain; 16 H. U. Puerta de Hierro, Madrid, Spain; 17 H. U. Germans<br />

Trias i Pujol, Badalona, Spain; 18 Clinica Universitaria de Navarra,<br />

Pamplona, Spain; 19 H. Moises Broggi, Sant Joan Despi, Spain;<br />

20 H. de Viladecans, Barcelona, Spain; 21 H. U. Virgen de la Arrixaca,<br />

Murcia, Spain; 22 H. Puerta del Mar, Cadiz, Spain; 23 Liver<br />

Unit, H.U. Central de Asturias, Oviedo, Spain; 24 H.U. Miguel Servet,<br />

Zaragoza, Spain<br />

Introduction. Between Oct’08-Jan’09 we performed a survey<br />

of HCC with 705 cases in 62 centers in Spain[Med Clin<br />

(Barc).2010;134(13):569-7].In that study 53% of patients<br />

were diagnosed out of surveillance with more advanced HCC<br />

and less radical therapies applied. We have performed a new<br />

study to analyze the reasons of not performing surveillance.<br />

Methods. We invited previous participants extending the invitation<br />

to any center in Spain via Spanish Association for the<br />

Study of the Liver website. Investigators prospectively registered<br />

demographic, clinical and tumor characteristics, first treatment<br />

and eligibility for liver transplantation (OLT). Results. Between<br />

Oct’14-Jan’15, 74 centers registered 720 new cases of primary<br />

liver cancer: HCC (n=686), intrahepatic cholangiocarcinoma<br />

(ICC) (n=29), mixed HCC-ICC (n=2), others (n=3).<br />

HCC characteristics were: men 82%; age 67 years; cirrhosis<br />

87%. Main etiologies: alcohol 35%, HCV 30%, HCV + alcohol<br />

15%, NAFLD 6%; Child-Pugh 6 (5-13); number of nodules 1<br />

(1-20); size of the largest 3cm (1–24); vascular invasion 18%;<br />

extrahepatic disease 8%. BCLC stages: 0: 11%, A: 43%, B:<br />

20%, C: 16% and D: 11%. Main treatments: percutaneous<br />

ablation (22%), TACE (23%), resection (11%) and sorafenib<br />

(11%). 10% (n=67) of patients were evaluated for OLT. 53%<br />

(n=316) were diagnosed outside surveillance: 269 (76%) by<br />

chance or at the same time that their liver disease; 63 (18%)


402A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

were not adherent to surveillance; and 21 (6%) had chronic<br />

liver disease without a clear cirrhosis. Patients diagnosed out<br />

of surveillance had more advanced BCLC (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 403A<br />

381<br />

Prediction of Hepatocellular Carcinoma Risk in Patients<br />

with Cirrhosis: Validation of ADRESS–HCC Model<br />

Ju Dong Yang, Hager Ahmed Mohammed, Jonggi Choi, Gregory<br />

J. Gores, Lewis R. Roberts, W. Ray Kim; Division of Gastroenterology<br />

and Hepatology, Mayo Clinic College of Medicine, Rochester,<br />

MN<br />

Background/Aims: ADRESS-HCC is a model to predict the risk<br />

of hepatocellular carcinoma (HCC) in patients with cirrhosis<br />

[Cancer 2014;3485]. It incorporates age, diabetes, race,<br />

etiology and severity of cirrhosis, and sex. In this study, we<br />

validate the model in an independent cohort of patients. Methods:<br />

Of all cirrhotic patients seen at Mayo Clinic Rochester<br />

between January 2006 and December 2011, those who had<br />

liver images (US, CT, or MRI) at baseline and at follow up more<br />

than 6 months from baseline were included. The ADRESS risk<br />

score was calculated by the following: score= [age]*0.0532+<br />

[1 for diabetes]*0.2135 + [1 for nonwhite/Hispanic]*0.2058<br />

+ [1 for metabolic/alcohol etiology]*0.3509 + [1 for viral<br />

etiology]*1.246 + [1 for male]*0.5114 + [Child-Turcotte-Pugh<br />

(CTP) score]*0.1170. Observed incidence of HCC was compared<br />

to the model’s prediction. Results: A total of 739 cirrhotic<br />

patients met the eligibility criteria. The mean age was 57.4<br />

years, with 59% men, 90% non-Hispanic White and 34% diabetic.<br />

The etiology of liver disease was alcoholic in 32%, viral<br />

in 23% and metabolic in 23%. The mean CTP score was 8.<br />

After a median follow up of 38 months, 69 (9%) patients developed<br />

HCC. The median ADRESS score was 5.01 (interquartile<br />

range: 4.44-5.46). Figure divides patients into 3 groups by<br />

the ADRESS score quartiles: the top (high risk) quartile (hazard<br />

ratio [HR]=6.8; 95%CI=2.9-20.2) and the middle two quartiles<br />

(HR=3.2; 95%CI=1.4-9.4) had a significantly higher incidence<br />

of HCC compare to the lowest risk quartile. When the surveillance<br />

threshold (ADRESS=4.67 with predicted annual incidence=1.5%)<br />

was applied, 0.97% of patients developed HCC<br />

in the first 18 months. Conclusions: Based on an independent<br />

cohort of cirrhotic patients, we validate ADRESS-HCC model as<br />

a useful tool for predicting the risk of HCC among patients with<br />

cirrhosis and identifying candidates for surveillance.<br />

Disclosures:<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

The following authors have nothing to disclose: Beatriz Minguez, Ayse Aytaman,<br />

Meritxell Ventura-Cots, Maaz B. Badshah, Caitlin C. Citti, Heather L. Platt, Ting-Yi<br />

Chen, Luciana Kikuchi, Sonja Marcus, Michael Yin, Judith Aberg, Myron E.<br />

Schwartz, Norbert Bräu<br />

Disclosures:<br />

Gregory J. Gores - Advisory Committees or Review Panels: Conatus<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

The following authors have nothing to disclose: Ju Dong Yang, Hager Ahmed<br />

Mohammed, Jonggi Choi<br />

382<br />

The life time frequency and interval of treatments define<br />

the prognosis of hepatocellular carcinoma after radical<br />

radiofrequency ablation therapy<br />

Takamasa Ohki, Yuki Hayata, Satoshi Kawamura, Daisaku Ito,<br />

Koki Kawanishi, Kentaro Kojima, Michiharu Seki, Nobuo Toda,<br />

Kazumi Tagawa; Gastroenterology, Mitsui Memorial Hospital,<br />

Tokyo, Japan<br />

Backgrounds and Aims: Hepatocellular carcinoma (HCC)<br />

frequently recurs even after curative radiofrequency ablation<br />

(RFA) therapy and the recurrence rate is approximately 50%<br />

within 3 years. Thus, the treatment strategy after HCC recurrence<br />

is very important to achieve a long term survival. We<br />

analyzed the life time frequency and interval of treatments to<br />

evaluate their impacts on overall survival (OS). Patients and<br />

methods: In this retrospective study, we enrolled 305 naïve<br />

HCC patients who were curatively treated with RFA between<br />

January 1 2000 and December 31 2010. The last follow-up<br />

date was December 31 2014. The rate of life time frequency of<br />

treatments per year (LFT rate) was calculated as total numbers<br />

of treatments divided by observational year. The total numbers<br />

of treatments were defined as sum of frequency of RFA and<br />

trans-catheter chemo-embolization or infusion (TACE or TAI)<br />

times. Patients were stratified into four groups to compare the<br />

OS according to LFT rate: group A; LFT rate < 0.5 N = 53,<br />

group B; 0.5 ≤ LFT rate < 1.0 N = 69, group C; 1.0 ≤ LFT rate<br />

< 2.0 N = 103, and group D; 2.0 ≤ LFT rate N = 80. Results:<br />

During the mean follow-up of 4.7 years, the life time frequency<br />

of treatments was 5.2 times on average. The frequency of each<br />

treatment modality was as follows: RFA 2.8 and TACE or TAI<br />

2.3 times. One-hundred fifty-eight patients died during the follow-up<br />

period, showing cumulative survival rates at 3, 5 and


404A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

10 years of 100%, 96.0%, 81.2% in group A, 95.3%, 77.6%<br />

and 45.5% in group B, 70.1%, 41.7% and 8.7% in group C,<br />

and 50.4%, 22.2% and 0% in group D, respectively (P < 0.01<br />

by the log-rank test). Adjusting for significant factors in univariate<br />

analysis, multivariate analysis indicated that frequency of<br />

receiving RFA (HR: 0.60 pre 1 time, P < 0.01), frequency of<br />

receiving TACE or TAI (HR: 0.88 per 1 time, P < 0.01) and<br />

LFT rate over 1.0 (HR: 29.4, P < 0.01) as independent factors<br />

which affected on OS. First recurrence, tumor sizes, and tumor<br />

numbers were not significant risk factors in the univariate analysis.<br />

Conclusion: Appropriate treatment at the time of each HCC<br />

recurrence remarkably prolonged OS even if HCC patients had<br />

multiple relapse. Repeat RFA and TACE or TAI were promising<br />

strategy to achieve a long term survival. However, LFT rate<br />

over 1.0 which indicated high rate of HCC recurrence was an<br />

independent factor of poor prognosis suggesting the malignant<br />

potential of HCC.<br />

Disclosures:<br />

Takamasa Ohki - Speaking and Teaching: Otsuka Pharmaceutical Co. Ltd.<br />

The following authors have nothing to disclose: Yuki Hayata, Satoshi Kawamura,<br />

Daisaku Ito, Koki Kawanishi, Kentaro Kojima, Michiharu Seki, Nobuo Toda,<br />

Kazumi Tagawa<br />

383<br />

Efficacy and safety of miriplatin versus epirubicin in<br />

transcatheter arterial chemoembolization therapy for<br />

advanced hepatocellular carcinoma: a randomized controlled<br />

trial<br />

Tatehiro Kagawa 1 , Shunji Hirose 1 , Yoshitaka Arase 1 , Kota<br />

Tsuruya 1 , Kazuya Anzai 1 , Sei-ichiro Kojima 1 , Koichi Shiraishi 1 ,<br />

Masako Shomura 3 , Jun Koizumi 2 , Tetsuya Mine 1 ; 1 Gastroenterology,<br />

Tokai University, Isehara, Japan; 2 Radiology, Tokai University<br />

School of Medicine, Isehara, Japan; 3 Nursing, Tokai University<br />

School of Medicine, Isehara, Japan<br />

Purpose: Miriplatin, a hydrophobic platinum-based anticancer<br />

agent, is used for transcatheter arterial chemoembolization<br />

(TACE), but its efficacy and safety are still unclear. To clarify<br />

the efficacy and safety of TACE with miriplatin in the treatment<br />

of patients with advanced hepatocellular carcinoma (HCC),<br />

we performed a randomized trial to compare miriplatin with<br />

epirubicin. Patients and Methods: We randomly assigned<br />

patients with advanced HCC to receive TACE with miriplatin<br />

or TACE with epirubicin after stratification by the tumor<br />

diameter (3 cm). Enrollment criteria were (1) presence of HCC<br />

with contrast-enhancement in the arterial phase of dynamic CT<br />

scan, (2) no indication of surgical resection or local therapy<br />

including radiofrequency ablation and percutaneous ethanol<br />

injection, (3) no previous treatment for the target HCC, (4) no<br />

extrahepatic metastasis, (5) no tumor thrombus in the portal<br />

or hepatic veins, (6) Child-Pugh class A or B, (7) ECOG PS<br />

score of 2 or less, and (8) serum bilirubin < 3 mg/dl, creatinine<br />

< 1.5 mg/dl, white blood cell count ≥ 3000/mm 3 ,<br />

platelet count ≥ 50000/mm 3 , and hemoglobin concentration<br />

≥ 9.5 g/dl. Dynamic CT was performed within 3 months after<br />

TACE and the efficacy was blindly evaluated by two radiologists<br />

according to the modified Response Evaluation Criteria in<br />

Solid Tumors (mRECIST). Overall survival and recurrence-free<br />

survival were compared by log-rank test. The treatment modality<br />

for the recurrence was decided by the primary doctors.<br />

This study was approved by the institutional ethics committee<br />

(UMIN000003987) and written informed consent was<br />

obtained from each patient. Results: 23 and 24 patients were<br />

allocated to the miriplatin (M) group and the epirubicin (E)<br />

group, respectively. No significant difference was observed in<br />

age gender, etiology, tumor size, Child-Pugh sore, and clinical<br />

stage between two groups. The rate of complete response in<br />

the M group was significantly lower than in the E group (33%<br />

vs 79%, P < 0.01). Recurrence-free survival was significantly<br />

shorter in the M group than in the E group (P < 0.05); 1-yr<br />

recurrence-free survival rate was 18% in the M group and 33%<br />

in the E group. OS was not significantly different between two<br />

groups; 1-yr and 3-yr OS were 95% and 58%, respectively, in<br />

the M group and 88% and 79%, respectively, in the E group.<br />

As adverse effects grade 2 fever was observed with similar<br />

frequency; 19% in the M group and 32% in the E group. Conclusions:<br />

This randomized trial demonstrated that the efficacy<br />

of the miriplatin-TACE was inferior to the conventional epirubicin-TACE,<br />

although miriplatin was as safe as epirubicin.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tatehiro Kagawa, Shunji Hirose,<br />

Yoshitaka Arase, Kota Tsuruya, Kazuya Anzai, Sei-ichiro Kojima, Koichi Shiraishi,<br />

Masako Shomura, Jun Koizumi, Tetsuya Mine<br />

384<br />

Identification of candidate extracellular non-coding RNA<br />

biomarkers for Hepatocellular Carcinoma<br />

Irene K. Yan, Tushar Patel; Mayo Clinic, Jacksonville, FL<br />

Hepatocellular carcinoma (HCC) has a poor prognosis. Early<br />

detection may improve outcomes by detecting the cancer at a<br />

stage when curative procedures may be possible. Our overall<br />

goals are to evaluate the utility of extracellular non-coding<br />

RNA as biomarkers for the early detection of HCC. Towards<br />

this goal, we sought to identify non-coding extracellular RNA<br />

(exRNA) that are associated with HCC cells, and to develop<br />

assays for their detection. METHODS: A systematic profiling<br />

approach evaluated non-coding RNA that are (1) associated<br />

with malignant HCC cells and (2) expressed within extracellular<br />

vesicles (EV) released from these cells. qPCR, Nanostring<br />

and droplet digital PCR were used for the detection of<br />

exRNA from culture supernatants or from serum of patients with<br />

HCC. Profiling was performed on RNA from cells and exRNA<br />

released by these cells for non-malignant (HH and THLE-2), and<br />

malignant HCC cells (Hep3B, HepG2, PLC/PRF/5, SNU-182,<br />

SNU-398). MicroRNA expression was assessed using qPCR<br />

panels from Exiqon to screen 752 miRNAs, and expression of<br />

90 selected lncRNAs was assessed using lncRNA qPCR assays.<br />

The expression of 800 miRNAs and 33 lncRNAs was analyzed<br />

in selected samples using Nanostring. Assay precision<br />

and reproducibility were individually determined for selected<br />

exRNA candidate biomarkers. RESULTS: HCC cells released<br />

a larger amount of extracellular RNA compared to normal<br />

hepatocytes (average of 1.23% vs. 0.55% of total donor cell<br />

RNA). Compared to exRNA from normal hepatocytes, qPCR<br />

analyses identified 14 lncRNAs and 33 miRNAs that were<br />

increased in exRNA from HCC cells. In addition, Nanostring<br />

analysis identified three miRNAs (miR-25, miR-302d, and miR-<br />

612) in exRNA preparations from all HCC cells, and from<br />

serum obtained from patients with HCC. These three microR-<br />

NAs were also detected using droplet digital PCR but not using<br />

qPCR. Droplet digital PCR assays for long non-coding RNA<br />

GAS5 and H19 in exRNA had a CV of 10.6% and 7.1% for<br />

precision, and a CV of 11% and 3% for reproducibility respectively.<br />

Digital PCR assays and nanostring were more sensitive<br />

than qPCR assays for the detection of low abundance exRNA<br />

in serum. SUMMARY and CONCLUSIONS: Non-coding RNAs<br />

that are selectively enriched within exRNA from HCC cells can<br />

be detected in the circulation. Using a novel biomarker discovery<br />

platform based on exRNA release, we have identified<br />

several candidate microRNA and long non-coding RNA biomarkers<br />

for HCC. Sensitive assay platforms have been devel-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 405A<br />

oped for use in <strong>studies</strong> to validate the utility of these exRNA as<br />

diagnostic markers for HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Irene K. Yan, Tushar Patel<br />

385<br />

The radiofrequency ablation is eligible treatment for<br />

the case of hepatocellular carcinoma more than 3 cm<br />

in diameter with Child-Pugh class A or des-gamma-carboxy<br />

prothrombin level less than 200 mAU/mL: Analysis<br />

of 1815 cases in a single center for fourteen-year<br />

experience<br />

Takashi Tanaka, Akira Anan, Daisuke Morihara, Kazuhide<br />

Takata, Keiji Yokoyama, Yasuaki Takeyama, Makoto Irie, Satoshi<br />

Shakado, Tetsuro Sohda, Shotaro Sakisaka; Fukuoka university<br />

school of medicine, Hepatology, Fukuoka, Japan<br />

Background and Aims: Most of the hepatocellular carcinoma<br />

(HCC) more than 3 cm in diameter is categorized into intermediate<br />

stage of Barcelona Clinic Liver Cancer (BCLC) staging<br />

system, and first line treatment in the intermediate stage is recommended<br />

transcatheter arterial chemoembolization (TACE).<br />

Although general indication of radiofrequency ablation (RFA)<br />

treatment for HCC is defined as tumor size less than 3 cm in<br />

diameter, some previous trials described RFA for HCC more<br />

than 3 cm in diameter. The aim of this study was to analyze the<br />

long term outcome and prognostic factors in the patients who<br />

had received RFA treatment for HCC more than 3 cm in diameter.<br />

Methods: A total of 1815 cases with liver tumors, from<br />

April 2000 to September 2014 were received RFA treatment<br />

in Fukuoka University Hospital. Among them, 94 cases were<br />

enrolled in this study, which was defined as patients with single<br />

or multinodular HCC more than 3 cm in diameter. Overall survival<br />

and recurrence rates were analyzed by the Kaplan-Meier<br />

method and risk factors on prognosis was assessed by multivariate<br />

Cox regression hazard model. Results: The median age of<br />

the study population (men 62 and women 32) was 71 (range:<br />

48-88) years old. Patient’s Child-Pugh class was in A (n=72)<br />

and B (n=22) and 77 patients related with viral hepatitis (HBV:<br />

n=4, HCV: n=73). The median tumor size was 3.5 cm in diameter<br />

(range: 3.1-6.2 cm), and the number of tumor was single:<br />

n=51, double: n=20, triple: n=6, more than four: n=7. We<br />

declined “up to 6” group who have the number of the tumor<br />

plus maximal size (cm) less than six, 66 patients were within<br />

up to 6 group. The median serum level of alpha-fetoprotein<br />

(AFP) and des-gamma-carboxy prothrombin (DCP) level was 35<br />

(range: 2-3061) and 99 (range: 8-12502) respectively. Cumulative<br />

overall survival rates at 1-, 3-, 5-, 10-years were 93.4%,<br />

62.7%, 38.4%, and 12.7%, respectively. Cumulative disease<br />

free survival rate at 1-, 3-, 5-, 10-years were 51.4%, 19.4%,<br />

15.3%, and 4.6%, respectively. Multivariate analysis revealed<br />

that factors associated with survival were tumor recurrence free,<br />

(HR=3.428; 95%CI 1.407-11.421, p=0.0044), Child-Pugh<br />

class A (HR=2.136; 95%CI 1.233-3.633, p=0.0084), and<br />

DCP levels less than 200 mAU/mL (HR=1.957; 95%CI 1.198-<br />

3.159, p=0.0078). Multivariate analysis disclosed that without<br />

up to 6 (HR=1.714; 95%CI 1.066-2.708, p=0.027) predicted<br />

higher incidence of developing recurrence. Conclusion: From<br />

our over 1800 RFA experiences in fourteen years, we revealed<br />

that radiofrequency ablation is eligible treatment for the case<br />

of hepatocellular carcinoma more than 3 cm in diameter with<br />

Child-Pugh class A or DCP level less than 200 mAU/ml.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takashi Tanaka, Akira Anan, Daisuke<br />

Morihara, Kazuhide Takata, Keiji Yokoyama, Yasuaki Takeyama, Makoto<br />

Irie, Satoshi Shakado, Tetsuro Sohda, Shotaro Sakisaka<br />

386<br />

Racial Differences in the Presentation and the Incidence<br />

of Hepatocellular Carcinoma (HCC) in a Large Cohort of<br />

Patients with Cirrhosis in the United States<br />

Long H. Nguyen 1 , Changqing Zhao 2,5 , Nghia H. Nguyen 3,2 ,<br />

Joseph K. Hoang 2 , Michael D. Nguyen 2 , Richard H. Le 2 , Peter T.<br />

Nguyen 2,6 , Derek Lin 1 , Vinh D. Vu 2 , Lee Ann Yasukawa 4 , Susan C.<br />

Weber 4 , Mindie H. Nguyen 2 ; 1 Department of Medicine, Stanford<br />

University Medical Center, Palo Alto, CA; 2 Division of Gastroenterology<br />

and Hepatology, Stanford University Medical Center, Palo<br />

Alto, CA; 3 Department of Medicine, University of California, San<br />

Diego, San Diego, CA; 4 Center for Clinical Informatics, Stanford<br />

University School of Medicine, Palo Alto, CA; 5 Department of Cirrhosis,<br />

Institute of Liver Disease, Shuguang Hospital, Shanghai<br />

University of T.C.M., Shanghai, China; 6 School of Medicine, University<br />

of Texas Medical Branch, Galveston, Galveston, TX<br />

Purpose: Cirrhosis regardless of etiology is a significant HCC<br />

risk factor. The purpose of our study is to clarify the largely<br />

unknown role of ethnicity in the development of HCC in patients<br />

with cirrhosis. Methods: We conducted a retrospective cohort<br />

study of 3208 consecutive white (n=1,692), Hispanic (n=772),<br />

Asian (n=614), and black (n=130) patients with cirrhosis seen<br />

at a single university tertiary care center from 2005-2010. Our<br />

primary endpoint was HCC diagnosis by histology or non-invasive<br />

criteria. Results: Among ethnic groups, patients were<br />

similarly aged (54-59 years), mostly male (61-64%), and with<br />

similar baseline MELD scores (17-23), while black patients had<br />

much shorter duration of follow up (22 vs. 31-40 months). At<br />

presentation, the highest HCC prevalence was seen in Asian<br />

(11%) then Hispanic (5%), white (4%), and black patients (4%,<br />

p


406A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

patients were less likely to meet Milan criteria for liver transplantation<br />

(LT) compared to HCV and nonviral patients: 34%,<br />

50%, 39%, P < 0.0001). HBV patients were more likely than<br />

HCV and non-viral patients to undergo resection (26%, 9%,<br />

and 13%, respectively; P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 407A<br />

older (mean ages 60 ± 13 vs. 55 ± 10, p


408A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

RFA vs. 52.5% in HR (p=0.677). On 3cm < size ≤ 5cm, cumulative<br />

DFS were 29.6% in TACE and RFA vs. 40.7% in HR<br />

(p=0.513). Most common complication on both group was<br />

elevation of AST/ALT (grade 2), most severe complication were<br />

happened 3 cases of HR group. One is severely elevated bilirubin<br />

without encephalopathy; the others were death due to<br />

renal failure or sepsis. In TACE with RFA group, liver segmental<br />

infarction were happened in 2 cases. Conclusions: Combined<br />

TACE and RFA could be alternative choice of initial treatment<br />

in early HCC patients.<br />

Disclosures:<br />

Kwan Soo Byun - Grant/Research Support: Gilead, BMS<br />

The following authors have nothing to disclose: Young Kul Jung, Sang Jun Suh,<br />

Hyung Joon Yim, Oh Sang Kwon, Yun Soo Kim, Duck Joo Choi, Ju Hyun Kim, Ji<br />

Hoon Kim, Jong Eun Yeon, Yeon Seok Seo, Soon Ho Um, Ho Sang Ryu<br />

392<br />

Myeloid-derived suppressor cells correlate with patient<br />

outcomes in hepatic arterial infusion chemotherapy for<br />

hepatocellular carcinoma<br />

Eishiro Mizukoshi, Tatsuya Yamashita, Kuniaki Arai, Takeshi Terashima,<br />

Masaaki Kitahara, Hidetoshi Nakagawa, Noriho Iida,<br />

Kazumi Fushimi, Shuichi Kaneko; Kanazawa Univ, Kanazawa,<br />

Japan<br />

Background: The modulation of tumor-associated host immune<br />

responses after cancer treatments has recently been reported.<br />

For example, hepatocellular carcinoma (HCC) treatments with<br />

radiofrequency ablation or transarterial chemoembolization<br />

were found to increase the frequency of tumor-associated antigen<br />

(TAA)-specific T cells due to the creation of an antigen<br />

source by the destruction of tumor cells. In addition, recent<br />

<strong>studies</strong> demonstrated that beyond their direct cytotoxic effects<br />

on cancer cells, several conventional chemotherapeutic drugs<br />

promoted the elimination or inactivation of regulatory T cells<br />

(Tregs) or myeloid-derived suppressor cells (MDSCs), resulting<br />

in enhanced host antitumor immunity. Hepatic arterial infusion<br />

chemotherapy (HAIC) has been employed as an alternative<br />

therapy to sorafenib for the patients with advanced HCC in<br />

Japan and other Asian countries. In the present study, we<br />

performed a comparative analysis of various immune cell<br />

responses including TAA-specific T cells, Tregs and MDSCs in<br />

advanced HCC patients treated with HAIC, and analyzed the<br />

relationship between these immune cell responses and patient<br />

prognosis. Methods: Thirty-six HLA-A24-positive patients who<br />

had clinically diagnosed advanced HCC were examined in the<br />

present study. After being diagnosed, all patients were treated<br />

with HAIC using interferon (IFN)/ 5-fluorouracil (FU) or IFN/<br />

FU + cisplatin. Interferon gamma ELISPOT assays were performed<br />

to examine the frequency of TAA-specific T cells using<br />

11 peptides derived from 5 TAAs. The frequencies of Tregs and<br />

MDSCs were examined by multicolor fluorescence-activated<br />

cell sorting analysis and analyzed the association with clinical<br />

factors and prognosis of patients. Results: The frequency of<br />

Tregs was significantly decreased after HAIC but no significant<br />

differences were observed in the frequency of TAA-specific T<br />

cells and MDSCs. An analysis of prognostic factors for overall<br />

survival identified diameter of the tumor (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 409A<br />

tic thrombosis had better survival when treated with anticoagulant<br />

(log-rank test p=0.001). Conclusion: Portal vein thrombosis<br />

in HCC patients is non-malignant in half of the patients, calling<br />

for anticoagulant therapy. The prognosis of patients with neoplastic<br />

and non neoplastic thrombosis remain poor.<br />

Disclosures:<br />

Massimo Colombo - Advisory Committees or Review Panels: BRISTOL-MEY-<br />

ERS-SQUIBB, SCHERING-PLOUGH, ROCHE, GILEAD, BRISTOL-MEYERS-SQUIBB,<br />

SCHERING-PLOUGH, ROCHE, GILEAD, Janssen Cilag, Achillion; Grant/<br />

Research Support: BRISTOL-MEYERS-SQUIBB, ROCHE, GILEAD, BRISTOL-MEY-<br />

ERS-SQUIBB, ROCHE, GILEAD; Speaking and Teaching: Glaxo Smith-Kline,<br />

BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH, ROCHE, NOVARTIS, GILEAD,<br />

VERTEX, Glaxo Smith-Kline, BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH,<br />

ROCHE, NOVARTIS, GILEAD, VERTEX, Sanofi<br />

The following authors have nothing to disclose: Angelo Sangiovanni, Cristina<br />

Bertelli, Sara Vavassori, Giuseppina Pisano, Massimo Iavarone, Silvia Fargion,<br />

Anna Ludovica Fracanzani<br />

394<br />

Preclinical characterization of a novel first in class<br />

GNS561: autophagy inhibitor therapeutic candidate for<br />

treatment of hepatocellular carcinoma<br />

Firas Bassissi 1 , Sonia Brun 1 , Jerome Courcambeck 1 , Gregory<br />

Nicolas 1 , Clarisse Dubray 1 , Antoine Beret 1 , Paul Castellani 2 ,<br />

Xavier Adhoute 2 , Marc Bourlière 2 , Jean-Luc Raoul 3 , Philippe Halfon<br />

1 ; 1 Genoscience Pharma, Marseille, France; 2 Hepatology,<br />

Hopital Saint-Joseph, Marseille, France; 3 Oncology, Institut Paoli<br />

Calmettes, Marseille, France<br />

Background: In spite of successful approval and wide application<br />

of sorafenib, the prognosis for patients with advanced<br />

hepatocellular carcinoma (HCC) remains poor. Fortunately,<br />

there have been renewed and continued interests and active<br />

research in developing other molecularly targeted agents in<br />

HCC during the past few years. While there is early evidence<br />

of antitumor activity of several agents in phase I/II <strong>studies</strong>,<br />

phase III efforts with a few targeted agents have failed, highlighting<br />

the challenges in new drug development in HCC. For<br />

this reason, development of innovative drugs with original<br />

mechanism of action could improve treatment of HCC patients<br />

Material and methods: We screened a library of autophagy<br />

inhibitor compounds and identified a new drug GNS561 as<br />

compound that kills tumor cells in a panel of HCC cell lines<br />

and primary tumor. Drug tolerance and plasma and liver pharmacokinetic<br />

were evaluated after single and repeated dosing<br />

in mice and rat. Results: GNS561 demonstrates autophagy<br />

inhibition and apoptosis induction activities related to lysosome<br />

disruption. GNS 561 shows potent anti-proliferation activity<br />

when assayed against a panel of human tumor cell lines, notably<br />

against HCC cell lines (Mean EC50 2mM). In addition<br />

GNS561 demonstrates potent activity against a panel of HCC<br />

patient tumors even in those with sorafenib resistance (Mean<br />

EC50 3mM vs 11mM for sorafenib). GNS561 is highly selectively<br />

trapped in the liver, via uptake hepatocyte transporters.<br />

Plasma and liver PK in mice and rats after single and repeated<br />

doses confirm this selectivity (exposure ratio liver/plasma<br />

about 20 in mice) with a good oral bioavailability. Tolerance<br />

of single and repeated doses is excellent in both rats and mice<br />

Conclusions: Our results provide a rationale for testing autophagy<br />

flux disruption as a novel therapeutic strategy for HCC.<br />

GNS561 is a liver selective drug which offers great promise<br />

for HCC treatment of non-responder or nontolerant patients to<br />

sorafenib. This compound is selected as a drug candidate for<br />

future investigation in HCC clinical trials.<br />

Disclosures:<br />

Xavier Adhoute - Speaking and Teaching: bayer<br />

Marc Bourlière - Advisory Committees or Review Panels: Schering-Plough,<br />

Bohringer inghelmein, Schering-Plough, Bohringer inghelmein, Transgene; Board<br />

Membership: Bristol-Myers Squibb, Gilead, Idenix; Consulting: Roche, Novartis,<br />

Tibotec, Abott, glaxo smith kline, Merck, Bristol-Myers Squibb, Novartis, Tibotec,<br />

Abott, glaxo smith kline; Speaking and Teaching: Gilead, Roche, Merck, Bristol-Myers<br />

Squibb<br />

Jean-Luc Raoul - Advisory Committees or Review Panels: Bayer SP, BMS, BTG,<br />

Arqule; Speaking and Teaching: Bayer SP<br />

The following authors have nothing to disclose: Firas Bassissi, Sonia Brun, Jerome<br />

Courcambeck, Gregory Nicolas, Clarisse Dubray, Antoine Beret, Paul Castellani,<br />

Philippe Halfon<br />

395<br />

Survival of Patients with Advanced Hepatocellular<br />

Carcinoma Treated with Sorafenib: An Analysis of<br />

SEER-Medicare Database<br />

Neehar D. Parikh 1 , Vincent D. Marshall 2 , Hari Nathan 3 , Rajesh<br />

Balkrishnan 2 , Vahakn Shahinian 4 ; 1 Division of Gastroenterology,<br />

University of Michigan, Ann Arbor, MI; 2 College of Pharmacy,<br />

University of Michigan, Ann Arbor, MI; 3 Department of Surgery,<br />

University of Michigan, Ann Arbor, MI; 4 Kidney Epidemiology and<br />

Cost Center, University of Michigan, Ann Arbor, MI<br />

Sorafenib, a multikinase inhibitor, is the only chemotherapeutic<br />

approved for use in the US for the treatment of advanced HCC.<br />

We aimed to examine the outcomes of elderly patients with<br />

advanced HCC in the US. We performed a secondary analysis<br />

of continuously enrolled Medicare beneficiaries with HCC diagnoses<br />

from 2007-2009, based on SEER diagnosis codes. Our<br />

primary aim was to determine survival in patients treated with<br />

sorafenib and predictors of survival. We compared advanced<br />

stage patients with HCC (stage III/IV) who received sorafenib<br />

within 6 months of diagnosis to advanced stage patients with<br />

HCC who received no therapy (control.) We calculated the<br />

Charlson comorbidity index 12 months prior to diagnosis and<br />

an aggregate variable for decompensated cirrhosis (presence<br />

of esophageal varices or variceal banding, ascites or paracentesis,<br />

hepatic encephalopathy or use of neomycin, lactulose,<br />

or rifaximin.) We performed univariate (Kaplan Meier with<br />

log rank test) and multivariate (Cox regression) analyses for<br />

predictors of survival. The baseline characteristics of the patient<br />

are shown in the Table. Survivial was singificantly better in the<br />

sorafenib treated group (p


410A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Vahakn Shahinian - Consulting: Amgen, Inc<br />

The following authors have nothing to disclose: Neehar D. Parikh, Vincent D.<br />

Marshall, Hari Nathan, Rajesh Balkrishnan<br />

396<br />

The Casein kinase II (CK2) inhibitor CX-4945 has<br />

an additive effect with gemcitabine and cisplatin on<br />

cholangiocarcinoma (CCA) cell lines mediated in part<br />

through PI3K/Akt inhibition<br />

Kais Zakharia 1 , Katsuyuki Miyabe 1 , Catherine D. Moser 1 , Sean<br />

O’Brien 3 , Mitesh J. Borad 2 , Lewis R. Roberts 1 ; 1 Division of Gastroenterology<br />

and Hepatology, Mayo Clinic College of Medicine,<br />

and Mayo Clinic Cancer Center, Rochester, MN; 2 Division of<br />

Hematology, Division of Oncology, Mayo Clinic, Scottsdale, AZ;<br />

3 Senhwa Biosciences, San Diego, CA<br />

Background: There is a need to develop effective therapies for<br />

CCA due to suboptimal outcomes with the current standard<br />

gemcitabine (Gem) and cisplatin (Cis) therapy. Casein kinase<br />

II (CK2) is a serine/threonine selective protein kinase that has<br />

been implicated in cancer progression, making it an attractive<br />

target for anticancer treatment. CX-4945, a first in class, orally<br />

available and highly selective inhibitor of CK2 has shown antitumor<br />

activity in other human cancers. In this study we investigated<br />

the anti-neoplastic effect of CX-4945 on 3 different<br />

CCA cell lines. Aims: To determine 1) whether CX-4945 has<br />

an anti-proliferative effect on CCA cells; 2) if CX-4945 induces<br />

apoptosis in CCA cells; 3) whether the anti-proliferative action<br />

of CX-4945 occurs through effects on PI3K/Akt signaling; and<br />

4) whether the effect of CX-4945 enhances that of Gem or Cis<br />

in vitro. Methods: CCK8 and colony formation assays were<br />

used to assess cell viability in cultures of HuCCT1 (intrahepatic<br />

CCA), EGI-1 (extrahepatic CCA), and a new primary patient<br />

derived CCA cell line LIV27 after treatment with CX-4945<br />

alone or in combination with Gem or Cis. The Caspase 3/7<br />

assay was used to assess the effect of CX-4945 on CCA apoptosis.<br />

The half maximal inhibitory concentrations (IC50) were<br />

calculated using the Chou-Talalay technique. Western immunoblotting<br />

was used to assess the effect of CX-4945 on Akt<br />

pathway signaling in CCA cells. Results: CX-4945 significantly<br />

decreased cell viability of HuCCT1, EGI-1 and LIV27 in a timeand<br />

dose-dependent manner (IC50 of 7.3, 9.5, and 9.4 μM<br />

respectively at 72 hours). CX-4945 inhibited colony formation<br />

in HuCCT1 and EGI-1 (IC50 of 3.8 and 1.6 μM respectively).<br />

CX-4945 did not significantly increase Caspase 3/7 activity<br />

in CCA cells until higher doses (15 and 20 μM) were used.<br />

CX-4945 inhibited Akt phosphorylation (at Serine473 and<br />

Serine129) and PTEN phosphorylation, suggesting that it acts<br />

through the PI3K/Akt pathway. Addition of CX-4945 to Cis<br />

or Gem decreased viability of the HuCCT1, EGI-1 and LIV27<br />

cells by 15-40% (depending on the dose and the cell line)<br />

with a combination index (CI) of approximately 1, indicating<br />

an additive effect. Addition of CX-4945 to both Gem and Cis<br />

together (1/4 IC50s) decreased viability of HuCCT1 cells by<br />

10-30%, also with a CI around 1. Conclusion: CX-4945 has an<br />

anti-proliferative effect on CCA cell lines. Only higher doses of<br />

CX-4945 increased Caspase 3/7 activity. The anti-proliferative<br />

effect of CX-4945 correlates with effects on PI3K/Akt signaling.<br />

CX-4945 enhanced the anti-proliferative effect of both Cis and<br />

Gem in vitro. An animal study is being performed to confirm<br />

these findings in vivo.<br />

Disclosures:<br />

Sean O’Brien - Employment: Senhwa Biosciences<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Kais Zakharia, Katsuyuki Miyabe,<br />

Catherine D. Moser, Mitesh J. Borad<br />

397<br />

Aspirin Use is Associated with Reduced Risk of Cholangiocarcinoma<br />

Jonggi Choi 1 , Hassan M. Ghoz 1 , Thoetchai Peeraphatdit 1,2 , Esha<br />

Baichoo 1,3 , Benyam D. Addissie 1 , William S. Harmsen 4 , Terry M.<br />

Therneau 4 , Janet E. Olson 4,5 , Roongruedee Chaiteerakij 1,6 , Lewis<br />

R. Roberts 1 ; 1 Division of Gastroenterology and Hepatology, Mayo<br />

Clinic College of Medicine, and Mayo Clinic Cancer Center, Rochester,<br />

MN; 2 Department of Internal Medicine, University of Minnesota,<br />

Minneapolis, MN; 3 Department of Medicine, Mount Sinai<br />

Roosevelt Hospital, New York, NY; 4 Department of Biomedical<br />

Statistics and Informatics, Mayo Clinic, Rochester, MN; 5 Division<br />

of Epidemiology, Department of Health Sciences Research, Mayo<br />

Clinic, Rochester, MN; 6 Department of Medicine, Faculty of Medicine,<br />

Chulalongkorn University and King Chulalongkorn Memorial<br />

Hospital, Thai Red Cross Society, Bangkok, Thailand<br />

Background & Aims: It is unclear whether aspirin reduces cholangiocarcinoma<br />

(CCA) risk. Previous <strong>studies</strong> on risk factors<br />

for CCA combined all the three CCA subtypes (intrahepatic,<br />

perihilar, and distal) together or combined perihilar and distal<br />

CCA as extrahepatic CCA. Whether these factors conferred<br />

the risk of all 3 subtypes to a similar magnitude is unknown.<br />

We aimed to determine the associations between aspirin use<br />

and other risk factors for each CCA subtype individually. Methods:<br />

We enrolled 2,395 CCA cases (1,169 intrahepatic, 995<br />

perihilar and 231 distal) seen at Mayo Clinic, Rochester, MN<br />

between 2000 and 2014. Controls selected from the Mayo<br />

Clinic Biobank were matched 2:1 with cases by age, sex,<br />

race, and residential area (n=4,769). Associations between<br />

aspirin use and other factors and CCA risk were determined<br />

using logistic regression analysis. Results: There were 591<br />

(24.7%) CCA cases and 2,129 (44.6%) controls who took<br />

aspirin. Of aspirin users, 560 (94.8%) of cases and 2008<br />

(94.3%) of controls were daily users; 384 (65.0%) of cases<br />

and 1,697 (79.7%) of controls took low-dose aspirin (81-162<br />

mg/day). Aspirin use had a significant inverse association<br />

for all CCA subtypes with adjusted odds ratios (AOR) (95%<br />

confidence interval, 95% CI) of 0.35 (0.29-0.42), 0.34 (0.27-<br />

0.42), and 0.29 (0.19-0.44), for intrahepatic, perihilar and<br />

distal CCA, respectively, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 411A<br />

Disclosures:<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Jonggi Choi, Hassan M. Ghoz,<br />

Thoetchai Peeraphatdit, Esha Baichoo, Benyam D. Addissie, William S. Harmsen,<br />

Terry M. Therneau, Janet E. Olson, Roongruedee Chaiteerakij<br />

398<br />

The autophagy marker LC3 is significantly associated<br />

with overall survival in human hepatocellular carcinoma<br />

underwent resection<br />

Chih-Wen Lin 1,3 , Jhy-Shrian Huang 1 , Chia-Yen Dai 2,3 , Jee-Fu<br />

Huang 2,3 , Wan-Long Chuang 2,3 , Ming-Lung Yu 2,3 ; 1 Division of<br />

Gastroenterology and Hepatology, Department of Medicine, E-DA<br />

Hospital/ I-SHOU University, Kaohsiung, Taiwan; 2 Hepatobiliary<br />

Division, Department of Internal Medicine, Kaohsiung Medical<br />

University Hospital, Kaohsiung Medical University,, Kaohsiung,<br />

Taiwan; 3 Graduate Institute of Medicine, College of Medicine,<br />

Kaohsiung Medical University, Kaohsiung, Taiwan<br />

BACKGROUND: Hepatocellular carcinoma (HCC) is one of<br />

the most common malignancies, with an increasing incidence<br />

and is the third leading cause of cancer-related mortality in<br />

the world. Most HCC patients are diagnosed at an advanced<br />

stage and have very poor prognosis and low survival. Despite<br />

the improvements of therapeutic modalities in HCC, the rate of<br />

5-years tumor-related death has remained as high as at 30-50<br />

%. Therefore, identification of prognostic factors to develop<br />

newer targeted therapy is urgently needed to improve HCC<br />

patient survival. Autophagy is a process through which longlived<br />

proteins and damaged organelles are conveyed to the<br />

lysosome for removal by degradation and recycling. Some<br />

<strong>studies</strong> have shown that autophagy plays a key role in the<br />

progression and development of cancer. But, the role of autophagy<br />

in the prognosis and metastasis of human HCC is not<br />

well unknown. AIMS: The study aims to explore the expression<br />

of markers of autophagy genes using immunohistochemistry<br />

(IHC) in human HCC tissues. We also investigate the autophagy<br />

markers associated with clinicopathological characteristics<br />

and prognosis. METHODS: We retrospectively analyzed<br />

200 patients diagnosed with HCC by histology after surgical<br />

resection at the Chunghua Christian Hospital, Chunghua, and<br />

Kaohsiung Medical University Hospital, Kaohsiung, Taiwan,<br />

from 2009 to 2014. The demographic data, recurrence, and<br />

survival were collected until December 2014. The expression<br />

of autophagy-related markers (LC3 and p62) were analyzed<br />

by IHC staining using HCC tumor tissues and non-tumor tissues.<br />

RESULTS: Two hundreds HCC patients were collected.<br />

The average age is 62.8 years old and the rate of male is<br />

74%. The rate of HBV, HCV, HBV+HCV, and Non-HBVHCV is<br />

48%, 28%, 4%, and 20%, respectively. Median survival was<br />

26.3 months (range 3-42 months). The positive rate of LC3 was<br />

significantly higher in HCC tumor tissues than non-tumor tissues<br />

(85.5 vs. 50.0%, P < 0.001). In HCC, The overall survival was<br />

significantly correlated with cirrhosis background, TMN stage,<br />

BCLC stage, Child-Pugh class, and LC3 tumor part staining in<br />

univariate Cox regression analysis. Furthermore, the overall<br />

survival was significantly correlated with cirrhosis background<br />

(P = 0.01), TMN stage, BCLC stage (P = 0.49), and LC3 tumor<br />

part staining (P = 0.015) in multivariate Cox regression analysis.<br />

The strong positive of LC3 staining is significantly associated<br />

with increasing the 3-year survival rates by Kaplan-Meier<br />

survival analysis (P=0.046). CONCLUSIONS: Our results show<br />

that the expression of autophagy marker, LC3, might be a<br />

strong prognostic factor of overall survival in HCC patients<br />

underwent liver resection.<br />

Disclosures:<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

The following authors have nothing to disclose: Chih-Wen Lin, Jhy-Shrian Huang,<br />

Chia-Yen Dai, Jee-Fu Huang, Ming-Lung Yu<br />

399<br />

Conversion therapy with hepatic arterial infusion<br />

chemotherapy and hepatic resection prolongs overall<br />

survival of patients with advanced hepatocellular carcinoma<br />

involving vascular invasion<br />

Hiroaki Nagamatsu 1,2 , Takuji Torimura 2 ; 1 Department of Gastroenterology<br />

and Hepatology, Yame general hospital, Yame, Japan;<br />

2 Division of Gastroenterology, Kurume University School of Medicine,<br />

Kurume, Japan<br />

[Objectives] Sorafenib is recognized as a standard treatment<br />

for advanced unresectable hepatocellular carcinoma (HCC)<br />

with vascular invasion worldwide. However, the therapeutic<br />

efficacy is marginal. We retrospectively evaluated the therapeutic<br />

effectiveness of conversion therapy with hepatic arterial<br />

infusion chemotherapy (HAIC) . [Subjects] From January 2004<br />

to December 2014, among 270 patients who received HAIC<br />

with cisplatin (CDDP) and 5-fluorouracil (5-FU) at our institution<br />

and affiliated Kurume university hospital, 189 patients with<br />

unresectable HCC involving vascular invasion without extrahepatic<br />

metastasis were enrolled in the study (mean age, 68.9<br />

years; Child-Pugh class A/B/C, 114/66/9 cases; portal vein<br />

tumor thrombosis in 2nd branch/1st branch/trunk, 80/66/43<br />

cases ). [Methods] All 189 patients received HAIC via the<br />

reservoir system. After chemolipiodolization, the patients were<br />

treated with CDDP 50 mg followed by 5-FU 1,250 mg via a<br />

balloon pump for 5 days (NFP therapy 1) ). For patients who<br />

achieved partial response (PR) after NFP and in whom conversion<br />

therapy was feasible, hepatectomy (Hr), trancecatheter<br />

arterial chemoembolization (TACE), or radiotherapy (RT) was<br />

performed to achieve cancer-free outcome. Tumor response<br />

by NFP was evaluated with m-RECIST. Cumulative overall survival<br />

(OS) after NFP was estimated using Kaplan-Meier survival<br />

curves, and multivariate analysis was performed to identify<br />

prognostic factors affecting OS using Cox proportional hazards<br />

models. For factors with significant difference in the multivariate<br />

analysis, OS was compared using log-rank test. [Results] 136<br />

patients (72%) responded to NFP therapy (CR, 17; PR,119).<br />

41 patients who showed PR received conversion therapy,(<br />

NFP +Hr, 18 patients; NFP +TACE, 18 patients, NFP +RT, 5<br />

patients). 58 patients (31%) achieved cancer-free outcome.<br />

Median OS (MST) after HAIC in all patients was 18 months.<br />

Significant factors for OS included cancer-free outcome (hazard<br />

ratio, 0.184; P=0.001) and addition of Hr (hazard ratio,<br />

0.105; P=0.003). MST in patients who responded was 29<br />

months, and in patients who achieved cancer-free outcome, it<br />

was extended to 57 months.Three - and five-year survival rates<br />

were 78% and 21% in patients treated with NFP only, 87%<br />

and 71% in NFP+Hr, 47% and 16% in NFP+TACE, and 67%<br />

and 33% in NFP+RT patients (P=0.028), respectively. [Conclusions]<br />

NFP therapy demonstrated favorable tumor response in<br />

patients with advanced hepatocellular carcinoma with vascular<br />

invasion. Addition of conversion therapy, specifically, Hr, targeting<br />

cancer-free outcome made possible long-term survival.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hiroaki Nagamatsu, Takuji Torimura


412A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

400<br />

Model to predict recurrence after living donor liver<br />

transplantation for hepatocellular carcinoma beyond<br />

the Milan criteria<br />

Yuri Cho 5 , Jeong-Hoon Lee 5 , Eun Ju Cho 5 , Donghyeon Lee 5 ,<br />

Jeong-Ju Yoo 5 , Minjong Lee 5 , Young Youn Cho 5 , Su Jong Yu 5 ,<br />

Nam-Joon Yi 4 , Kwang-Woong Lee 4 , Seong Hoon Kim 1 , Jong Man<br />

Kim 2 , Jae Won Joh 2 , June Sung Lee 3 , Yoon Jun Kim 5 , Kyung-Suk<br />

Suh 4 , Jung-Hwan Yoon 5 ; 1 Center for Liver Cancer, National Cancer<br />

Center, Goyang-Si, Korea (the Republic of); 2 Department of<br />

Surgery, Samsung Medical Center, Seoul, Korea (the Republic of);<br />

3 Department of Internal Medicine, Inje University, Ilsan Paik Hospital,<br />

Goyang-Si, Korea (the Republic of); 4 Department of Surgery,<br />

Seoul National University College of Medicine, Seoul, Korea (the<br />

Republic of); 5 Department of Internal Medicine and Liver Research<br />

Institute, Seoul National University College of Medicine, Seoul,<br />

Korea (the Republic of)<br />

Background: Some subgroups of hepatocellular carcinoma<br />

(HCC) exceeding the Milan criteria (MC) experience substantial<br />

benefit from living donor liver transplantation (LDLT). This<br />

study aimed to develop and validate a model to predict tumor<br />

recurrence after LDLT (MoRAL) for HCC beyond the MC. Methods:<br />

This multicenter study included a total of 566 consecutive<br />

patients who underwent LDLT in Korea: the beyond-MC cohort<br />

(n=205, the derivation [n=92] and validation [n=113] sets)<br />

and the within-MC cohort (n=361). Results: Using multivariate<br />

Cox proportional hazard model, we derived the MoRAL score<br />

(11 x √protein induced by vitamin K absence-II + 2 x √alpha-fetoprotein),<br />

which provided a good discriminant function on<br />

time-to-recurrence (c-index=0.88). C-index was maintained similarly<br />

on both internal and external validations (mean 0.87 and<br />

0.84, respectively). At cut-off of 314.8 (75th percentile value),<br />

a low MoRAL score (≤314.8) was associated with significantly<br />

longer recurrence-free (vs. >314.8, hazard ratio [HR]=5.29,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 413A<br />

right upper quadrant pain (24%), and hepatomegaly (24%).<br />

Routine screening detected carcinoma in 9/25 patients; 8/9<br />

of those screened cases were caught at an early stage with a<br />

single tumor. Prior to diagnosis with HCC, the median interval<br />

since prior imaging was 12 (1-42) months. Patients without<br />

prior or recent (within 12 months) liver imaging were more<br />

likely to present with multiple liver masses or metastatic disease<br />

(91%). Moreover, 88% of patients deceased before 1 year<br />

after diagnosis had no prior or recent screening. Conclusion:<br />

Patients with Fontan circulation are at increased risk to develop<br />

late hepatocellular cancer and should be screened according<br />

to guidelines. Early diagnosis of hepatocellular carcinoma may<br />

allow for improved treatment strategies and prolonged survival.<br />

Presenting Stages of Hepatocellular Carcinoma<br />

for Epi-TACE group and 869 days for Cis-TACE group (HR:<br />

1.18, 95% CI:0.57-2.44, p = 0.658). In subgroup analysis for<br />

the primary treatment, response rate was significantly better in<br />

Epi-TACE group than in Cis-TACE group (64.3% versus 8.3%,<br />

respectively, p = 0.002), though they were comparable in<br />

subgroup analysis for the non-primary treatment (16.7% versus<br />

26.7%, respectively, p = 0.531). Conclusions This prospective<br />

randomized controlled study demonstrated that Epi-TACE might<br />

be more effective than Cis-TACE. Epirubicin might be used<br />

for the primary TACE, and cisplatin might be alternatives as<br />

non-primary TACE.<br />

Disclosures:<br />

Etsuro Hatano - Speaking and Teaching: Bayer<br />

The following authors have nothing to disclose: Yosuke Kasai, Kohta Iguchi,<br />

Rinpei Imamine, Takahiro Nishio, Masayuki Okuno, Satoru Seo, Kojiro Taura,<br />

Kentaro Yasuchika, Toshiya Shibata, Shinji Uemoto<br />

Disclosures:<br />

Daniel Ganger - Advisory Committees or Review Panels: Bristol Myers, Abbvie;<br />

Grant/Research Support: Merck, Ocera; Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Laura J. Dickmeyer, Barbara J.<br />

Deal, Andrew Defreitas<br />

402<br />

A randomized controlled study of effectiveness between<br />

transcatheter arterial chemoembolization with epirubicin<br />

versus cisplatin for multiple hepatocellular carcinoma<br />

Yosuke Kasai 1 , Etsuro Hatano 1 , Kohta Iguchi 1 , Rinpei Imamine 2 ,<br />

Takahiro Nishio 1 , Masayuki Okuno 1 , Satoru Seo 1 , Kojiro Taura 1 ,<br />

Kentaro Yasuchika 1 , Toshiya Shibata 2 , Shinji Uemoto 1 ; 1 Department<br />

of Surgery, Graduate School of Medicine, Kyoto University,<br />

Kyoto, Japan; 2 Department of Diagnostic Imaging and Nuclear<br />

Medicine, Graduate School of Medicine, Kyoto University, Kyoto,<br />

Japan<br />

Background and Aim In transcatheter arterial chemoembolization<br />

(TACE) for hepatocellular carcinoma (HCC), it has not been<br />

clarified which chemotherapy agent is an effective combined<br />

drug with lipiodol. We previously showed that lipiodol-TACE<br />

combined with cisplatin (Cis-TACE) provided a higher response<br />

rate than lipiodol-TACE combined with epirubicin (Epi-TACE)<br />

for multiple HCC in a retrospective cohort study (Hepatol Res.<br />

2011). We conducted a randomized controlled study comparing<br />

the effectiveness of Epi-TACE and Cis-TACE. Method<br />

Sixty patients with multiple HCC which were not indicated for<br />

curative liver resection or percutaneous ablation therapy were<br />

randomly assigned to Epi-TACE or Cis-TACE group, stratified<br />

by the primary or non-primary treatment and the presence or<br />

absence of the history of liver resection. Finally, 26 patients for<br />

Epi-TACE group and 27 patients for Cis-TACE group met the<br />

eligibility criteria and were analyzed. The primary endpoint<br />

was response rate, and the secondary endpoints were progression<br />

free survival (PFS) and overall survival (OS). Results<br />

Patient characteristics were almost comparable between the<br />

two groups, except for larger tumor size in Epi-TACE group.<br />

The response rates were 42.3% for Epi-TACE group and<br />

18.5% for Cis-TACE group (p = 0.057). When adjusted by<br />

tumor size, the odds ratio for Epi-TACE group against Cis-TACE<br />

group was 3.47 [95% confidence interval (CI): 0.98-13.6, p =<br />

0.054]. The median PFS times were 224.5 days for Epi-TACE<br />

group and 165 days for Cis-TACE group [hazard ratio (HR) for<br />

Epi-TACE group against Cis-TACE group: 0.83, 95% CI: 0.47-<br />

1.45, p = 0.503], and the median OS times were 1128 days<br />

403<br />

Overexpression of periostin is related to poor prognosis<br />

of hepatocellular carcinoma<br />

Se Young Jang 1 , Soo Young Park 1 , Won Young Tak 1 , Young Oh<br />

Kweon 1 , Keun Hur 2 , Gyeonghwa Kim 2 , Yong-Hun Choi 2 , Jung Gil<br />

Park 3 , Yu Rim Lee 1 , Sun Kyung Jang 1 , Su Hyun Lee 1 ; 1 Kyungpook<br />

National University Hospital, Daegu, Korea (the Republic of); 2 Biochemistry<br />

and Cell Biology, Kyungpook National University School<br />

of Medicine, Daegu, Korea (the Republic of); 3 Gastroenterology<br />

and Hepatology, CHA University, Gumi CHA medical center,<br />

Gumi, Korea (the Republic of)<br />

Background: Periostin is an extracellular matrix protein and<br />

known to be related to the metastatic potential and prognosis<br />

of human cancers in recent <strong>studies</strong>. However, few <strong>studies</strong><br />

showed the expression level of periostin and its association<br />

with prognosis of hepatocellular carcinoma (HCC). We analyzed<br />

periostin overexpression and its implication for prognosis<br />

of HCC. Methods: A total of 149 patients who underwent<br />

surgical resection in Kyungpook National University Hospital<br />

between 2006 and 2010 were selected and retrospectively<br />

reviewed. All tissue specimens were formalin-fixed and paraffin-embedded<br />

for immunohistochemistry staining. A tissue<br />

microarray (TMA) containing tissue from HCC (n=149), adjacent<br />

non-tumor tissues of HCC. Results: Expression of periostin<br />

was associated with microvascular invasion (OR=3.383;<br />

p=0.001), multiple tumors (OR=2.631; p=0.020), and higher<br />

modified UICC stage (OR=2.945, p=0.004). The patients with<br />

positive periostin expression have significantly shorter overall<br />

survival (p=0.021) and recurrence-free survival (p=0.032).<br />

Cumulative overall survival was significantly different between<br />

periostin negative and positive groups when microvascular<br />

invasion was absent (p=0.018). Conclusion: We found that<br />

high periostin expression on hepatocellular carcinoma tissues is<br />

correlated with poor prognosis. Especially when microvascular<br />

invasion is absent, periostin can be a better prognostic marker<br />

for overall survival.


414A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Cumulative overall survival rates of periostin positive and negative<br />

gruops according to the existence of microvascular invasion.<br />

Disclosures:<br />

Won Young Tak - Advisory Committees or Review Panels: Gilead Korea; Grant/<br />

Research Support: SAMIL Pharma; Speaking and Teaching: Bayer Korea<br />

The following authors have nothing to disclose: Se Young Jang, Soo Young Park,<br />

Young Oh Kweon, Keun Hur, Gyeonghwa Kim, Yong-Hun Choi, Jung Gil Park,<br />

Yu Rim Lee, Sun Kyung Jang, Su Hyun Lee<br />

404<br />

Impact of screening on Hepatocellular Carcinoma survival<br />

in patients with Chronic Hepatitis B - A regional<br />

population based study<br />

Rasham Mittal, Bechien U. Wu; Kaiser Permanente Los Angeles<br />

Medical Center, Los Angeles, CA<br />

Background: The impact of screening for hepatocellular carcinoma<br />

(HCC) in patients with chronic hepatitis B (CHB) remains<br />

controversial. Aim: To determine if screening is associated with<br />

improved survival among CHB patients with HCC in a diverse<br />

regional U.S.population. Methods: Retrospective cohort study<br />

(2006-2013) from an integrated health care system in Southern<br />

California. All CHB (age ≥18 years) patients identified<br />

using ICD codes and confirmed by serology. HCC patients<br />

were identified using ICD code and cross referenced through<br />

prospective internal cancer registry. Patients were considered<br />

to have undergone screening if they had USG, CT scan or MRI<br />

abdomen with contrast imaging at 6-18 months prior to HCC<br />

diagnosis. Patients were excluded if they did not have active<br />

Kaiser Membership for at least 18 months prior to HCC diagnosis.<br />

Survival curves estimated using Kaplan-Meyer analysis.<br />

Cox proportional hazard regression analysis was performedt o<br />

adjust for baseline characteristics. Results: From a total of 9811<br />

CHB patients, we identified 185 HCC cases [median age at<br />

HCC diagnosis 61 years, IQR 53, 68 years; 147 (79.4%)<br />

male; 136 (73.5%) Asians; 123 (66.5%) liver cirrhosis]. Out<br />

of 185, 110 patients had screening prior to HCC while 75 presented<br />

with HCC without prior screening. The median observed<br />

survival was higher in screening versus unscreened group (5.9<br />

years versus 3.2 years; P=0.036). In regression analysis, HCC<br />

screening (HR 0.60; 95%CI, 0.37-0.96; P=0.03) were associated<br />

with improved survival after adjusting for baseline characteristics<br />

such as age (HR 0.97; 95%CI, 0.95-0.99; P=0.259),<br />

female gender (HR 0.40; 95%CI, 0.20-0.85; P=0.012) Charlson<br />

co-morbidity score (HR 1.14; 95%CI, 1.04-1.24; P=0.003)<br />

and liver cirrhosis (HR 1.14; 95%CI, 0.66-1.98; P=0.623).<br />

Conclusion: HCC surveillance was associated with improved<br />

survival in CHB patients in a large regional U.S. population.<br />

Disclosures:<br />

The following authors have nothing to disclose: Rasham Mittal, Bechien U. Wu<br />

405<br />

Influence of hepatitis B virus DNA elevation on recurrence<br />

of hepatocellular carcinoma after surgical resection<br />

and the preventive role of antiviral therapy<br />

Yang J. Yoo, Ji Hoon Kim, Seong Hee Kang, Young-Sun Lee, Tae<br />

Suk Kim, Sang Jun Suh, Young Kul Jung, Yeon Seok Seo, Hyung<br />

Joon Yim, Jong Eun Yeon, Kwan Soo Byun; Department of Internal<br />

Medicine, Korea University College of Medicine, Seoul, Korea (the<br />

Republic of)<br />

Aims Surgical resection is the treatment of choice for early<br />

stage hepatocellular carcinoma (HCC). Previous <strong>studies</strong><br />

revealed that reactivation of hepatitis B virus is associated with<br />

the recurrence of hepatitis B virus (HBV) related HCC after surgical<br />

resection. We aimed to investigate the influence of HBV<br />

DNA elevation on HCC recurrence and the preventive role of<br />

antiviral therapy. Methods One hundred eleven patients who<br />

had BCLC stage 0 or A and received surgical resection as primary<br />

therapy were enrolled. HBV DNA elevation was defined<br />

as reactivation (increase >1log 10<br />

IU/ml or re-emergence of<br />

HBV DNA) in patients without preoperative antiviral therapy or<br />

virologic breakthrough in patients with preoperative antiviral<br />

therapy (n=62). Results Mean age was 54.1±9.4 years and<br />

male consisted 76.6% (n=85). All patients belonged in Child<br />

class A. BCLC 0 consisted 29.7% (n=33) and 30.6% (n=34)<br />

had Hepatitis B envelope antigen positivity and 42.3% (n=<br />

47) had HBV DNA level ≥ 2000 IU/mL at the time of surgery.<br />

Overall 1 year, 3 year, 5 year recurrence was 17.3%, 36.4%,<br />

40%. In multivariate analysis for risk factors of recurrence,<br />

multiple tumor, HBV DNA elevation, ALT ≥ 30 IU/L, were independently<br />

associated with HCC recurrence. In subgroup analysis,<br />

patients with preoperative antiviral therapy showed similar<br />

HCC recurrence rate with patients who start antiviral therapy<br />

after resection and significantly higher recurrence rate than<br />

patients with no antivirals during follow up. In multivariate analysis<br />

for risk factors of HBV DNA elevation, Age < 50 years,<br />

no preoperative antiviral therapy, HBV DNA ≥ 2000 remained<br />

as risk factors. To rule out the antiviral effect, we investigated<br />

risk factors for HCC recurrence in each group with or without<br />

preoperative antiviral therapy. In patients with preoperative<br />

antiviral therapy, multiple tumor was the only risk factor for<br />

HCC recurrence. In patients without preoperative antiviral therapy,<br />

male sex and HBV DNA elevation were independent risk<br />

factors for HCC recurrence. Risk factors for HBV DNA elevation<br />

were further analyzed in patients without preoperative antiviral


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 415A<br />

therapy. Age < 50 years and HBV DNA ≥ 2000 IU/mL were<br />

independent risk factors for HBV DNA elevation. Conclusion<br />

HBV DNA elevation after resection increases the risk of HCC<br />

recurrence irrespective of preoperative antiviral therapy. However,<br />

HBV DNA elevation is an independent risk factor for<br />

recurrence in patients without preoperative antiviral therapy.<br />

Therefore, perioperative antiviral therapy should be considered<br />

to prevent HBV DNA elevation and recurrence, especially in<br />

patients with Age < 50 years and/or HBV DNA ≥ 2000 IU/<br />

mL.<br />

Disclosures:<br />

Kwan Soo Byun - Grant/Research Support: Gilead, BMS<br />

The following authors have nothing to disclose: Yang J. Yoo, Ji Hoon Kim, Seong<br />

Hee Kang, Young-Sun Lee, Tae Suk Kim, Sang Jun Suh, Young Kul Jung, Yeon<br />

Seok Seo, Hyung Joon Yim, Jong Eun Yeon<br />

survival (P = 0.024). Conclusions: Gene alterations in TERT<br />

promoter, TP53, CTNNB1, and HBV integration were closely<br />

associated with HCC development, and mutations in TERT promoter<br />

are related to poor prognosis. These results are useful for<br />

understanding the underlying mechanism of hepatocarcinogenesis,<br />

diagnosis, and predicting outcomes of patients with HCC.<br />

Disclosures:<br />

Sei Kakinuma - Grant/Research Support: The Japanese Society of Gastroenterology,<br />

MSD Co., Ltd.<br />

Mamoru Watanabe - Grant/Research Support: MSD Co., Ltd.<br />

The following authors have nothing to disclose: Fukiko Kawai-Kitahata, Yasuhiro<br />

Asahina, Shinji Tanaka, Miyako Murakawa, Sayuri Nitta, Takako Watanabe,<br />

Satoshi Otani, Fumio Goto, Hiroko Nagata, Shun Kaneko, Seishin Azuma,<br />

Yasuhiro Itsui, Mina Nakagawa, Minoru Tanabe, Shinya Maekawa, Nobuyuki<br />

Enomoto<br />

406<br />

Comprehensive analyses of mutations and hepatitis B<br />

virus integration in hepatocellular carcinoma with clinicopathological<br />

features<br />

Fukiko Kawai-Kitahata 1 , Yasuhiro Asahina 1 , Shinji Tanaka 1 , Sei<br />

Kakinuma 1 , Miyako Murakawa 1 , Sayuri Nitta 1 , Takako Watanabe<br />

1 , Satoshi Otani 1 , Fumio Goto 1 , Hiroko Nagata 1 , Shun<br />

Kaneko 1 , Seishin Azuma 1 , Yasuhiro Itsui 1 , Mina Nakagawa 1 ,<br />

Minoru Tanabe 1 , Shinya Maekawa 2 , Nobuyuki Enomoto 2 ,<br />

Mamoru Watanabe 1 ; 1 Tokyo Medical and Dental University,<br />

Tokyo, Japan; 2 University of Yamanashi, Yamanashi, Japan<br />

Background & Aims: Genetic alterations in specific genes<br />

are critical events in carcinogenesis and hepatocellular carcinoma<br />

(HCC) progression. However, the genetic alterations<br />

responsible for HCC development, progression, and survival<br />

are unclear. Methods: The extracted DNA from 104 patients<br />

was amplified by multiplexed PCR targeting 54 cancer-related<br />

genes containing 2820 mutational hotspots. The mutations<br />

were analyzed using a semiconductor-based DNA sequencer.<br />

Nucleotide variants that were detected in tumors and absent<br />

in corresponding non-tumor tissue were determined as somatic<br />

mutations. We also investigated TERT promoter mutation at<br />

2 hot spots through direct sequencing. HBV oligonucleotide<br />

probes were designed primarily within the conserved regions<br />

of different HBV strains including genotypes Aa, Ae, Ba, Bj,<br />

and C. After incubation with custom capture oligos, hybridized<br />

fragments were sequenced on a MiSeq instrument generating<br />

300 bp paired-end reads. Results: We found recurrent<br />

mutations in several genes such as TERT (65%), TP53 (38%),<br />

CTNNB1 (30%), AXIN1 (2%), PTEN (2%), and CDKN2A<br />

(2%). TERT promoter mutations were associated with older<br />

age (P = 0.005), presence of hepatitis C virus (HCV) infection<br />

(P = 0.003), and absence of hepatitis B virus (HBV) infection<br />

(P < 0.0001). In patients with hepatitis B surface antigen<br />

(HBsAg) positive, a total of 78 HBV integration breakpoints<br />

were detected in 25/27 (92.6%) patients. Most (89%) of HBV<br />

integrant was the HBx region. HBV integration was frequently<br />

found in particular genes including TERT (n = 9), KMT2B (MLL4;<br />

n = 2), and MYO7A (n = 2). Seven (77.8%) of TERT locus<br />

breakpoints were paired with HBx gene, all at the 3’-end of<br />

HBx gene. HBV integration into TERT locus was mutually exclusive<br />

to TERT promoter mutations. The HBV integration number<br />

was significantly associated with HCC at younger age (Spearman’s<br />

rank order correlation, r s<br />

= −0.5342, P < 0.01). TP53<br />

mutations were associated with HBV infection (P = 0.0001)<br />

and absence of HCV infection (P = 0.002). CTNNB1 mutations<br />

were associated with absence of HBV infection (P = 0.010).<br />

Moreover, TERT promoter mutation was significantly associated<br />

with shorter disease-free survival (P = 0.005) and poor overall<br />

407<br />

Treatment not needed for hypovascular tumors associated<br />

with hepatocellular carcinoma<br />

Yutaka Midorikawa, Tadatoshi Takayama; Nihon Univ., Tokyo,<br />

Japan<br />

Aim: Hypovascular tumors associated with hepatocellular<br />

carcinoma (HCC) can be diagnosed, but it remains unknown<br />

whether such lesions should be treated immediately after diagnosis.<br />

This study aimed to clarify whether treating hypovascular<br />

liver nodules contributes to the prolongation of patient survival,<br />

and sought to directly calculate a lead time from the diagnosis<br />

of hypovascular liver nodules to their transformation into<br />

hypervascular HCC. Methods: One hundred and four patients<br />

with hypovascular liver nodules smaller than 3 cm underwent<br />

liver resection immediately after diagnosis (Group 1), while<br />

80 patients with hypovascular liver nodules were treated after<br />

appearance of vascularization or other hypervascular HCC<br />

lesions on imaging <strong>studies</strong> (Group 2). To avoid lead-time bias<br />

for tumor vascularization, the survival rates of patients after<br />

diagnosis of hypovascular liver nodules as well as the survival<br />

rates after liver resection in the two groups were compared.<br />

Results: After a median follow-up period of 3.3 years (range,<br />

0.6–11.2), the median overall survival rates “after liver resection”<br />

were 9.7 years (95% confidence interval [CI], 6.7–11.9)<br />

and 8.5 years (4.8–12.1; P = 0.017), respectively, and recurrence-free<br />

survivals were 3.8 years (95% CI, 2.7–6.3) and 1.9<br />

years (1.5–2.9; P


416A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Yutaka Midorikawa, Tadatoshi<br />

Takayama<br />

408<br />

Elevated Serum Wisteria floribunda agglutinin-positive<br />

human Mac-2 binding protein predict all-cause mortality<br />

in hepatitis C patients with hepatocellular carcinoma<br />

Shigemune Bekki, Kazumi Yamasaki, Yuuki Sonoda, Yuki Kugiyama,<br />

Satoru Hashimoto, Akira Saeki, Shinya Nagaoka, Seigo<br />

Abiru, Atsumasa Komori, Hiroshi Yatsuhashi; National Organization<br />

Hospital Nagasaki Medical Center, Omura, Japan<br />

Background & Aims: Serum glycosylated Wisteria floribunda<br />

agglutinin-positive Mac-2 binding protein (WFA(+)-M2BP)<br />

has been recently reported as a noninvasive and useful surrogate<br />

marker for the degree of liver fibrosis and for the risk of<br />

hepatocellular carcinoma (HCC). We investigated the utility<br />

of WFA(+)-M2BP to predict all-cause mortality in the patients<br />

with hepatitis C virus (HCV)-related hepatocellular carcinoma<br />

(HCC). Patients & Methods: A total of 466 HCV-related HCC<br />

patients who had been admitted to our hospital from 1999<br />

to 2013 were enrolled. We evaluated association between<br />

mortality and clinicolaboratory variables, included age, sex,<br />

asparate and alanine aminotransferase levels (AST and ALT),<br />

bilirubin, albumin, prothrombin activity, platelet count, Child-<br />

Pugh grade, alpha-fetoprotein (AFP), des-γ-carboxy prothrombin<br />

(DCP), TNM stage, and WFA(+)-M2BP. The average<br />

observation period was 3.7 years. The association between<br />

WFA(+)-M2BP and mortality was evaluated using the Cox proportional<br />

hazards regression analysis. Results: The baseline<br />

characteristics of the 466 patients were 295 (63.2%) male with<br />

a median age of 70.0 years. The median value of WFA(+)-<br />

M2BP was 4.6 (range, 0.2-18.36) cutoff index (COI). Serum<br />

WFA(+)-M2BP levels significantly correlated with bilirubin<br />

(r=0.26, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 417A<br />

410<br />

Significant Suppression of Local Tumor Recurrence in<br />

Treatment for Hepatocellular Carcinoma with Peritumoral<br />

Vessels by Radiofrequency Ablation with Percutaneous<br />

Ethanol Injection: Propensity Score Matching<br />

Analysis<br />

Kazuhide Takata 1 , Akira Anan 1 , Takashi Tanaka 1 , Keiji<br />

Yokoyama 1 , Daisuke Morihara 1 , Yasuaki Takeyama 1 , Makoto<br />

Irie 1 , Satoshi Shakado 1 , Kaoru Iwata 2 , Tetsuro Sohda 1 , Shotaro<br />

Sakisaka 1 ; 1 Gastroenterology and Medicine, Fukuoka University<br />

Faculty of Medicine, Fukuoka, Japan; 2 Internal medicine,<br />

Meotoiwa Hospital, Fukuoka, Japan<br />

Background and Aims: Radiofrequency ablation (RFA) is a<br />

major modality for the treatment of early stage hepatocellular<br />

carcinoma (HCC). However, serious problems of RFA remain<br />

in local recurrence of HCC with peritumoral vessels, probably<br />

due to heat-sink effect. Therefore, we have undergone percutaneous<br />

ethanol injection (PEI) to the peritumoral vessels side of<br />

the tumor to complete tumor necrosis by ethanol, with RFA. The<br />

aim of the study was to evaluate the effect PEI combination RFA<br />

for HCC with peritumoral vessels. Methods: From Jun 2001 to<br />

Mar 2015, 371 patients, who diagnosed newly HCC (total<br />

525 nodules), 3 cm or less and within 3 nodules, were treated<br />

RFA as an initial treatment in our hospital. Among 190 nodules<br />

with peritumoral vessels located less than 5 mm from the tumor,<br />

patients in group of RFA monotherapy (61 nodules) or group<br />

of PEI combination (61 nodules) were 1 : 1 matched according<br />

to their propensity scores. Ethanol was injected inside the<br />

tumor close to the peritumoral vessels in the combination therapy<br />

group. Cumulative incidences of HCC local recurrence<br />

were analyzed. Results: The overall median follow-up period<br />

was 42 months (range, 3-134). The cumulative HCC local<br />

recurrence rates were 7.2%, 18.7%, and 24.1% at 1, 3, 5<br />

years, respectively, in the PEI combination group, and 15.1%,<br />

34.9%, 46.9% in the RFA monotherapy group; showing significantly<br />

lower recurrence rates in PEI combination group (P <<br />

0.05). In Cox regression analysis, presence of PEI combination<br />

was significantly associated with the risk reduction of HCC<br />

local recurrence (hazard ratio, 0.44; 95% CI: 0.21, 0.88; P <<br />

0.05). Conclusions: RFA combined with PEI reduced the risk of<br />

a local recurrence rate by half, compared with RFA alone, in<br />

early stage HCC with peritumoral vessels. Further prospective<br />

<strong>studies</strong> based on cumulative local recurrence rate may be helpful<br />

in improving the quality of RFA for HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kazuhide Takata, Akira Anan,<br />

Takashi Tanaka, Keiji Yokoyama, Daisuke Morihara, Yasuaki Takeyama, Makoto<br />

Irie, Satoshi Shakado, Kaoru Iwata, Tetsuro Sohda, Shotaro Sakisaka<br />

411<br />

Comparison of balloon-occluded Transarterial chemoembolization<br />

(TACE) with conventional TACE for treatment<br />

of hepatocellular carcinoma<br />

Takuro Niinomi, Masatoshi Ishigami, Yoji Ishizu, Teiji Kuzuya,<br />

Takashi Honda, Kazuhiko Hayashi, Yoshiki Hirooka, Hidemi Goto;<br />

Gastroenterology and Hepatology, Nagoya Univercity Graduate<br />

School of Medicine, Nagoya, Japan<br />

AIM: Transarterial chemoembolization (TACE) has been widely<br />

performed as a treatment for hepatocellular carcinoma (HCC)<br />

in patients who don’t have the indication for curative treatment.<br />

Selecting the anti-cancer agents, treatment methods and equipment<br />

for the best therapeutic effect is becoming more important<br />

for performing better quality of TACE. In recent years, balloon-occluded<br />

TACE (B-TACE) by using microballoon catheters<br />

with small diameters has been widely applied, primarily in<br />

Japan. The purpose of this study was to compare the early<br />

therapeutic efficacy and lipiodol deposition of B-TACE with<br />

those of conventional TACE (C-TACE). Patients and Methods:<br />

From April 2013 to November 2014, 200 consecutive patients<br />

with unresectable HCC were treated with TACE at our institution.<br />

The patients were divided into two groups based on the<br />

usage of microballoon catheter. Treatment efficacy and lipiodol<br />

deposition were retrospectively compared between two<br />

groups. Of these patients, 20 patients with 51 nodules were<br />

treated with B-TACE and 39 patients with 99 nodules were<br />

treated with conventional TACE (C-TACE). Mean lesion density<br />

and lipiodol emulsion concentration ratio of HCC to embolized<br />

liver parenchyma (LECHL ratio) in Hounsfield units (HU) were<br />

evaluated on non-enhanced CT imaging immediately after<br />

TACE. The treatment effect of TACE was evaluated by dynamic<br />

CT 1 to 3 months after TACE, and was evaluated based on<br />

the Response Evaluation Criteria in Cancer Study Group of<br />

Japan (RECICL2009). Results: The median lesion density on CT<br />

immediately after TACE was higher in B-TACE group than in<br />

C-TACE group (B-TACE 533 HU; range 121-1117 HU versus<br />

C-TACE 422 HU; range 136-1322 HU, P=0.057, Mann-Whitney<br />

U test). LECHL ratio on CT immediately after TACE was<br />

remarkably higher in B-TACE group than in C-TACE group<br />

(B-TACE 3.69; range 1.29-9.7 versus C-TACE 2.82; range<br />

0.96-12.83, P=0.017, Mann-Whitney U test). The treatment<br />

effect on the target nodule classified as TE4, TE3, TE2 and<br />

TE1 were 68%, 16%, 16% and 0% in the B-TACE group, and<br />

58%, 9%, 28% and 5% in the C-TACE group respectively.<br />

Response rate (TE4+TE3) was significantly higher in B-TACE<br />

group than in C-TACE group (84% versus 67%, P=0.0216,<br />

Fisher’s exact test). Conclusion: Mean lesion density was higher<br />

in B-TACE group than in C-TACE group. Notably, LECHL ratio<br />

which imply selective deposition of lipiodol, were significantly<br />

higher in B-TACE group than in C-TACE group. These results<br />

suggested that compared with C-TACE, B-TACE significantly<br />

improved lipiodol deposition on HCC nodule during TACE<br />

procedure, and could obtain better treatment efficacy.<br />

Disclosures:<br />

Hidemi Goto - Grant/Research Support: MSD, Roche, Bayer, Bristol-Myers, Eisai,<br />

Ajinomoto, Otsuka, Astra, Tanabe, Takeda<br />

The following authors have nothing to disclose: Takuro Niinomi, Masatoshi<br />

Ishigami, Yoji Ishizu, Teiji Kuzuya, Takashi Honda, Kazuhiko Hayashi, Yoshiki<br />

Hirooka<br />

412<br />

Hepatic Decompensation After Transarterial Chemoembolization<br />

For Hepatocellular Carcinoma<br />

Dempsey Hughes, Fredric D. Gordon, Keith E. Stuart, FW Nugent,<br />

Sebastian Flake, Mary Ann Simpson, Amir A. Qamar; Transplantation,<br />

Lahey Hospital and Medical Center, Burlington, MA<br />

Background:Hepatic decompensation(HD)is a concern in<br />

patients with cirrhosis who undergo transarterial chemoembolization(TACE)for<br />

hepatocellular carcinoma.Identifying predictors<br />

associated with treatment-related HD would allow better<br />

risk assessment risk associated with TACE.The aim of this study<br />

is to assess HD after TACE and to identify any pre-treatment<br />

factors related to HD. Methods:A retrospective cohort study<br />

of 184 patients with cirrhosis and HCC who underwent TACE<br />

at our institution from 2009-14 was analyzed.TACE-related<br />

HD was defined as:1)new onset variceal hemorrhage(VH),<br />

ascites or hepatic encephalopathy(HE), 2)treatment change<br />

for known history of VH,ascites (increased diuretic or paracentesis<br />

requirement), HE(increase or addition of lactulose or<br />

rifaximin), or spontaneous bacterial peritonitis within 90 days<br />

of first TACE.TACE was performed using standard protocol.


418A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Univariate analysis compared baseline factors.Baseline factors<br />

independently associated with HD after TACE were identified<br />

with Cox regression. Results:51 patients(28%)experienced an<br />

HD event within 90 days of TACE.Median time to decompensation<br />

event after first TACE was 52 days.On univariate analysis,<br />

alkaline phosphatase(ALP), total bilirubin(TB), albumin,<br />

INR, MELD, being within Milan or UCSF criteria, history of<br />

HD event within 1 year of TACE significantly differed between<br />

decompensated and compensated groups(TABLE).Cox regression<br />

model identified a decompensation event within 1 year<br />

prior to TACE(HR 2.19,p=0.02), ALP(HR 1.003,p=0.007)and<br />

being outside UCSF criteria(HR 0.49,p=0.02)as independently<br />

associated with an HD event after TACE.TB, albumin, INR and<br />

MELD were not. Conclusion:For patients with cirrhosis being<br />

considered for TACE, absence of decompensating event within<br />

1 year of TACE, HCC within UCSF criteria and low ALP are<br />

associated with low risk of TACE-related HD.<br />

Univariate Analysis<br />

Disclosures:<br />

The following authors have nothing to disclose: Dempsey Hughes, Fredric D.<br />

Gordon, Keith E. Stuart, FW Nugent, Sebastian Flake, Mary Ann Simpson, Amir<br />

A. Qamar<br />

between pathological and radiological findings regarding<br />

the number of nodules was found in 103/184(56%): 78<br />

were underestimated and 25 overestimated by radiology.<br />

10/40(25%) patients who were within Milan criteria, resulted<br />

outside according to histology due to discrepancy in number<br />

of nodules. Considering the diameter of the biggest nodule on<br />

explant, discordance was observed in 158/181(87%): 93<br />

were underestimated and 65 overestimated. 20 of 40(50%)<br />

within Milan criteria on imaging, exceeded these due to<br />

discrepancy in biggest nodule diameter. HCC recurrence<br />

occurred in 29/184(16%) patients in a median of 37m (4-157)<br />

post-LT. In Cox regression, pre-LT factors associated with HCC<br />

recurrence were aFP>100 IU/ml (p3cm<br />

(p3cm but different<br />

aFP levels (22% in aFP100, p=0.003).<br />

The same pre-LT factors were also significantly associated with<br />

HCC-related mortality post-LT. Conclusion: In HCC transplanted<br />

patients there was a marked discrepancy between radiological<br />

findings and pathological description of the explant. The best<br />

predictors of HCC recurrence were pre-LT aFP and diameter<br />

of the biggest nodule. Combined together could allow us to<br />

include in the current transplant criteria patients with biggest<br />

nodule diameter>3cm if aFP


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 419A<br />

on explant (>3 tumors, largest tumor >5cm, vascular invasion,<br />

metastases, or poor differentiation). NASH patients were less<br />

likely to undergo locoregional therapy (LRT) pre-LT (86 vs 92%;<br />

p


420A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

was 53 +/-0.6 days. By multivariate analysis, factors related<br />

to all causes of death were patients’ age >60 years (OR=1.2,<br />

95%CI; 1.1-1.3), length of hospital stay of >7 days (OR=1.1,<br />

95%CI; 1.02-1.2), male (OR=1.3, 95%CI; 1.2-1.4), living<br />

in Northern part of Thailand (OR=1.5, 95%CI; 1.3-1.8) and<br />

presence of complications during admission (OR=1.3, 95%CI;<br />

1.1-1.5. CONCLUSIONS: The disease burden of patients with<br />

cholangiocarcinoma in Thailand is significant with 0.65% of<br />

in-patients with GI. diseases and high mortality rate of 14%.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sombat Treeprasertsuk, Kittiyod<br />

Poovorawan, Ngamphol Soonthornworasiri, Roongruedee Chaiteerakij, Pisaln<br />

Mairiang, Kookwan Sawadpanich, Varocha Mahachai, Kamthorn Phaosawasdi<br />

417<br />

Short-term preoperative treatment with sorafenib for<br />

patients with resectable hepatocellular carcinoma (HCC):<br />

results of the BIOSHARE neoadjuvant phase II study<br />

from GERCOR IRC<br />

Mohamed Bouattour 1 , Laetitia Fartoux 2 , Olivier Rosmorduc 2 , Eric<br />

Vibert 3 , Charlotte E. Costentin 4 , Olivier Soubrane 5 , Olivier Scatton<br />

6 , Maxime Ronot 7 , Laurent Castera 1 , Muriel Garnier 8 , Armand<br />

De Gramont 8 , Jacques Belghiti 5 , Valerie Paradis 9 , Annemilaï<br />

Tijeras-Raballand 10 , Alexandra Hadengue 11 , David Brusquant 11 ,<br />

Benoist Chibaudel 11 , Eric Raymond 8 , Sandrine Faivre 8 ; 1 Hepatology,<br />

Beaujon University Hospital, Clichy, France; 2 Hepatology,<br />

Saint-Antoine University Hospital, Paris, France; 3 Hepato-Biliary<br />

Surgery, Paul-Brousse University Hospital, Villejuif,, France; 4 Hepatology,<br />

Henri-Mondor University Hospital, Creteil, France; 5 Hepato-Biliary<br />

Surgery, Beaujon University Hospital, Clichy, France;<br />

6 Hepato-Biliary Suregry, Saint-Antoine University Hospital, Paris,<br />

France; 7 Radiology, Beaujon University Hospital, Clichy, France;<br />

8 Medical Oncology, Centre Hospitalo-Univesitaire Vaudois<br />

(CHUV), Lauzanne, Switzerland; 9 Pathology, Beaujon University<br />

Hospital, Clichy, France; 10 AREC FILIA RESEARCH,, Paris, France;<br />

11 GERCOR, Paris, France<br />

Background. Neoadjuvant therapy offers the opportunity of<br />

investigating new agents in early stage and allows direct<br />

access to tumor biology. This open-label, multicenter, phase<br />

II study aimed at investigating the activity of sorafenib including<br />

radiological, pathological and molecular changes in tumor<br />

from patients with operable HCC. Patients and methods. A preoperative<br />

treatment consisting of 400 mg bid orally sorafenib<br />

was administered for 4 consecutive weeks followed by surgery<br />

1 week after sorafenib discontinuation. Primary endpoints<br />

were antitumor activity and histological changes between<br />

paired biopsies and plasma biomarkers, from baseline and<br />

post sorafenib treatment. Secondary endpoints were safety<br />

profile, R0 surgery and post-surgical complications. Residual<br />

disease characteristics were analyzed on surgical specimens.<br />

Results. Among 30 patients enrolled, 28 were evaluable for<br />

safety and 25 patients (21 men; median age: 61.5 years)<br />

were evaluable for the primary endpoint. All evaluable patients<br />

were Child Pugh A, 14 (56%) with chronic liver disease, 9<br />

(36%) with cirrhosis and 2 patient (8%) with no underlying<br />

liver disease. The baseline median tumor size was 37 mm<br />

(17-220 mm) and 21 patients (84%) had a single lesion. The<br />

median duration and dosing of sorafenib for evaluable patients<br />

were 28 days (21-35d) and 793 mg/day (477 – 843 mg/<br />

day) respectively. Overall, the safety profile of preoperative<br />

sorafenib was good. According to RECIST criteria, all patients<br />

showed stable disease after 4 weeks of sorafenib. Among 14<br />

patients evaluated according to mRECIST and Choi criteria,<br />

objective responses were observed for 4 (29%) and 7 (50%)<br />

patients respectively. All evaluable patients went on liver resection;<br />

median hospitalization duration was 8 days (4-59d) and<br />

no unexpected postoperative complication was observed. R0<br />

tumor resection was achieved in 22 patients (88%). Surgical<br />

specimen showed macrovascular and microvascular invasion<br />

as well as satellite nodules in 12%, 48%, and 36% respectively.<br />

Intratumor necrosis was observed in 13 patients of 20<br />

evaluable patients. Biomarkers from pre- and post-treatment<br />

tissue (Vimentin, CK19, E-Cadherin, N-Cadherin, CK7, MIB1,<br />

CD31, VEGF, CD133, CA9, p-S6, c-MET, CXCR4) and plasma<br />

(VEGF-C, Ang2, PlGF, c-KIT, SDF-1, HGF, TGFb1) are currently<br />

assessed and will be presented during the meeting. Conclusion.<br />

Preoperative BIOSHARE trial showed appropriate safety<br />

regarding both tolerance to sorafenib and surgical procedures.<br />

Short-term neoadjuvant treatment with sorafenib showed promising<br />

efficacy in terms of tumor response (mRECIST and Choi)<br />

and will allow detailed evaluation of molecular changes in<br />

patient tumor tissues.<br />

Disclosures:<br />

Mohamed Bouattour - Advisory Committees or Review Panels: Bayer Schering<br />

Pharma; Speaking and Teaching: Bayer Schering Pharma<br />

Olivier Rosmorduc - Advisory Committees or Review Panels: Syrtex, IPSEN;<br />

Speaking and Teaching: Bayer<br />

Laurent Castera - Speaking and Teaching: Gilead, BMS, Janssen, Echosens,<br />

Abbvie, Biopredictive<br />

The following authors have nothing to disclose: Laetitia Fartoux, Eric Vibert,<br />

Charlotte E. Costentin, Olivier Soubrane, Olivier Scatton, Maxime Ronot, Muriel<br />

Garnier, Armand De Gramont, Jacques Belghiti, Valerie Paradis, Annemilaï<br />

Tijeras-Raballand, Alexandra Hadengue, David Brusquant, Benoist Chibaudel,<br />

Eric Raymond, Sandrine Faivre<br />

418<br />

Analytic Morphomics Predicts Survival in Patients with<br />

Hepatocellular Carcinoma Treated with Transarterial<br />

Chemoembolization<br />

Neehar D. Parikh 1,3 , Amit G. Singal 2 , Peng Zhang 3 , Venkat Krishnamurthy<br />

3 , Pranab Barman 1 , Akbar K. Waljee 1,4 , Grace L. Su 1,3 ;<br />

1 Division of Gastroenterology, University of Michigan, Ann Arbor,<br />

MI; 2 Division of Digestive and Liver Diseases, Parkland Health and<br />

Hospital System, Dallas, TX; 3 VA Ann Arbor Healthcare System,<br />

Ann Arbor, MI; 4 Veterans Affairs Center for Clinical Management<br />

Research, VA Ann Arbor Healthcare System, Ann Arbor, MI<br />

The prognosis of patients with hepatocellular carcinoma (HCC)<br />

undergoing transarterial chemoembolization (TACE) is often<br />

uncertain. The existing HCC staging criteria are limited in their<br />

ability to predict survival after therapy. Analytic morphomics<br />

has been shown to predict patient outcomes in cirrhosis and<br />

other disease states. We aimed to determine the ability of<br />

analytic morphomics to predict outcome after TACE for HCC.<br />

We included patients from a single center (Ann Arbor VA Medical<br />

Center) who had TACE as the primary treatment for their<br />

HCC and who had a CT chest or abdomen prior to therapy.<br />

Analytic morphomic measurements were taken at the bottom of


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 421A<br />

the 11th thoracic vertebral level. We performed univariate and<br />

conditional inference tree analysis to identify the morphomic<br />

characteristics predictive of survival. Our primary outcome was<br />

actuarial survival at one year. We performed validation with<br />

patients who underwent TACE at two other centers (University<br />

of Michigan [n=85] and University of Texas Southwestern<br />

[n=72].) In total 76 patients (74 male; 2 female) were identified<br />

from the derivation cohort. Baseline characteristics, liver<br />

function and tumor stage are shown in the Table. Median survival<br />

was 564 days (95% CI: 416-713) from performance of<br />

the CT scan. Conditional inference tree analysis revealed that<br />

interstitial density (ITHU) was the only morphomic factor predictive<br />

of survival. Patients with ITHU above 0.709 had a 1-year<br />

survival of 51% (95% CI 43-59%) and those below 0.709 had<br />

a 1-year survival of 87% (95% CI: 81-93%) (p=0.001). External<br />

validation in two diverse external cohorts, showed that the<br />

ITHU cutoff was also predictive of one year survival in patients<br />

undergoing TACE at each site. In an exploratory analysis of the<br />

derivation cohort, we found that patients who had decompensation<br />

after TACE had a significantly higher ITHU than those<br />

who did not (p3 and FBG≥2.68g/L<br />

were independent risk factors of recurrence of HCC after LT.<br />

A scoring model was built to predict the risk of tumor recurrence,<br />

with a sensitivity of 68.3% and a specificity of 87.5%.<br />

Conclusion: Pretransplant elevated plasma fibrinogen level significantly<br />

increases the risk for tumor recurrence in patients<br />

after liver transplantation for HCC. The scoring model we built<br />

based on fibrinogen level and tumor number strongly correlates<br />

with tumor recurrence and may aid in the selection of patients<br />

that would most benefit from transplantation for HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Guo-Ying Wang, Yang Yang,<br />

Gui-Hua Chen<br />

420<br />

Racial Disparity in Hepatocellular Carcinoma Patients<br />

Diagnosed in an Urban Medical Center<br />

Kirthi Lilley, Paul H. Naylor, Omar Sadiq, Sarah Stern, Sindhuri<br />

Benjaram, Karthik Ravindran, Samran Haider, Ryan Morton, Ravi<br />

Anand, Anupama Devara, Karan Mathur, Philip A. Philip, Murray<br />

N. Ehrinpreis, Milton G. Mutchnick; Internal Medicine, Wayne<br />

State University School of Medicine, Detroit, MI<br />

Background: Hepatocellular carcinoma (HCC) is a significant<br />

liver disease outcome and based on a previous report from<br />

our group, occurs in 23% of untreated African Americans (AA)<br />

with chronic hepatitis C (CHC) who were followed for an average<br />

of 8 years in an urban GI clinic. The aim of this study was<br />

to evaluate racial disparity with respect to the role of CHC and<br />

the development of HCC. Tumor size at first diagnosis and the<br />

role of surveillance prior to diagnosis of HCC were the primary<br />

foci of the study. Methods: The electronic medical records of<br />

the largest health care provider in Southeast Michigan and<br />

its associated multi-specialty group were used to identify all<br />

patients with HCC between 2012 and 2014. Using ICD-9<br />

codes 115.0 (hepatocellular carcinoma) and 115.2 (malignant<br />

neoplasm of liver, NOS), a patient pool of 281 potential<br />

HCC patients (67% AA) was identified. From the pool, 125<br />

patients were confirmed to have HCC by imaging (US /CT/<br />

MRI) and/or biopsy. The remaining patients either had metastastic<br />

tumors or no confirmation of a HCC diagnosis. Race was<br />

determined by self-reporting, based primarily on skin color and<br />

was analyzed as AA (black) or Non-AA (non-black; Caucasian,<br />

Hispanic, Asian and other) Results: The 125 confirmed<br />

HCC patients were primarily AA (76%) and male (80%). The<br />

average age at diagnosis of HCC was 62 years (range for<br />

95% was 61-64 years). HCV was a dominant risk factor for<br />

the development of HCC in AA but not in non-AA (89% vs<br />

45%; p


422A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

the majority of patients were not diagnosed in a surveillance<br />

protocol. The majority of patients undergoing surveillance also<br />

had non-curable tumors at diagnosis. The optimal method for<br />

preventing HCC in AA remains effective treatment with the new<br />

anti-viral therapies which, unlike previous therapies are highly<br />

effective in AA.<br />

Disclosures:<br />

Milton G. Mutchnick - Advisory Committees or Review Panels: BMS; Grant/<br />

Research Support: Janssen, Gilead; Speaking and Teaching: Janssen, Gilead,<br />

Abbvie, CLDF, Simply Speaking<br />

The following authors have nothing to disclose: Kirthi Lilley, Paul H. Naylor,<br />

Omar Sadiq, Sarah Stern, Sindhuri Benjaram, Karthik Ravindran, Samran<br />

Haider, Ryan Morton, Ravi Anand, Anupama Devara, Karan Mathur, Philip A.<br />

Philip, Murray N. Ehrinpreis<br />

421<br />

In patients with HCC and macrovascular invasion<br />

treated with sorafenib, AFP value≤1000 UI/L at baseline<br />

and vascular invasion not including 1 st order<br />

branches or main trunk (Vp1/2) are associated with<br />

prolonged survival.<br />

Charlotte E. Costentin 1 , Françoise Roudot-Thoraval 2 , Thomas<br />

Decaens 3 , Nathalie Ganne-Carrié 4 , Bernard Paule 5 , Eric Vibert 5 ,<br />

Mélanie Chiaradia 6 , Christian Letoublon 7 , Julien Calderaro 8 , Giuliana<br />

Amaddeo 1 , Alexis Laurent 9 , Ivan Bricault 10 , Olivier Seror 11 ,<br />

Didier Samuel 5 , Ariane Mallat 1 , Christophe Duvoux 1 ; 1 Hepatology,<br />

Hôpital Henri Mondor, Creteil, France; 2 Public Health, Henri<br />

Mondor Hospital, Creteil, France; 3 Hepato-gastroenterology, CHU<br />

Grenoble, Grenoble, France; 4 Hospital, Jean Verdier, Bondy,<br />

France; 5 Hepatology-hepatobiliary surgery, Paul Brousse Hospital,<br />

Villejuif, France; 6 Radiology, Henri Mondor Hospital, Creteil,<br />

France; 7 Surgery, CHU Grenoble, Grenoble, France; 8 Pathology,<br />

Henri mondor Hospital, Creteil, France; 9 Surgery, Henri Mondor<br />

Hospital, Creteil, France; 10 Radiology, CHU Grenoble, Grenoble,<br />

France; 11 Radiology, Jean Verdier Hospital, Bondy, France<br />

Introduction: According to EASL-OERTC guidelines, sorafenib<br />

is the standard of care for advanced hepatocellular carcinoma<br />

(HCC) BCLC-C although in the SHARP trial, macrovascular<br />

invasion (MVI) was associated with a poor prognosis. However,<br />

survival had been assessed for patients with MVI with<br />

or without extrahepatic spread (EHS) (8 months) but not in<br />

the group of patients with MVI but no EHS. The aim of this<br />

study was to describe outcome and predictors of survival<br />

in patients with HCC and MVI but no extra-hepatic spread<br />

treated with sorafenib. Methods: 67 patients with HCC and<br />

MVI but no EHS treated with sorafenib were identified in the<br />

database of 4 French centers (Bondy, Créteil, Grenoble,Villejuif)<br />

and retrospectively studied. Results: Patients (pts) were<br />

mostly males (88.1%), mean age 65.5±9.6 with underlying<br />

cirrhosis (89.4%). Etiology of liver disease was alcohol ±NASH<br />

(38.8%), viral hepatitis (35.8%), NASH without alcohol<br />

(14.9%), other (10.4%). Child-Pugh score was A and B in 59<br />

(86.4%) and 8 (13.6%) pts, respectively. HCC was uni-nodular<br />

in 17 pts (25.4%), infiltrative in 32 pts (47.8%), bilobar<br />

in 22 pts (33.3%). Median size of the larger nodule was 70<br />

[50-100] mm. MVI distribution was: Vp1+2 (invasion of or<br />

distal to the 2 nd order branches of the portal vein (PV)) in 8<br />

pts (12.0%), Vp3 (invasion of 1 st order branches of the PV)<br />

in 20 pts (30%), Vp4 (invasion of the main trunk and/or contra-lateral<br />

PV branch) in 32 pts (47.5%) and hepatic vein±Vp<br />

in 7 pts (10.5%). Median AFP level at baseline was 408 [32-<br />

2695] ng/ml. Median follow up was 6.7 [4.0-15.1] months.<br />

Median treatment duration on sorafenib was 6.4 months, and<br />

20 (30%) pts had additional therapies during follow up, either<br />

during sorafenib therapy or as second line (including TACE,<br />

systemic therapy, radiotherapy or radioembolization). Median<br />

overall survival was 12.6 months (IC95%: 8.4-16.7). In univariate<br />

analysis, AFP>1000 UI/L and 1 st order branches/troncular<br />

portal vascular invasion (Vp3-Vp4) were negatively associated<br />

with survival (p=0.049 and 0.031 respectively). In multivariate<br />

analysis, both parameters were independently associated<br />

with death: AFP>1000 (HR 2.188 [1.09-4.41], p=0.029),<br />

Vp3-Vp4 (HR 2.874[1.19-6.94], p=0.019). Median overall<br />

survival in pts with these 2 prognostic factors was 8.4 [4.2-<br />

12.5] months vs 16.9 [11.6-22.2] months in pts with one or<br />

none of these prognostic factors (p 1000<br />

UI/L and Vp3-Vp4 MVI are independent predictors of reduced<br />

survival. Long-term survival (median 16 months) was observed<br />

in pts with one or none of these 2 prognostic factors.<br />

Disclosures:<br />

Françoise Roudot-Thoraval - Advisory Committees or Review Panels: Roche; Consulting:<br />

LFB biomedicaments; Speaking and Teaching: gilead, Janssen, BMS,<br />

Roche<br />

Nathalie Ganne-Carrié - Advisory Committees or Review Panels: Roche; Speaking<br />

and Teaching: BMS, Gilead, Bayer<br />

Olivier Seror - Board Membership: Bayer Healthcare; Consulting: Olympus;<br />

Speaking and Teaching: Angiodynamics<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

Christophe Duvoux - Advisory Committees or Review Panels: Novartis, Roche,<br />

Novartis, Roche, Novartis, Roche, Novartis, Roche; Speaking and Teaching:<br />

Astellas, Astellas, Astellas, Astellas<br />

The following authors have nothing to disclose: Charlotte E. Costentin, Thomas<br />

Decaens, Bernard Paule, Eric Vibert, Mélanie Chiaradia, Christian Letoublon,<br />

Julien Calderaro, Giuliana Amaddeo, Alexis Laurent, Ivan Bricault, Ariane Mallat<br />

422<br />

Lamivudine vs. Entecavir for the newly diagnosed hepatitis<br />

B-virus related Hepatocellular carcinoma<br />

jung hee kim, Dong Hyun Sinn, Kyunga Kim, Geum-Youn Gwak,<br />

Yong-Han Paik, Moon Seok Choi, Joon Hyeok Lee, Kwang Cheol<br />

Koh, Seung Woon Paik; samsung medical center, Seoul, Korea<br />

(the Republic of)<br />

Antiviral therapy is a key element in the management of hepatitis<br />

B virus (HBV)-related hepatocellelular carcinoma (HCC)<br />

patients. However, little is known whether potent drug, entecavir,<br />

is more effective than a less potent drug, lamivudine,<br />

in reducing the risk of death, hepatic decompensation and<br />

recurrence in HBV-related HCC. A retrospective cohort of 451<br />

newly-diagnosed, HBV-related HCC patients without antiviral<br />

therapy at diagnosis, and started antiviral therapy with either<br />

entecavir (n = 249) or lamivudine (n=202) were enrolled.<br />

Overall survival, new-onset hepatic decompensation (ascites,<br />

variceal bleeding, hepatic encephalopathy), and recurrence<br />

after curative therapy were compared. Baseline HBV DNA levels<br />

were not diffference between two groups but rescue therapy<br />

was more frequent in lamvudine group.(45% vs. 6.4%,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 423A<br />

vious hepatic decompensation, and was also independent risk<br />

factor for recurrence after curative therapy [HR (95% CI): 1.84<br />

(1.25-2.72), p = 0.002] among 117 patients who received<br />

curative therapy. The overall survival, decompensation-free<br />

survival, and recurrence-free survival was better in entecavir<br />

treated patients than lamivudine treated patients . The survival<br />

benefit of choosing high potent drug over less potent drug will<br />

be higher when patient survival is long enough. Low-potent<br />

drug may be a valuable option for those with low expected<br />

survival, especially in resource-limited setting, but high-potent<br />

drug should be the preferred choice in the rest of HBV-related<br />

HCC patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: jung hee kim, Dong Hyun Sinn,<br />

Kyunga Kim, Geum-Youn Gwak, Yong-Han Paik, Moon Seok Choi, Joon Hyeok<br />

Lee, Kwang Cheol Koh, Seung Woon Paik<br />

423<br />

Long term outcome after resection of intermediate<br />

hepatocellular carcinoma: comparison to transarterial<br />

chemoembolization<br />

DAEGEON AHN, Dong Hyun Sinn, Geum-Youn Gwak, Moon<br />

Seok Choi, Joon Hyeok Lee, Kwang Cheol Koh, Seung Woon<br />

Paik, Byung Chul Yoo; Gastroenterology, Samsung Medical Center,<br />

Seoul, Korea (the Republic of)<br />

Background and Aims: Transarterial chemoembolization<br />

(TACE) is the first-line recommended therapy for patients with<br />

intermediate hepatocellular carcinoma (HCC) in the BCLC staging<br />

system. However, in clinical practice, some patients receive<br />

hepatectomy. We compared long-term outcome in these<br />

patients. Methods: A total of 277 patients with intermediate<br />

stage HCC who were treated with either hepatectomy or TACE<br />

were analyzed. Results: The median survival was significantly<br />

better for resection than TACE (5.3 vs. 2.6 years, p = 0.002),<br />

however, when adjusted for age, Child-Pugh Class, maximal<br />

tumor size, tumor number, uni- or bi-lobal involvement, and AFP<br />

levels, there was no significant survival difference by treatment<br />

modality [Hazard ratio (HR): 1.42, 95% CI (0.91-2.22), p =<br />

0.11]. In sub-group analysis, resection showed better survival<br />

than TACE in Child-Pugh score 5, maximal tumor size ≤ 5cm<br />

or > 10cm, and tumor number ≤ 3, however, there was no<br />

survival difference in patients with Child-Pugh score ≥ 6, tumor<br />

size between 5-10cm, and tumor number ≥ 4. When patients<br />

were sub-grouped according to the Child-Pugh Class, tumor<br />

number and size, resection (n = 26) provided survival benefit<br />

over TACE (n = 58) for those with Child-Pugh class 5, less than<br />

3 tumors, and tumor size ≤ 5cm or > 10cm [adjusted HR: 0.41<br />

(95% CI: 0.20-0.84), p = 0.015], but not the rest of patients<br />

[adjusted HR for resection (n = 26) vs. TACE (n = 167): 0.73<br />

(95% CI: 0.44-1.21), p = 0.22]. Conclusion: Resection can<br />

be first-line treatment option for intermediate stage HCC when<br />

certain criteria are met. In this study, child-Pugh score 5, tumor<br />

number ≤ 3, and tumor size ≤ 5 or > 10 cm were the criteria.<br />

Disclosures:<br />

The following authors have nothing to disclose: DAEGEON AHN, Dong Hyun<br />

Sinn, Geum-Youn Gwak, Moon Seok Choi, Joon Hyeok Lee, Kwang Cheol Koh,<br />

Seung Woon Paik, Byung Chul Yoo<br />

424<br />

TBI 302: A first-in-class therapy for the treatment of<br />

advanced stage liver cancer<br />

Steve Brookes 1 , Gary Levy 1,2 , Jin Seog Seo 1 , Murray J. Cutler 1 ,<br />

Yvonne Frater 1 , Gord Adamson 1 , David Bell 1 ; 1 Drug Development,<br />

Therapure Biopharma Inc., Mississauga, ON, Canada; 2 Multi<br />

Orgon Transplant Program, University of Toronto Transplant Institute,<br />

Toronto, ON, Canada<br />

Hepatocellular carcinoma (HCC) is the second most common<br />

cause of death from cancer worldwide; few treatment options<br />

are available, especially for advanced stage disease. Current<br />

standard of care provides only a modest improvement in overall<br />

survival and is associated with poor quality of life. Therapure<br />

has developed a novel drug delivery platform based upon<br />

the attachment of therapeutic drugs to hemoglobin (Hb) as a<br />

means of targeting the liver. TBI 302 is a hemoglobin-drug conjugate<br />

(HDC) designed to deliver the anticancer drug floxuridine<br />

(FUdR) to the liver to improve the effectiveness of treatment<br />

for HCC. Systemic toxicity associated with free FUdR restricts its<br />

use to locoregional administration (0.2-0.5 mg/kg/d) via continuous<br />

hepatic arterial infusion (HAI). Although HAI of FUdR<br />

can reduce hepatic tumor burden, toxicity from HAI FUdR and<br />

complications associated with direct hepatic infusion pumps<br />

can be significant. HDC technology uses Hb as an innovative<br />

drug carrier that exploits the natural pathway for hemoglobin<br />

clearance predominantly through the liver to provide selective<br />

drug targeting while preserving FUdR activity following standard<br />

intravenous (IV) infusion. To establish a safe starting dose<br />

for a first-in-human Phase 1 trial, a GLP-compliant repeat dose<br />

preclinical safety study of TBI 302 in cynomolgus monkeys was<br />

conducted. TBI 302 administered by 1-hour IV infusion once<br />

per week for 8 weeks at doses of 2, 5, and 10.5 mg/kg (0.08,<br />

0.19, and 0.4 mg/kg FUdR, respectively) was well tolerated<br />

at all dose levels. Increasing doses of TBI 302 resulted in proportional<br />

increases in area under the concentration-time curve<br />

(AUC) and maximum concentration (Cmax) of total plasma<br />

FUdR. No clinical signs or biochemical toxicity was associated<br />

with IV infusion of TBI 302. The no-observed-adverse-effect<br />

level (NOAEL) of TBI 302 was determined to be the highest<br />

dose level of 10.5 mg/kg (0.4 mg/kg FUdR). TBI 302 is projected<br />

to have a half-life of approximately 5 hours in humans. A<br />

Phase I safety study of TBI 302 as second-line therapy in HCC<br />

has been approved by the US FDA. The primary objective is<br />

to determine safety and tolerability of TBI 302. The secondary<br />

objectives are to determine TBI 302 pharmacokinetics and<br />

effects on tumor burden. Therapure’s HDC platform represents<br />

a new class of therapy that offers liver-specific targeting for<br />

a wide range of hepatic diseases, while potentially reducing<br />

extra-hepatic toxicity of drugs.<br />

Disclosures:<br />

Jin Seog Seo - Employment: Therapure Biopharma Inc<br />

Murray J. Cutler - Employment: Therapure Biopharma<br />

Yvonne Frater - Management Position: Therapure<br />

Gord Adamson - Employment: Therapure Biopharma Inc.<br />

David Bell - Employment: Therapure Biopharma Inc.<br />

The following authors have nothing to disclose: Steve Brookes, Gary Levy


424A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

425<br />

Staging Hepatocellular Carcinoma in Non-Cirrhotic<br />

Patients<br />

Philippe Kolly 1,2 , Helen Reeves 3,4 , Marina Knöpfli 2 , Jean-Francois<br />

Dufour 1,2 ; 1 Hepatology, Department of Clinical Research, University<br />

of Bern, Bern, Switzerland; 2 University Clinic for Visceral<br />

Surgery and Medicine, Inselspital Bern, Bern, Switzerland; 3 The<br />

Northern Institute for Cancer Research, Newcastle University,<br />

Newcastle-upon-Tyne, United Kingdom; 4 The Liver Unit, Freeman<br />

Hospital, Newcastle-upon-Tyne, United Kingdom<br />

To design a staging system for HCC in non-cirrhotic patients,<br />

we used a pooled set of 864 patients combining 2 prospective<br />

European cohorts. Among them, 221 had a non-cirrhotic<br />

liver. Cirrhosis was excluded based on clinical presentation<br />

and imaging. The staging and treatment algorithm was constructed<br />

by using Cox regression, regression tree analysis and<br />

evidenced based literature. We created the Non-Cirrhotic Liver<br />

Cancer (NCLC) staging system. Based on the ECOG PS, tumor<br />

burden (size and number of nodules), presence of vascular<br />

invasion and metastasis, this staging system proposes 4 stages.<br />

Each stage is associated with treatment options, separated in<br />

first and second intention treatments (Fig.). We assessed the<br />

discriminatory power of the staging system by Log-Rank (p<br />

< 0.001). The prognostic ability of the staging system was<br />

assessed with Harrell’s concordance index (0.729). HCC<br />

developing in non-cirrhotic patients represents 10-40% of<br />

HCC. With the growing incidence of NAFLD-associated HCC,<br />

this proportion will rise. The BCLC staging system, which has<br />

been endorsed by AASLD and EASL, has not been developed<br />

for non-cirrhotic patients. The statistical analysis revealed that<br />

the performance status and the tumor burden are discriminating<br />

parameters to classify non-cirrhotic patients with HCC in 4<br />

different stages with different prognostics and treatment allocation.<br />

A staging system for HCC developing in non-cirrhotic<br />

patients has become essential. The NCLC has been designed<br />

specifically for this purpose and will need to be further developed.<br />

Figure legend: NCLC staging and treatment algorithm.<br />

This algorithm stages HCC in non-cirrhotic patients according<br />

to ECOG PS, tumor burden and the existence of vascular invasion<br />

or metastasis. It suggests 4 stages and the corresponding<br />

treatments, separated in first and second intention.<br />

426<br />

Clinical significance of microRNA-31 expression in<br />

human hepatocellular carcinoma<br />

Keun Hur 1 , Se Young Jang 2 , Soo Young Park 2 , Gyeonghwa Kim 1 ,<br />

Yong-Hun Choi 1 , Yu Rim Lee 2 , Su Hyun Lee 2 , Sun Kyung Jang 2 ,<br />

Won Young Tak 2 , Young Oh Kweon 2 ; 1 Department of Biochemistry<br />

and Cell Biology, Kyungpook National University School of<br />

Medicine, Daegu, Korea (the Republic of); 2 Department of Internal<br />

Medicine, Kyungpook National University Hospital, Daegu, Korea<br />

(the Republic of)<br />

Background & aims Hepatocellular carcinoma (HCC) is one<br />

of the most leading cause of cancer-related death worldwide.<br />

It is therefore important to understand the mechanistic roles of<br />

biomolecules involved in HCC development. MicroRNAs (miR-<br />

NAs) are small noncoding RNAs that act as an important regulator<br />

of gene expression in various tumor development as well<br />

as metabolic diseases. This study aim to investigate the role<br />

of miR-31 in HCC and non-alcoholic steatohepatitis (NASH).<br />

Methods We analyzed clinical specimens from HCC, matched<br />

normal liver, and NASH tissues. MicroRNAs expression levels<br />

were evaluated by quantitative real-time PCR (qRT-PCR) and<br />

miR-31 expression was normalized relative to RNU6B expression.<br />

In addition, we determined the clinical significance of<br />

the miR-31 expression in matching tissue and serum samples<br />

from HCC patients. Results MiR-31 expression was significantly<br />

upregulated in HCC and NASH tissues compared with normal<br />

liver tissues (p=0.0015). In HCC patients, lower level of<br />

miR-31 (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 425A<br />

Pearson chi squared tests we examined the rates of false-positive<br />

HCC by UNOS region and transplant center. Results:<br />

Among 3,852 transplant recipients with T2MELD exceptions<br />

and available explant data, 262 (6.8%) had no evidence of<br />

HCC identified on explant pathology. Rates of false-positive<br />

HCC were higher for patients who underwent prior HCC treatment,<br />

10.1% vs 6.5%, p = 0.014. Rates of false-positive HCC<br />

varied by UNOS region: 2.1% false-positive rate in region 6<br />

versus 12.0%% in region 7 (p = 0.001). There was significant<br />

among-center variability (Figure) in false-positive rates among<br />

those centers who transplanted ≥30 recipients with T2MELD<br />

exceptions during the study period: ranging from 0% in four<br />

centers to 46.0%at one center, p90%, even in patients with advanced<br />

hepatic fibrosis. On the other hand, hepatocellular carcinoma<br />

(HCC) develops in up to 4% of patients after successful treatment<br />

of continuous HCV infection using interferon therapy. It is<br />

important to evaluate hepatic pathology after HCV eradication.<br />

The aim of the present study was to characterize microRNA<br />

(miRNA) expression in the liver tissue of patients who achieved<br />

an SVR. Patients and Methods: 71 tissue samples from patients<br />

who underwent HCC resection were evaluated in the present<br />

study: 61 HCCs from patients with continuously infected HCV<br />

(HCV-HCC) and 10 from patients who had achieved SVR (SVR-


426A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

HCC). We also included non-tumorous tissues (SVR-NT) from<br />

four patients with SVR-HCC and liver tissue (SVR-CH) from four<br />

patients with SVR but without HCC. Total RNA was extracted<br />

from the liver samples. The miRNA expression patterns were<br />

analyzed using microarray. Using leave-one-out cross-validation,<br />

we picked up several miRNAs to classify samples between<br />

SVR-HCC and HCV-HCC and between SVR-NT and SVR-CH. In<br />

addition, target gene expression was quantified after miRNA<br />

overexpression in HEK293 cells. Results: We were able to<br />

discriminate between SVR-HCC and HCV-HCC with 75.36%<br />

accuracy using the expression pattern of six specific miRNAs<br />

(let-7a, let-7f, miR-21, miR-122, miR-720, and miR-923). The<br />

expression levels of 37 miRNAs were significantly lower in<br />

HCV-HCC than in SVR-HCC, whereas the expression of 25<br />

miRNAs was significantly higher in HCV-HCC than in SVR-<br />

HCC (P < 1.0e-05). The expression of thrombospondin 1 was<br />

regulated in an opposing manner by miR-30a-3p in SVR-HCC<br />

and HCV-HCC. In non-tumorous tissues, the expression pattern<br />

of seven miRNAs distinguished between SVR-CH and SVR-NT<br />

with 87.50% accuracy. Conclusions: Comprehensive miRNA<br />

expression analyses can not only differentiate between SVR-<br />

HCC and HCV-HCC, but can also predict hepatocarcinogenesis<br />

after achieving an SVR.<br />

Determining SVR-HCC using 6 miRNAs (let-7a, let-7f, miR-21,<br />

miR-122, miR-720, and miR-923) expression pattern in LOOCV<br />

analysis.<br />

The column and row showed prediction and result, respectively.<br />

For example in 5 years observation, 47 examples of SVR-HCC<br />

expected as HCV-HCC, and it were 3 examples in HCV-HCC.<br />

The prediction accuracy between HCV-HCC and SVR-HCC-5<br />

years was 75.36%.<br />

Disclosures:<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

The following authors have nothing to disclose: Akihiro Tamori, Yoshiki<br />

Murakami, Shoji Kubo, Saori Itami, Sawako K. Uchida, Hiroyasu Morikawa,<br />

Masaru Enomoto, Shigekazu Takemura, Toshihito Tanahashi, Y-h Taguchi<br />

430<br />

Exploration of cancer-prone gut microbiota in chronic<br />

hepatitis C patients<br />

Noriho Iida, Eishiro Mizukoshi, Shuichi Kaneko; Kanazawa Univ,<br />

Kanazawa, Japan<br />

Purpose: Recently it has been recognized from animal <strong>studies</strong><br />

that dysbiosis of gut microbiota is one of the causes of liver<br />

carcinogenesis. However, character of gut microbiota prone<br />

to liver carcinogenesis in human patients remains unclear. In<br />

this study we aimed to characterize gut microbiota of chronic<br />

hepatitis C patients with hepatocellular cancers (HCC) and<br />

determine whether the microbiota derived from HCC patients<br />

actually cause liver carcinogenesis in mice or not. Methods:<br />

Feces were collected from 11 chronic hepatitis C patients with<br />

primary HCC, 8 chronic hepatitis C patients without HCC<br />

(CLD), and 6 healthy donors (HD). After DNA was extracted<br />

from the fecal samples, libraries were generated with Illumina<br />

Nextera DNA kits to do whole genome shotgun sequencing.<br />

The sequencing was performed with Illumina Miseq. After quality<br />

control, removal of human genome sequences and PCR<br />

replicative sequences, the sequences were analyzed with<br />

MetaPhLAN or HUMAnN, to know taxonomic profile or presence<br />

of genes, respectively. To find significantly differential<br />

bacteria species or genes, linear discriminative analyses (LDA)<br />

were done. To test causality of the gut microbiota derived from<br />

HCC patients on liver carcinogenesis, fecal suspension made<br />

from the patients or HD was transplanted to C57BL6 mice in<br />

which gut bacteria were depleted by oral antibiotic cocktail.<br />

The human feces-transplanted mice were injected with diethylnitrosamine<br />

(DEN) and CCl4 to induce liver tumors. Results:<br />

Total 57,259,209 sequences were analyzed on MetaPhLAN<br />

and HUMAnN. Alpha-Diversity in the sample (Simpson’s index)<br />

was not significantly different among the groups. HCC group<br />

was not differentially clustered from HD group. However, in<br />

the comparison of HCC vs HD on LDA analysis, 8 taxa were<br />

significantly increased and 2 taxonomy decreased in HCC<br />

(P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 427A<br />

non-HCC liver tissue in relation to RNU6B. Results: Considering<br />

all 45 HCC patients, lower levels of hsa-miR- 125a-5p were<br />

observed in HCC tissue than in non-HCC liver tissues (M+SD<br />

4.84+3.69 vs. 8.27+4.5 A.U., p=0.0002). This difference<br />

was highly significant to statistical analysis in the 35 HCV-patients,<br />

4.21+ 3.47 vs. 7.79+4.75 AU, p=0.0009 and lower<br />

and not statistically significant significance in the 5 HBV-patients,<br />

7.38+2.08, vs. 10.7+2.4AU, and. in the 5 patients<br />

with NASH-related cirrhosis, 6.72+5.24 vs. 9.2+3.99AU. In<br />

addition, the level of miR-125a-5p in the HCC tissue of the 35<br />

HCV-related cirrhosis was lower than that observed in the 10<br />

non-HCV patients (4.21+3.47 vs. 7.05+3.77 AU, p=0.023).<br />

Besides, the levels of has-miR-125a-5p in HCC tissue were<br />

moderately lower in the 41 patients with Child-Pugh score A<br />

(4.67+3.47 AU) than in the 4 with score B (7.05+3.77 AU,<br />

p=0.7), such as in the 40 patients with BCLC class A as compared<br />

with the 5 with class B or C (4.66+3.78 vs. 6.31+2.79<br />

AU, p=0.21). Conclusions: These data suggest an oncosuppressor<br />

effect of microRNA hsa-miR-125a-5p on HCV-related<br />

HCC and possibly even in non-HCV HCC,, an observation<br />

deserving further investigation<br />

Disclosures:<br />

The following authors have nothing to disclose: Nicola Coppola, Giorgio de<br />

Stefano, Nicola Mosca, Valentina Iodice, Carmine Minichini, Filomena Castiello,<br />

Nunzia Farella, Mario Starace, Nicoletta Potenza, Giulia Liorre, Lorenzo Onorato,<br />

Evangelista Sagnelli, Aniello Russo<br />

432<br />

Surgical resection versus radiofrequency ablation plus<br />

drug-eluting beads transcatheter arterial chemoembolization<br />

for the treatment of solitary large hepatocellular<br />

carcinoma. A single center experience<br />

Antonio Saviano 1 , Roberto Iezzi 2 , Felice Giuliante 3 , Lucia Salvatore<br />

1 , Caterina Mele 3 , Alessandro Posa 2 , Mariachiara Campanale<br />

1 , Emanuele Rinninella 1 , Valentina Cesario 1 , Federico<br />

Barbaro 1 , Marco Biolato 1 , Maria Assunta Zocco 1 , Laura Riccardi 1 ,<br />

Brigida E. Annicchiarico 1 , Massimo Siciliano 1 , Anna Maria De<br />

Gaetano 2 , Antonio Grieco 1 , Gian Ludovico Rapaccini 1 , Antonio<br />

Gasbarrini 1 , Lorenzo Bonomo 2 , Maurizio Pompili 1 ; 1 Department<br />

of Internal Medicine, Catholic University, Rome, Italy; 2 Department<br />

of Bioimaging and Radiological Sciences, Institute of Radiology,<br />

Catholic University, Rome, Italy; 3 Hepatobiliary Surgery Unit,<br />

Department of Surgery, Catholic University, Rome, Italy<br />

Background and aims: Aim of this study was to compare<br />

liver resection (RES) and single-step combined therapy with<br />

radiofrequency ablation and drug-eluting beads transarterial<br />

chemoembolization (RFA+TACE) in patients with single hepatocellular<br />

carcinoma (HCC) larger than 3 cm and compensated<br />

cirrhosis. The primary end-points were overall survival (OS)<br />

and tumour recurrence (TR) rates. Methods: This retrospective<br />

study involved 56 Child-Pugh A cirrhotic patients (25 in RES<br />

group and 33 in RFA +TACE group) with single HCC observed<br />

between 2010 and 2014 in our Center. All the patients did<br />

not show macroscopic vascular invasion and/or extrahepatic<br />

tumor spreading before treatment. In all cases the treatment<br />

choice was made after multidisciplinary evaluation. RFA and<br />

TACE were performed using a single step procedure (RFA<br />

during balloon artery occlusion of the tumour feeding artery<br />

followed by TACE) . The cumulative OS and TR rates were analysed<br />

using the Kaplan Meyer method. Results: The median age<br />

and follow up period of the whole population were, respectively,<br />

70 yrs and 16.5 months. The median HCC size was<br />

4,5 cm in RES group (range 3-7.4 cm) and 4 cm in RFA+TACE<br />

group (range 3-6.8 cm) (p 0.150). The tumours larger than<br />

5 cm were 40.0% in RES group and 21.9% in RFA group (p<br />

0.138). Demographic, clinical and laboratory parameters did<br />

not differ significantly between groups but RFA+TACE patients<br />

showed clinically evident portal hypertension in a higher rate<br />

of cases (32.0% in RES group, 63.6% in RFA+TACE group, p<br />

0.017).Treatment-related complications did not differ significantly<br />

between groups. OS rates at 1-3 years were 91.8%<br />

- 75.2% in RES group and 86.3% - 61.6% in RFA+TACE group<br />

(p 0.463). Overall TR rates at 1-3 years were significantly<br />

lower in resected patients (RES group: 32.7% - 42.3%; RFA+-<br />

TACE group 43% - 74.7%, p 0.039). Both 3-year local tumour<br />

progression (RES group 31.1%, RFA+TACE group 74.6%, p<br />

0.004) and distant intrahepatic recurrence (RES group 34.8%,<br />

RFA+TACE group 77.0%, p 0.043) occurred significantly more<br />

frequently in RFA+TACE group. Considering only patients with<br />

HCC ≤ 5 cm, the OS rates at 1-3 years were 100.0% - 84.6%<br />

in RES group and 87.1% - 62.3% in RFA+TACE group (p<br />

0.218). Furthermore, TR rates at 1-3 years were significantly<br />

lower in RES group than in RFA+TACE group (20.6% - 33.8%<br />

versus 45.3% - 65.8%, p 0.049). Conclusions: RES appears<br />

to be the choice treatment in compensated cirrhotic patients<br />

with single HCC larger than 3 cm. One step combined RFA+-<br />

TACE treatment is an effective alternative therapeutic option in<br />

patients who are not candidates to RES after multidisciplinary<br />

evaluation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Antonio Saviano, Roberto Iezzi,<br />

Felice Giuliante, Lucia Salvatore, Caterina Mele, Alessandro Posa, Mariachiara<br />

Campanale, Emanuele Rinninella, Valentina Cesario, Federico Barbaro,<br />

Marco Biolato, Maria Assunta Zocco, Laura Riccardi, Brigida E. Annicchiarico,<br />

Massimo Siciliano, Anna Maria De Gaetano, Antonio Grieco, Gian Ludovico<br />

Rapaccini, Antonio Gasbarrini, Lorenzo Bonomo, Maurizio Pompili<br />

433<br />

Efficacy of Radiotherapy in addition to Chemoembolization<br />

and Hepatic Arterial Infusion Chemotherapy versus<br />

Sorafenib in Advanced Hepatocellular Carcinoma with<br />

Portal Vein Thrombosis<br />

Sun Young Yim 1 , Soon Ho Um 1 , Tae Hyung Kim 1 , Jem Ma Ahn 1 ,<br />

Beom Jin Park 2 , Yeon Seok Seo 1 , Ho Sang Ryu 1 ; 1 Department of<br />

Internal Medicine, Korea University College of Medicine, Seoul,<br />

Korea (the Republic of); 2 Radiology, Korea University College of<br />

Medicine, Seoul, Korea (the Republic of)<br />

Background and aim: Sorafenib is the only standard treatment<br />

for advanced hepatocellular carcinoma with (HCC). However,<br />

its efficacy is not satisfactory and other treatment options are<br />

required. This study investigated the efficacy of chemoembolization<br />

and hepatic arterial infusion chemotherapy (HAIC)<br />

with or without radiotherapy versus sorafenib alone. Methods:<br />

This single-center retrospective study involved 105 patients<br />

of advanced HCC with portal vein tumor thrombosis (PVT).<br />

Enrolled patients had either child-pugh (CP) class A or B liver<br />

cirrhosis whom were classified into 3 groups: 1) Sorafenib<br />

alone, n=20; 2) Chemoembolization and HAIC, n=26; 3)<br />

Chemoembolization and HAIC with radiotherapy, n=59.<br />

Sorafenib was initiated with 400mg twice daily and HAIC<br />

was based on cisplatin and 5-fluorouracil regimen, performed<br />

every 4 weeks. Response of PVT was determined 3 months<br />

after completion of treatment and was regarded as responsive<br />

when there is at least partial response. Overall survival (OS)<br />

was analyzed among the treatment groups and factors associated<br />

with mortality were evaluated using multivariate analysis.<br />

Result: The median radiation dose, sorafenib treatment duration<br />

and chemoembolization sessions were 50 Gy, 40 days,<br />

and 4 sessions, respectively. Proportion of patients according<br />

to the degree of PVT (main/both vs. first order vs. segmental<br />

PV) and bilobar tumor mass involvement did not differ among<br />

the three groups of treatment. However, PVT response rate


428A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

was significantly higher in group 3 (13.5% vs. 6.7% vs. 55%,<br />

P=0.014) with lower incidence of solid organ metastasis (60%<br />

vs. 23.1% vs. 18.6%, P=0.001) and CP class B (70%, 50%,<br />

25.4%%, P=0.001) compared to group 2 and 1. In univariate<br />

cox analysis, treatment modalities, presence of either lymph<br />

node or other organ metastasis, CP class B, and decrease in<br />

AFP levels were associated with survival. However, multivariate<br />

analysis revealed that treatment modalities (group 1 vs.<br />

2, HR, 0.244; 95% CI, 0.06-0.999, P=0.05, group 1 vs. 3,<br />

HR, 0.121; 95% CI 0.028-0.51, P=0.004 and group 2 vs.<br />

3, HR 0.495; 95% CI, 0.25-0.98, P=0.044) and decrease in<br />

serum AFP level within 2 months of treatment (HR, 1.813; 95%<br />

CI, 1.204-2.72; P=0.004) were the only independent factors<br />

associated with survival. Median OS was significantly higher<br />

in patient treated with radiotherapy (group3, 11.1 months)<br />

than group 2 (3.6 months, log rank P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 429A<br />

Disclosures:<br />

The following authors have nothing to disclose: Andrea Casadei Gardini, Giorgia<br />

Marisi, Paola Ulivi, Luca Faloppi, Emanuela Scarpi, Mario Scartozzi, Francesco<br />

G. Foschi, Massimo Iavarone, Gianfranco Lauletta, Jody Corbelli, Elena<br />

Tenti, Floriana Facchetti, Cristina Della Corte, Stefano Tamberi, Oriana Nanni,<br />

Stefanu Cascinu, Dino Amadori, Giovanni Luca Frassineti<br />

other liver diseases (4%). There was no significant difference in<br />

frequency of miR-26a loss in patients with advanced versus less<br />

advanced HCC regardless of treatment (79% vs 74% in resection<br />

cases and 50% vs 60% in transplanted cases). Conclusions:<br />

Loss of miR-26a in the tumor tissue relative to surrounding<br />

liver tissue is a common phenomenon in HCC regardless the<br />

etiology or prior liver directed therapy but it was not associated<br />

with an aggressive tumor phenotype. More <strong>studies</strong> are needed<br />

to understand the biological significance of this observation.<br />

Disclosures:<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

The following authors have nothing to disclose: Sarah C. Nabinger, Samer Gawrieh,<br />

Maaz B. Badshah, Sandra K. Althouse, Smitha Marri, Sangbin Lee, Smiti S.<br />

Sahu, Susan Perkins, Romil Saxena, Janaiah Kota<br />

436<br />

MicroRNA-26 expression is significantly lower in hepatocellular<br />

carcinoma relative to surrounding liver tissue<br />

in humans with chronic liver disease<br />

Sarah C. Nabinger 1 , Samer Gawrieh 3 , Maaz B. Badshah 3 , Sandra<br />

K. Althouse 2 , Smitha Marri 3 , Sangbin Lee 1 , Smiti S. Sahu 1 ,<br />

Susan Perkins 2 , Romil Saxena 4 , Naga P. Chalasani 3 , Janaiah<br />

Kota 1 ; 1 Medical and Molecular Genetics, Indiana University<br />

School of Medicine, Indianapolis, IN; 2 Biostatistics, Indiana University<br />

School of Medicine, Indianapolis, IN; 3 of Gastroenterology/<br />

Hepatology, Indiana University School of Medicine, Indianapolis,<br />

IN; 4 Pathology and Laboratory Medicine, Indiana University<br />

School of Medicine, Indianapolis, IN<br />

Background & Aim: Hepatocellular Carcinoma (HCC) is the<br />

third leading cause of cancer-related deaths worldwide. We<br />

have previously shown that loss of microRNA-26a (miR-26)<br />

is common in murine HCC and systemic delivery of miR-26a<br />

suppressed liver tumor development in a mouse model of HCC.<br />

In this study, we investigated if miR-26 loss occurs with HCC<br />

in a large cohort of patients with underlying liver disease and<br />

assessed its correlation with advanced tumor features. Methods:<br />

We identified 136 paired tumor and normal adjacent<br />

liver tissues from HCC patients undergoing tumor resection<br />

(n=87) or liver transplantation (n=49) at Indiana University<br />

Medical Center. Total RNA was isolated from formalin fixed<br />

paraffin embedded tumor and normal adjacent liver samples,<br />

and subjected to quantitative PCR analysis to estimate miR-26a<br />

levels in these patients. The prevalence of miR-26 loss, defined<br />

by a relative fold change


430A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

438<br />

Sorafenib Therapy in HIV-infected Patients with Hepatocellular<br />

Carcinoma (HCC)<br />

Luciana Kikuchi 1 , Caitlin C. Citti 2,3 , Ting-Yi Chen 4 , Heather L.<br />

Platt 3 , David E. Kaplan 5 , Meritxell Ventura-Cots 6 , Mamta K. Jain 4 ,<br />

Eugenia Vispo 7 , Jorge Daruich 8 , Marina Nunez 9 , Pablo Barreiro 7 ,<br />

Beatriz Minguez 6 , Ayse Aytaman 10 , David Rimland 11 , Norbert<br />

Bräu 3,2 ; 1 Gastroenterology, Universidade de São Paulo, São<br />

Paulo, Brazil; 2 Infectious Diseases and Liver Diseases, Icahn School<br />

of Medicine at Mount Sinai, New York, NY; 3 James J. Peters VA<br />

Medical Center, Bronx, NY; 4 University of Texas Southwestern<br />

Medical Center, Dallas, TX; 5 University of Pennsylvania, Philadelphia,<br />

PA; 6 Hospital Universitario Vall d’Hebron, Barcelona, Spain;<br />

7 Hospital Carlos III, Madrid, Spain; 8 Universidad de Buenos Aires,<br />

Buenos Aires, Argentina; 9 Wake Forest University, Winston-Salem,<br />

NC; 10 VA New York Harbor HCS, Brooklyn, NY; 11 Atlanta VA<br />

Medical Center, Atlanta, GA<br />

BACKGROUND: Sorafenib is the only systemic chemotherapy<br />

for HCC shown to decrease all-cause mortality. Little is known<br />

about its effect in HIV-infected patients with HCC. METHODS:<br />

HIV-infected patients with HCC who received sorafenib (with or<br />

without TACE) were retrospectively identified from 1992-2013<br />

in 38 centers in 8 countries. They were compared to HIV-infected<br />

patients from the same cohort who received no effective<br />

HCC therapy. RESULTS: Compared to 132 untreated patients,<br />

the 22 patients receiving sorafenib were of similar age (51 vs.<br />

53 years), but tended to have less frequently alcohol abuse<br />

(18% vs. 38%, p=0.079), were diagnosed through HCC<br />

screening more often (59 vs. 30%, p=0.009), and had lower<br />

mean Child-Pugh scores (6.6 vs. 7.8, p=0.001). Sorafenib<br />

patients had lower mean HIV viral loads (1.7 vs. 2.7 log10<br />

copies/ml, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 431A<br />

dictors of RFS in overall patients include cirrhosis (hazard ratio<br />

(HR)=2.337, p=0.006) and AFP level >20 ng/mL (HR=1.896,<br />

p=0.026). The median RFS in NUC secondary prevention failure<br />

and tertiary prevention groups were 54 and 36 months,<br />

respectively (p=0.467). The 1-, 2- and 5-year RFS rates were<br />

85.3%, 72.7%, 46.1% in the NUC secondary prevention failure<br />

group, and 76.4%, 61.3%, 46.6% in tertiary prevention<br />

group, respectively. Baseline viral loads, HBsAg levels and<br />

virological response were not associated with RFS. In the subgroup<br />

patients without cirrhosis, NUC secondary prevention<br />

failure group had a trend of better RFS (p=0.057), whereas<br />

the RFS was comparable between the two group in patients<br />

with cirrhosis. Conclusions Continue NUC treatment has the<br />

potential to reduce the risk of HCC recurrence in non-cirrhotic<br />

patients despite secondary prevention had failed. Closely monitoring<br />

is recommended for HBV-HCC patients with cirrhosis<br />

after curative surgical resection even under NUC treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: I-Cheng Lee, Yi-Hsiang Huang,<br />

Gar-Yang Chau, Teh-Ia Huo, Chien-Wei Su, Han-Chieh Lin<br />

440<br />

A Pilot Safety Study of Nature Killer Cells From Sibship<br />

to Treat the Recurrence of Hepatocellular Carcinoma<br />

After Liver Transplantation<br />

Guo-Ying Wang, Yang Yang, Gui-Hua Chen; Liver Transplantation<br />

Center, the Third Affiliated Hospital, Sun Yat-sen University, Organ<br />

Transplantation Research Institute, Guangzhou, China<br />

Objective: Nature Killer (NK) Cells have been thought to play<br />

a pivotal role in innate immunity. However, the safety and efficacy<br />

of NK cells for the treatment of hepatocellular carcinoma<br />

(HCC) recurrence after liver transplantation (LT) is unknown. In<br />

this study, we investigated whether the injection of activated<br />

NK cells (CD3 - CD56 + cells) from the sibship with the same<br />

blood type is safe for the treatment of HCC recurrence after LT.<br />

Methods: Six patients with HCC recurrence after LT were eligible<br />

and enrolled in this study. The patients received injections<br />

of 5 × 10 9 NK cells from the sibship with the same blood type<br />

4 times at a frequency of once every two weeks. Lymphocytes<br />

were extracted from the peripheral blood mononuclear cells<br />

and cultured with IL-2 and other cytokines for 2 weeks. The<br />

purity of lymphocytes was assessed by flow cytometric analysis,<br />

and only the CD3 - CD56 + cells greater than 70% were<br />

used. The adverse effects, laboratory tests, overall survival<br />

were assessed. Results: No serious adverse effects, laboratory<br />

abnormalities were identified as related to the treatment of<br />

NK cells. Fix patients were alive in the follow-up period of 26<br />

months while one died of liver failure due to tumor progression<br />

after 2 months of NK cells infusion. There was no graft-versushost<br />

disease in all 6 patients during follow-up. Conclusion:<br />

In conclusion, we have demonstrated for the first time to our<br />

knowledge that NK cells from sibship with the same blood type<br />

can be used safely as adoptive immunotherapy for the treatment<br />

of HCC recurrence after liver transplantation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Guo-Ying Wang, Yang Yang,<br />

Gui-Hua Chen<br />

441<br />

Serological markers of extracelllular matrix remodelling<br />

for the diagnosis of HCC and the evaluation of tumour<br />

pathogenesis<br />

Roslyn Vongsuvanh 1 , Jacob George 1 , David van der Poorten 1 ,<br />

Morten A. Karsdal 2 , Diana J. Leeming 2 ; 1 Storr Liver Unit, Westmead<br />

Millenium Institute, Westmead Hospital, Westmead, NSW,<br />

Australia; 2 Nordic Bioscience, Fibrosis Biology and Biomarkers,<br />

Herlev, Denmark<br />

Liver fibrosis is associated with HCC, however, whether markers<br />

of matrix turnover may be exploited for HCC evaluation,<br />

remains poorly understood. Fragments of the extracellular<br />

matrix (ECM) are released systemically during the process of<br />

hepatic fibrosis. In this study, we assessed protein fingerprint<br />

markers of the ECM in serum of patients with HCC compared<br />

to cirrhosis, chronic liver disease and healthy controls. We<br />

determined the diagnostic performance of these markers for<br />

HCC in order to gain insight into the mechanisms linking matrix<br />

turnover to HCC diagnosis. Methods Plasma from 86 HCC<br />

patients, age and sex-matched to 86 cirrhotics, 86 hepatitis B<br />

(HBV) non-cirrhotics and 86 healthy controls, were collected in<br />

a cross-sectional study. ECM markers of matrix metaloproteinases<br />

(MMP)-2, 7 or -9, derived collagen turnover of type I, III,<br />

IV, VI (C1M, C3M, C4M2, C6M) and elastin (ELM7), as well<br />

as formation of type III and IV collagen (Pro-C3, P4NP 7S) were<br />

measured by ELISA. Results Mean plasma Pro-C3 levels were<br />

significantly higher in HCC (33.12 ng/ml) than in cirrhotics<br />

(22.2 ng/ml), HBV non-cirrhotics (12.18 ng/ml) and healthy<br />

controls (9.03 ng/ml) (p


432A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

442<br />

Hepatocellular Carcinoma Surveillance Is Associated<br />

With Improved Survival And Can Be Targeted To High<br />

Risk Populations<br />

Thai Hong 1 , Paul Gow 2,10 , Michael A. Fink 3,11 , Anouk T. Dev 4 ,<br />

Stuart K. Roberts 5 , Amanda J. Nicoll 6,7 , John Lubel 6,12 , Ian Kronborg<br />

8 , Niranjan J. Arachchi 8 , Marno C. Ryan 1,10 , William W.<br />

Kemp 5 , Virginia H. Knight 4 , Helen Farrugia 9 , Vicky Thursfield 9 ,<br />

Paul V. Desmond 1,10 , Alex J. Thompson 1,10 , Sally Bell 1 ; 1 Gastroenterology,<br />

St Vincent’s Hospital, Melbourne, Melbourne, VIC,<br />

Australia; 2 Gastroenterology, Austin Hospital, Melbourne, VIC,<br />

Australia; 3 Surgery, Austin Hospital, Melbourne, VIC, Australia;<br />

4 Gastroenterology & Hepatology, Monash Medical Centre, Melbourne,<br />

VIC, Australia; 5 Gastroenterology & Hepatology, The<br />

Alfred Hospital, Melbourne, VIC, Australia; 6 Gastroenterology &<br />

Hepatology, Eastern Health, Melbourne, VIC, Australia; 7 Gastroenterology<br />

& Hepatology, The Royal Melbourne Hospital, Melbourne,<br />

VIC, Australia; 8 Gastroenterology & Hepatology, Western<br />

Hospital, Melbourne, VIC, Australia; 9 Victorian Cancer Registry,<br />

Cancer Council Victoria, Melbourne, VIC, Australia; 10 Medicine,<br />

The University of Melbourne, Melbourne, VIC, Australia; 11 Surgery,<br />

The University of Melbourne, Melbourne, VIC, Australia;<br />

12 Eastern Health Clinical School, Monash University, Melbourne,<br />

VIC, Australia<br />

Background: Hepatocellular carcinoma (HCC) incidence is heterogeneously<br />

distributed worldwide and is rising rapidly in<br />

many developed countries where the complete range of therapeutic<br />

options is available. Local epidemiology is important<br />

to accurately characterise the disease and allow strategies to<br />

effectively manage the problem. We performed the first population-based<br />

study in Australia to capture a clinical cohort and<br />

study the effect of aetiology, ethnicity and advanced therapeutic<br />

options on survival outcomes. Method: Incident cases of<br />

HCC (defined by AASLD diagnostic criteria or histology) were<br />

prospectively identified over a 12-month period (1 July, 2012<br />

to 30 June, 2013) from the population of Melbourne, Australia.<br />

Cases were captured from multiple sources including admissions<br />

and outpatient attendances to any of Melbourne’s seven<br />

tertiary hospitals, gastroenterology department, radiology,<br />

pathology and pharmacy databases as well as those registered<br />

by the Victorian Cancer Registry. Patients were followed up<br />

for at least 24 months with the primary endpoint being overall<br />

survival. Results: There were 272 incident cases identified. The<br />

major risk factors for liver disease were hepatitis C virus (HCV)<br />

infection (41%), alcohol (39%), and hepatitis B virus (HBV)<br />

infection (22%). Patients born overseas had up to 20 times the<br />

incidence rate of Australian-born patients. Regional variations<br />

of incidence within the study area correlated with percentage<br />

of overseas-born patients and HBV seroprevalence in those<br />

regions. Overall survival at 12 months was 57%. On multivariate<br />

analysis, the independent predictors of survival were age,<br />

the absence of cirrhosis, MELD score, AFP, tumour size, Barcelona<br />

Clinic Liver Clinic (BCLC) stage and being involved in an<br />

HCC surveillance program. Patients undergoing surveillance<br />

had significantly smaller tumour size (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 433A<br />

444<br />

Hepatic function and tumour burden are predictive<br />

of outcome in patients with hepatocellular carcinoma<br />

treated with transarterial chemoembolisation<br />

Martha Kirstein 1,2 , Nora Schweitzer 1,2 , Nazli Ay 1,2 , Christina<br />

Boeck 1,2 , Bernhard Meyer 1,3 , Thomas Rodt 1,3 , Wacker Frank 1,3 ,<br />

Michael P. Manns 1,2 , Arndt Vogel 1,2 ; 1 Medical School Hannover,<br />

Hannover, Germany; 2 Gastroenterology, Hepatology, Endocrinology,<br />

Hannover, Germany; 3 Radiology, Hannover, Germany<br />

Background: Hepatocellular carcinoma (HCC) is one of the<br />

most lethal and prevalent cancers worldwide. Transarterial<br />

chemoembolization (TACE) is commonly used to act locally in<br />

the intermediate disease stage. Early randomized trials and<br />

more recent reviews and meta-analyses reported improved survival<br />

rates of patients with unresectable lesions managed with<br />

TACE so that TACE has been accepted as the standard treatment<br />

for intermediate stage disease. Methods: In this study,<br />

we characterized 487 HCC patients from Hannover Medical<br />

School, Germany, treated with TACE and evaluated the<br />

asscociation of their characteristics with survival. Results: 487<br />

HCC patients were treated at least once with TACE between<br />

2000 and 2013. Most patients (91.4%) were treated with conventional<br />

TACE using doxorubicin, ciplatin and/or mitomycin<br />

and lipiodol either alone or in combination with degradable<br />

starch microspheres (DSM, Spherex). Only minorities received<br />

transarterial embolisation (3.5%), transarterial chemoperfusion<br />

(3.1%) or TACE with drug-eluting beads (1.6%) as first transarterial<br />

treatment. 387 patients (79.3%) received TACE as first<br />

HCC-specific therapy. The mean number of received TACEs<br />

for all patients was 2.24±1.95(1-17) with a mean interval of<br />

140±257(33-2376) days between the first and second TACE.<br />

19.7%, 11.3% and 9% of the patients were diagnosed with<br />

one, two and three HCC nodules, respectively, whereas the<br />

remaining patients were diagnosed with more than three nodules.<br />

The mean diameter of the largest lesion was 59±36mm<br />

(4-226). 28.9% of the patients had at least one lesion ≥70mm.<br />

Hepatic function was well preserved in most patients with a<br />

mean Child-Pugh-Score A (6 points) and ALBI grade 2. After<br />

repeated TACEs there was no significant decrease of liver<br />

function observed as assessed by both classification systems.<br />

However, values of cholinesterase (CHE) were significantly<br />

decreased indicating CHE as a sensitive marker for patients<br />

treated with TACE (p


434A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Overall survival after the diagnosis of HCC according to the HCV<br />

genotype.<br />

Disclosures:<br />

Raoel Maan - Consulting: AbbVie<br />

Adriaan J. van der Meer - Consulting: Gilead; Speaking and Teaching: MSD,<br />

Gilead<br />

Jordan J. Feld - Advisory Committees or Review Panels: Merck, Janssen, Gilead,<br />

AbbVie, Theravance, Bristol Meiers Squibb; Grant/Research Support: AbbVie,<br />

Boehringer Ingelheim, Janssen, Gilead, Merck<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

Jean-Francois Dufour - Advisory Committees or Review Panels: Bayer, BMS, Gilead,<br />

AbbVie, Novartis, Sillagen, Genfit<br />

Andres Duarte-Rojo - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Vital Therapies; Speaking and Teaching: Roche<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Bart J. Veldt - Board Membership: GSK, Janssen Therapeutics<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

The following authors have nothing to disclose: Frank Lammert, Bettina E. Hansen<br />

446<br />

Validation of the Metroticket calculator in liver transplant<br />

recipients that had a prior chemoembolization for<br />

hepatocellular carcinoma<br />

Hyung-Don Kim 1 , Ju Hyun Shim 2 , Seungbong Han 3 , Mi-Jung Jun 2 ,<br />

Yeonjung Ha 2 , Jihyun An 2 , Danbi Lee 2 , Kang Mo Kim 2 , Young-<br />

Suk Lim 2 , Han Chu Lee 2 , Young-Hwa Chung 2 , Yung Sang Lee 2 ,<br />

Dong Jin Suh 4 ; 1 Department of Internal Medicine, Asan Medical<br />

Center, University of Ulsan College of Medicine, Seoul, Korea<br />

(the Republic of); 2 Department of Gastroenterology, Asan Medical<br />

Center, University of Ulsan College of Medicine, Seoul, Korea (the<br />

Republic of); 3 Department of Applied statistics, Gachon University,<br />

Gyeonggi-do, Korea (the Republic of); 4 Department of Internal<br />

Medicine, Vievis Namuh Hospital, Seoul, Korea (the Republic of)<br />

Background & Aims: The Metroticket calculator provides a continuum<br />

of survival probabilities for transplanted hepatocellular<br />

carcinoma (HCC) patients on the basis of tumor number, largest<br />

tumor diameter, and presence of microvascular invasion<br />

(MVI). This method had not yet been validated however in HCC<br />

patients with a previous history of transarterial chemoembolization<br />

(TACE). We performed this validation in our current study<br />

in liver transplant (LT) recipients. Methods: We enrolled 142<br />

patients with arterial enhancing HCC(s) on dynamic images<br />

treated with TACE between 1997 and 2011 at Asan Medical<br />

Center who were subsequent LT recipients. Tumor parameters<br />

measured by the enhancement radiological method pre-LT<br />

or by explant pathology post-LT were incorporated into the<br />

Metroticket calculator. The calculator was validated in terms<br />

of calibration and discrimination capacities for the entire study<br />

population and for subgroups undergoing living donor-related<br />

LT and infected with hepatitis B virus. Results: Study subjects<br />

received a median of 2 TACE sessions (range, 1-18) prior to<br />

LT, and most of these patients received a living donor graft<br />

(97.2%) and had HBV infection (89.4%). Through images and<br />

explants, 121 (85.2%) and 102 (71.8%) of our study subjects<br />

met the Milan criteria, respectively. Seventeen patients<br />

(12%) had no pathologic viable nodule, and 15 (10.6%) cases<br />

showed microvascular invasion. The mean 3-year and 5-year<br />

survival rates predicted by the radiological model for all 142<br />

patients were 76.4% and 70.1%, respectively, and fell perfectly<br />

within the 95% confidence intervals (CI) of the observed<br />

estimates (72.8-86.2% and 68.6-83.2%, respectively). In the<br />

pathological model incorporating microvascular invasion, the<br />

anticipated 5 year survival rate of 120 patients with viable<br />

nodules on explants was 69.5%, which also was inside the<br />

95% CI of the actuarial rates (67.9-83.5%). The c-index as a<br />

measure of discriminatory power was 0.61 and 0.62, respectively,<br />

for the 3- and 5-year predictions in the radiological<br />

model, and 0.72 for the 5-year prediction in the pathological<br />

model. All corresponding findings were comparable for<br />

subgroups of living donor-LT recipients and hepatitis B virus<br />

infected patients. Conclusion: A Metroticket calculation based<br />

on explant data can accurately predict post-LT survival in HCC<br />

patients with prior TACE. Estimate-based predictions before LT<br />

using imaging may have poorer discriminatory power compared<br />

to calibration.<br />

Disclosures:<br />

Young-Suk Lim - Advisory Committees or Review Panels: Bayer Healthcare, Gilead<br />

Sciences; Grant/Research Support: Bayer Healthcare, BMS, Gilead Sciences,<br />

Novartis<br />

Han Chu Lee - Grant/Research Support: Medigen Biotechnology Co., Novartis,<br />

Roche, Bayer HealthCare, Bristol-Myers Squibb, INC research, Boehringer Ingelheim,<br />

Taiho Pharmaceutical Co., Yuhan Co.<br />

The following authors have nothing to disclose: Hyung-Don Kim, Ju Hyun Shim,<br />

Seungbong Han, Mi-Jung Jun, Yeonjung Ha, Jihyun An, Danbi Lee, Kang Mo<br />

Kim, Young-Hwa Chung, Yung Sang Lee, Dong Jin Suh<br />

447<br />

Retreatment with TACE: Transition of Model to Estimate<br />

Survival in Ambulatory Hepatocellular carcinoma<br />

patients helps decision<br />

Young Youn Cho 1 , Su Jong Yu 1 , June Sung Lee 2 , Jeong-Ju Yoo 1 ,<br />

Minjong Lee 1 , Donghyeon Lee 1 , Yuri Cho 1 , Eun Ju Cho 1 , Jeong-<br />

Hoon Lee 1 , Yoon Jun Kim 1 , Jung-Hwan Yoon 1 ; 1 Department of<br />

Internal Medicine and Liver Research Institute, Seoul National<br />

University College of Medicine, Seoul, Korea (the Republic of);<br />

2 Department of Internal Medicine, Ilsan Paik Hospital, Ilsan, Korea<br />

(the Republic of)<br />

Background/Aims: Transarterial chemoembolization (TACE) is<br />

a major therapeutic modality for patients with unresectable<br />

hepatocellular carcinoma (HCC), but prognosis is variable.<br />

Model to Estimate Survival in Ambulatory Hepatocellular carcinoma<br />

patients (MESIAH) was recently developed and validated<br />

as a survival model to predict prognosis of patients with<br />

HCC. We aimed to develop a novel index using MESIAH for<br />

predicting who can benefit from repeated TACE. Methods:


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 435A<br />

From 2005 to 2008, 783 patients initially treated with TACE<br />

in Seoul National University Hospital (Seoul, Korea) were<br />

included in this study. Among them, 356 patients received<br />

second TACE session (TACE-2) and 196 patients received<br />

third TACE session (TACE-3) for recurred HCC. We assessed<br />

MESIAH score at baseline and prior to each TACE session, and<br />

designated patients into MESIAH transition group if MESIAH<br />

score increased before each TACE session. Cox regression and<br />

survival analysis were performed. Results: The mean age was<br />

56.3 ± 10.3 and hepatitis B virus were present in 76.0%. There<br />

were 43 (12.1%) patients in the pre-TACE-2 MESIAH transition<br />

group. There were no significant differences of Model for Endstage<br />

Liver Disease score (8.27 vs 8.89, P=0.28) and baseline<br />

MESIAH score (5.14 vs 5.02, P=0.41) between MESIAH transition<br />

and non-transition groups, respectively. MESIAH transition<br />

group showed shorter overall survival than the MESIAH<br />

non-transition group (6.0 versus 16.0 months, respectively;<br />

hazard ratio, 1.73; 95% confidence interval, 1.20-2.41).<br />

Similar results were observed when the definition of MESIAH<br />

transition was applied before TACE-3. Conclusions: MESIAH<br />

transition is a simple method which can be implicated in practice.<br />

MESIAH transition identifies patients who are unsuitable<br />

for retreatment with TACE and external validation is warranted.<br />

Disclosures:<br />

Yoon Jun Kim - Grant/Research Support: Bristol-Myers Squibb, Roche, JW Creagene,<br />

Bukwang Pharmaceuticals, Handok Pharmaceuticals, Hanmi Pharmaceuticals,<br />

Yuhan Pharmaceuticals; Speaking and Teaching: Bayer HealthCare<br />

Pharmaceuticals, Gilead Science, MSD Korea, Yuhan Pharmaceuticals, Samil<br />

Pharmaceuticals, CJ Pharmaceuticals, Bukwang Pharmaceuticals, Handok Pharmaceuticals<br />

The following authors have nothing to disclose: Young Youn Cho, Su Jong Yu,<br />

June Sung Lee, Jeong-Ju Yoo, Minjong Lee, Donghyeon Lee, Yuri Cho, Eun Ju<br />

Cho, Jeong-Hoon Lee, Jung-Hwan Yoon<br />

448<br />

Survival of Patients with Advanced Hepatocellular Carcinoma<br />

on Sorafenib: Retrospective Analysis from a<br />

Single Asian Center<br />

Yeonjung Ha, Danbi Lee, Ju Hyun Shim, Young-Suk Lim, Han Chu<br />

Lee, Young-Hwa Chung, Yung Sang Lee, Kang Mo Kim; Asan<br />

Medical Center, Seoul, Korea (the Republic of)<br />

Background: Sorafenib is the current standard of care for<br />

patients with advanced hepatocellular carcinoma (HCC);<br />

however actual treatment pattern varies from region to region<br />

as well as among physicians. Therefore, we retrospectively<br />

investigated the practice pattern and compared the effect of<br />

sorafenib on survival and progression with other treatment<br />

strategies. Methods: We conducted a single center retrospective<br />

study involving patients of Barcelona Clinic Liver Cancer<br />

C stage HCC on sorafenib with or without other treatments.<br />

Baseline characteristics, pre and intratreatment variables, and<br />

features regarding sorafenib administration were evaluated.<br />

The primary outcome measure was the overall survival (OS),<br />

and time-to-progression (TTP) was also calculated according<br />

to each treatment strategy. Results: A total of 658 patients<br />

(mean age 54.5 years, 83.3% male) were analyzed, with 293<br />

initially treated with sorafenib, 236 treated with transarterial<br />

chemoembolization (TACE) at first then switched to sorafenib,<br />

and 129 who received a combination therapy of TACE and<br />

sorafenib. The baseline characteristics of patients at the time<br />

of initial sorafenib administration were comparable across the<br />

three groups, Tumor-Node-Metastasis stage was IV in 96.3%<br />

and liver function was Child-Pugh A in 77.1%. The indication<br />

for sorafenib administration was mostly distant metastasis<br />

in all groups, 88.5% in the initial treatment group, 89.7%<br />

in the switched group and 80.6% in the combination group<br />

(p=0.43). The average dosage of sorafenib was similar across<br />

the groups (mean, 661.6 mg, p=0.18), whereas the treatment<br />

duration was significantly shorter in the switched group.<br />

Patients were withdrawn from sorafenib mostly because of disease<br />

progression, while 12.8% stopped the medication due<br />

to adverse reactions, most commonly hand-foot skin reaction<br />

(43.0%). The median OS was 11.8 months in sorafenib alone<br />

as compared with 13.5 months in the switched group and<br />

16.2 months in the combination group (p=0.13). In a subgroup<br />

analysis of patients with portal vein invasion but not with distant<br />

metastasis, combination therapy was associated with better OS<br />

(p=0.001) and longer TTP (p=0.020). In a multivariate model,<br />

only combined treatment strategy was significantly associated<br />

with longer OS (odds ratio [OR], 0.33, p=0.007) and TTP (OR,<br />

0.40; p=0.014). Conclusion: Neither switched therapy (TACE<br />

to sorafenib) nor combined approach was associated with a<br />

longer OS than sorafenib alone in patients with BCLC C HCC.<br />

However, in a subset of patients with portal vein invasion but<br />

not with distant metastasis, combined treatment with TACE and<br />

sorafenib showed better survival.<br />

Disclosures:<br />

Young-Suk Lim - Advisory Committees or Review Panels: Bayer Healthcare, Gilead<br />

Sciences; Grant/Research Support: Bayer Healthcare, BMS, Gilead Sciences,<br />

Novartis<br />

Han Chu Lee - Grant/Research Support: Medigen Biotechnology Co., Novartis,<br />

Roche, Bayer HealthCare, Bristol-Myers Squibb, INC research, Boehringer Ingelheim,<br />

Taiho Pharmaceutical Co., Yuhan Co.<br />

The following authors have nothing to disclose: Yeonjung Ha, Danbi Lee, Ju Hyun<br />

Shim, Young-Hwa Chung, Yung Sang Lee, Kang Mo Kim<br />

449<br />

In patients with hepatocellular carcinoma (HCC) and<br />

macrovascular invasion (MVI) amenable to surgical<br />

resection, a survival similar to that observed on<br />

sorafenib can be achieved<br />

Charlotte E. Costentin 1 , Eric Vibert 2 , Françoise Roudot-Thoraval 3 ,<br />

Thomas Decaens 4 , Nathalie Ganne-Carrié 5 , Bernard Paule 2 ,<br />

Christian Letoublon 6 , Mélanie Chiaradia 7 , Julien Calderaro 8 ,<br />

Rene Adam 2 , Ivan Bricault 9 , Giuliana Amaddeo 1 , Daniel Cherqui<br />

2 , Olivier Seror 10 , Didier Samuel 2 , Ariane Mallat 1 , Christophe<br />

Duvoux 1 , Alexis Laurent 11 ; 1 Hepatology, Hôpital Henri Mondor,<br />

Creteil, France; 2 Hepatobiliary surgery, Paul Brousse Hospital,<br />

Villejuif, France; 3 Public Health, Henri Mondor Hospital, Creteil,<br />

France; 4 hepato-gastroenterology, CHU Grenoble, Grenoble,<br />

France; 5 Hepatology, Jean Verdier Hospital, Bondy, France; 6 Surgery,<br />

CHU Grenoble, Grenoble, France; 7 Radiology, Henri Mondor<br />

Hospital, Creteil, France; 8 Pathology, Henri Mondor Hospital,<br />

Creteil, France; 9 Radiology, CHU Grenoble, Grenoble, France;<br />

10 Radiology, Jean Verdier Hospital, Bondy, France; 11 Surgery,<br />

Henri Mondor Hospital, Creteil, France<br />

Introduction: A median survival of 8 months has been reported<br />

in the SHARP trial in HCC+MVI treated with Sorafenib. In<br />

contrast, surgical retrospective <strong>studies</strong> report overall survivals<br />

ranging from 12 to 20 month in this setting. The aim of this<br />

study was to describe characteristics and overall survival in<br />

HCC+MVI patients treated with surgery or sorafenib as the<br />

first treatment step. Methods: 142 patients with HCC+MVI,<br />

without extra-hepatic spread, treated either with surgical resection<br />

(group 1; n=75) or sorafenib (group 2; n=67) in 4 French<br />

centers as first-line treatment after diagnosis of MVI were retrospectively<br />

reviewed. Results: Compared to patients in group<br />

2, patients in group 1 were younger (58.3±11.7 vs 65.5±9.6<br />

years, p


436A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

7.5% of the patients (n=5) in group 2 died within 3 months<br />

post-sorafenib initiation (p=0.196). In group 1, 29 patients<br />

(45%) had additional therapy either adjuvant or for recurrence<br />

(including TACE, chemotherapy (CT), sorafenib, thermo-ablation,<br />

additional resection) and one was transplanted. In group<br />

2, 20 (30%) patients had additional therapy either in addition<br />

to sorafenib or as second line (including TACE, CT, radiotherapy<br />

or radioembolization). Median overall survival (OS) was<br />

12.0 months (IC95%:5.5-18.5) in group 1 and 12.9 months<br />

(IC95%: 8.3-17.5) in group 2 (p=0.61). Taking into account<br />

the imbalance in baseline characteristics, a propensity analysis<br />

was performed, identifying 57 patients with similar characteristics<br />

with a median OS of 7.2 months (IC95%: 3.7-10.6) in the<br />

“resection” group and 12.6 months (IC95%: 7.1-18.0) in the<br />

“sorafenib” group (p=0.89). In multivariate analysis, prognostic<br />

factors associated with survival in the whole population (resection<br />

+ sorafenib) were PT time 50 mm HR=1.69, p=0.046), MVI involving<br />

1st order branch or main trunk (HR=1.948, p=0.042)<br />

and additionnal therapy (HR=0.500, p=0.024). Median OS<br />

in the resection+additional therapy group of patients (n=29)<br />

was 25.2 [8.8-41.6] months and 24.3 [15.9-32.7] months<br />

in the sorafenib+additional therapy group (n=20), (p=0.82).<br />

Conclusion: In patients with HCC+ MVI amenable to surgical<br />

resection, a survival similar to that observed on standard of<br />

care, sorafenib, can be achieved. Palliative surgery may be<br />

a valuable alternative in this setting. A multimodal treatment<br />

strategy was associated with prolonged survival.<br />

Disclosures:<br />

Françoise Roudot-Thoraval - Advisory Committees or Review Panels: Roche; Consulting:<br />

LFB biomedicaments; Speaking and Teaching: gilead, Janssen, BMS,<br />

Roche<br />

Nathalie Ganne-Carrié - Advisory Committees or Review Panels: Roche; Speaking<br />

and Teaching: BMS, Gilead, Bayer<br />

Olivier Seror - Board Membership: Bayer Healthcare; Consulting: Olympus;<br />

Speaking and Teaching: Angiodynamics<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

Christophe Duvoux - Advisory Committees or Review Panels: Novartis, Roche,<br />

Novartis, Roche, Novartis, Roche, Novartis, Roche; Speaking and Teaching:<br />

Astellas, Astellas, Astellas, Astellas<br />

The following authors have nothing to disclose: Charlotte E. Costentin, Eric Vibert,<br />

Thomas Decaens, Bernard Paule, Christian Letoublon, Mélanie Chiaradia,<br />

Julien Calderaro, Rene Adam, Ivan Bricault, Giuliana Amaddeo, Daniel Cherqui,<br />

Ariane Mallat, Alexis Laurent<br />

450<br />

Prediction of liver decompensation and survival after<br />

TACE for hepatocellular carcinoma. Is there a role for<br />

non-invasive markers of fibrosis?<br />

Jean-Baptiste Hiriart 1 , François Franques 1 , Julien Vergniol 1 , Christophe<br />

Cassinotto 2 , Faiza Chermak 1 , Bruno Lapuyade 2 , Juliette<br />

Foucher 1 , Wassil Merrouche 1 , Jean-Frédéric Blanc 3 , Victor de<br />

Ledinghen 1 ; 1 Hepatology, Hôpital Haut-Lévêque, Pessac, France;<br />

2 Diagnostic and Interventional Imaging, Hôpital Haut-Lévêque, Pessac,<br />

France; 3 Hepatology, Hôpital Saint-André, Bordeaux, France<br />

Introduction: Non-invasive markers of liver fibrosis are useful to<br />

predict survival and liver-related events in patients with chronic<br />

liver disease. The aim of this study was to investigate the performance<br />

of these markers to predict hepatic decompensation and<br />

death after transarterial chemoembolization (TACE) for hepatocellular<br />

carcinoma (HCC). Patients and methods: From 2008<br />

to 2012, all consecutive patients undergoing a first TACE for<br />

HCC with a FibroScan and serum biomarkers (Fibrotest, APRI,<br />

Hepascore, FIB-4, AST/ALT ratio, NAFLD score) performed at<br />

baseline were included. Liver decompensation was defined as<br />

the occurrence of at least one of the following events within 6<br />

weeks after TACE: ascites, increase of bilirubin levels above 50<br />

mmol/L, acute kidney injury. Child-Pugh score and MELD score<br />

were calculated the day before TACE and one month after.<br />

Incident cases of death were prospectively collected. Results:<br />

103 patients were included (mean age 66 years, 85% men).<br />

In univariate analysis, factors associated with decompensation<br />

were albumin level 200 ng/ml (OR=2.51, CI95% 1.24-5.08, p=0.01)<br />

and albumin


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 437A<br />

for 70.9% and 12.6% of the total HCC cases (1292 patients)<br />

respectively. Out of the total cohort enrolled from 1980 to<br />

2005, the ratio of cryptogenic to CHB patients was 1:6.7;<br />

while from 2006 to 2015, the ratio of cryptogenic to CHB<br />

patients has increased significantly to 1:3.9. Both groups were<br />

comparable with regards to gender, Chinese ethnicity, smoking<br />

status, Child’s score, stage of HCC, portal vein invasion,<br />

AFP and platelet count. Relative to CHB group, cryptogenic<br />

HCC patients were older (68 versus 60 years old; p


438A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Atsuo Takigawa, Ryotaro Sakamori,<br />

Tomohide Tatsumi, Takayuki Yakushijin, Tadashi Kegasawa, Yuki Makino,<br />

Seiichi Tawara, Yoshinobu Saito, Yoshiki Onishi, Satoshi Tanaka, Kunimaro<br />

Furuta, Minoru Shigekawa, Naoki Hiramatsu<br />

Disclosures:<br />

The following authors have nothing to disclose: Li Huang, Kunsong Zhang, Wei<br />

Chen, Mengjun Hou, Yuyan Han, Fanyin Meng, Sharon DeMorrow, Xiaoyu Yin,<br />

Jiaming Lai, Bingyan Tan, Lijian Liang<br />

454<br />

Dysregulated peripheral and local endocannabinoid<br />

system promote tumor pathogenesis in biliary tract cancers<br />

Li Huang 1 , Kunsong Zhang 1 , Wei Chen 1 , Mengjun Hou 2 , Yuyan<br />

Han 3 , Fanyin Meng 4 , Sharon DeMorrow 5 , Xiaoyu Yin 1 , Jiaming<br />

Lai 1 , Bingyan Tan 2 , Lijian Liang 1 ; 1 Department of Pancreatobiliary<br />

Surgery, The First Affiliated Hospital, Sun Yat-sen University,<br />

Guangzhou, China; 2 Department of Nutrition, School of Public<br />

Health, Sun Yat-Sen University, Guangzhou, China; 3 Department<br />

of Medicine, Baylor and Scott & White Digestive Disease Research<br />

Center, Texas A&M HSC College of Medicine and Scott & White<br />

Hospital, Temple, TX; 4 Department of Medicine; Department of<br />

Research, Scott & White Digestive Disease Research Center; Texas<br />

A&M Health Science Center College of Medicine; Central Texas<br />

Veterans Health Care System, Temple, TX; 5 Department of Internal<br />

Medicine, Texas A&M Health Science Center College of Medicine;<br />

Baylor Scott & White Health Digestive Disease Research Center;<br />

Central Texas Veterans Health Care System, Temple, TX<br />

Background: The endocannabinoid system (ECS) is dysregulated<br />

in various liver diseases. Biliary tract cancers (BTCs)<br />

encompass gallbladder carcinoma, intrahepatic, perihilar and<br />

distal cholangiocarcinoma. Previously we have shown that the<br />

two major endocannabinoids (ECs), anandamide (AEA) and<br />

2-arachidonoyl glycerol (2-AG) exert opposing effects on proliferation<br />

of BTCs cell lines. However, regulation and clinical<br />

significance of ECS in BTCs remain inconclusive. Method: Levels<br />

of AEA and 2-AG in preoperative peripheral plasma and<br />

bile in 22 BTCs patients and 8 patients with benign biliary<br />

diseases, and paired resection samples in 15 BTCs patients<br />

were quantitated by gas chromatography/electron ionization<br />

(EI)-mass spectrometry (GC/MS). Expression and activity of<br />

the enzymes involved in ECs biosynthesis (phospholipase D<br />

for AEA and diacylglycerol lipases (DAGLs, including DAGLα<br />

and DAGLβ) for 2-AG) and ECs degradation (fatty acid amide<br />

hydrolase-1 (FAAH-1) for AEA and monoaclyglycerol lipase<br />

(MAGL) for 2-AG) were also analyzed. Levels of ECs were correlated<br />

with patients’ clinicopathologic features. Results: Level<br />

of AEA was lower in plasma and bile (median [interquartile<br />

range]: 1.804 pmol/mL [0.502 to 5.004 pmol/mL] vs 10.150<br />

pmol/mL [2.684 to 17.491 pmol/mL]) in BTCs than in benign<br />

biliary diseases, though the former didn’t reach significance.<br />

AEA was significantly lower in BTCs cancer nest than in adjacent<br />

normal tissue (22.01 fmol/mg [16.55 to 53.61 fmol/mg]<br />

vs 58.68 fmol/mg [25.36 to 68.97 fmol/mg]). Level of 2-AG<br />

was upregulated in plasma (180.0 pmol/mL [140.7 to 283.8<br />

pmol/mL] vs 42.3 pmol/mL [25.6 to 148.9 pmol/mL]) and<br />

bile in BTCs, though the latter didn’t reach significance. 2-AG<br />

was significantly higher in BTCs cancer nest than in adjacent<br />

normal tissue (1.967 pmol/mg [1.439 to 5.433 pmol/mg] vs<br />

1.101 pmol/mg [0.358 to 1.473 pmol/mg]). Upregulated<br />

expressions and activities of DAGLs, FAAH-1 and MAGL were<br />

demonstrated in BTCs. Lower cancer nest level of AEA, higher<br />

plasma and cancer nest level of 2-AG were correlated with<br />

more advanced tumor stage in BTCs. Conclusion: ECS is dysregulated<br />

in BTCs and it may foster a tumor promoting microenvironment.<br />

Modulation of ECS could be regards as a promising<br />

therapeutic strategy for BTCs.<br />

455<br />

Prospective cohort study for evaluating clinical effects<br />

and safety of intra-arterial infusion therapy of Cisplatin<br />

suspension in lipiodol combined with 5-fluorouracil and<br />

sorafenib for advanced hepatocellular carcinoma with<br />

macroscopic vascular invasion, without extra-hepatic<br />

spread<br />

Niizeki Takashi, Masahito Nakano, Takuji Torimura; Gastroenterology,<br />

Kurume University, school of medicine, Kurume, Japan<br />

Purpose Sorafenib is the standard first line treatment for patients<br />

with BCLC stage C hepatocellular carcinoma (HCC). However<br />

sorafenib monotherapy confers less than three months of<br />

actual survival benefit in both Western and Asian populations.<br />

Hepatic arterial infusion chemotherapy (HAIC) is recognized<br />

as a useful therapeutic option for advanced HCC in Japan, and<br />

response to HAIC is known as an important prognostic factor.<br />

While sorafenib has been proven to improve the prognosis in<br />

HCC patients with macroscopic vascular invasion (MVI), it is<br />

still unknown which is better, Sorafenib or HAIC. The aim of<br />

this multicenter none- randomized prospective cohort study was<br />

to investigate the efficacy and safety of HAIC compared with<br />

Sorafenib in patients with advanced HCC with MVI, without<br />

extra-hepatic spread (EHS) and Child-Pugh class A disease.<br />

Method: This study was conducted from April 2009 to March<br />

2014, total 64 HCC patients were registered. 44 were treated<br />

with HAIC, 20 were treated with Sorafenib. HAIC regimen<br />

comprised a combination of 50 mg fine powder formulation of<br />

Cisplatin in 5-10 ml lipiodol and contimuous infusion of 5-fluorouracil<br />

(1,500 mg/5 days). The primary efficacy endpoint<br />

was progression-free survival (PFS), while the secondary endpoints<br />

were Median survival time (MST), tumor response rate,<br />

and safety. PFS and MST were estimated by Kaplan- Meier<br />

method. Therapeutic effect was evaluated according to RECIST.<br />

Results: There were no statistica differences in clinical factors<br />

between two groups. PFS in HAIC and Sorafenib was 9.3<br />

months and 4.1 months respectively. (P = 0.002, 95% CI,<br />

0.221- 0.713) CR or PR rate in HAIC and Sorafenib was 71%<br />

and 10%.(P 0.001, 95% CI, 4.32- 106.09) MST in HAIC and<br />

Sorafenib was 30.0 months and 13.0 months respectively. (P<br />

= 0.016, 95% CI, 0.219- 0.854) Severe adverse events were<br />

observed in 5 patients. In Sorafenib group, 2 had hepatic<br />

failure, in HAIC group, 2 had hepatic failure, 1 had pseudo<br />

aneurysm. In HAIC group, treatment- related mortality was not<br />

observed. By multivariate analysis, independent predictor of<br />

survival were therapeutic effect (CR or PR, P= 0.009, 95% CI,<br />

0.220- 0.752), Child- Pugh score (score 5, P= 0.022, 95%<br />

CI, 0.191- 0.752), grade of portal vein invasion (trunk, P=<br />

0.002, 95% CI, 0.118- 0.614), and independent predictor of<br />

therapeutic effect was therapeutic regimen (HAIC, P 0.0001).<br />

Conclusion: In HAIC group, PFS was 9.3 months, MST was<br />

30.0 months and response rate was 74%, and these results<br />

were superior to Sorafenib. In conclusion, this regimen should<br />

be the first choice for patients with advanced HCC with MVI,<br />

without EHS and Child Pugh A disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Niizeki Takashi, Masahito<br />

Nakano, Takuji Torimura


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 439A<br />

456<br />

Impact of Hepatic p62 Overexpression on Survival of<br />

Patients with Hepatocellular Carcinoma<br />

Yasuji Komorizono 1 , Masaki Kitazono 2 , Sadao Tanaka 3 , Kazuhisa<br />

Nakashima 1 , Toshihiko Shibatou 1 , Katsumi Sako 1 ; 1 Hepatology,<br />

Nanpuh Hospital, Kagoshima, Japan; 2 Digestive Surgery, Nanpuh<br />

Hospital, Kagoshima, Japan; 3 Pathology, Nanpuh Hospital,<br />

Kagoshima, Japan<br />

Purpose: p62 is an adaptor or a scaffolding substrate between<br />

the ubiquitin–proteasome and the autophagy–lysosome pathways.<br />

p62 overexpression has been demonstrated to play a<br />

critical role in tumorigenesis of various malignancies, including<br />

hepatocellular carcinoma (HCC); however, the association<br />

between p62 overexpression and survival in HCC is currently<br />

unknown. Our purpose was to clarify whether overexpression<br />

of p62 impacts survival of patients with HCC. Methods:<br />

Between May 2003 and July 2013, we retrospectively identified<br />

78 patients with primary HCC (51 men and 27 women<br />

aged 34 to 81 years; mean, 68.4 years) who had undergone<br />

surgical resection and enrolled them in this study. We performed<br />

immunohistochemical analyses of their p62 expression<br />

and correlated the findings with annotated clinicopathologic<br />

data. Fifty-four patients (69%) were positive for hepatitis virus C<br />

antibody and 11 (14%) for hepatitis B antigen; the remaining<br />

13 patients (17%) were negative for both markers. We performed<br />

immunohistochemical analyses with antibody against<br />

p62 on surgically resected specimens and compared the overall<br />

survival of the p62-positive and -negative groups using the<br />

Kaplan–Meier method. We assessed differences between these<br />

two groups using the log-rank test. A Cox proportional hazards<br />

model was used to evaluate the interactions between the baseline<br />

characteristics and the impact of p62 overexpression on<br />

overall survival. Results:We detected p62 overexpression in 54<br />

patients (69%). Overall, 38 deaths (49%) occurred during follow-up<br />

(26 in the p62-positive and 12 in the -negative group).<br />

In the baseline characteristics, there was significant difference<br />

in the histological differentiation (poorly versus well and moderately)<br />

between the p62 positive group and the p62 negative<br />

group (p = 0.047). The overall survival rates at 5, 7, and 10<br />

years were 56%, 45%, and 23 % in the p62 positive group<br />

and 86%,58%, and 46% in the p62 negative group, respectively.<br />

The log-rank test revealed that the overall survival was<br />

significantly shorter in the p62 positive group than in the p62<br />

negative group (p = 0.046). Multivariate analyses showed that<br />

only p62 overexpression (HR, 2.29; 95% CI, 1.03–5.12; P =<br />

0.043) and main tumor size (HR, 1.17; 95% CI, 1.0–1.35; p<br />

= 0.044) were independently associated with overall survival.<br />

Conclusion: The results of the current study indicated that overexpression<br />

of p62 was closely associated with overall survival<br />

in patients with HCC. P62 could be used as a prognostic biomarker<br />

in patients with HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yasuji Komorizono, Masaki Kitazono,<br />

Sadao Tanaka, Kazuhisa Nakashima, Toshihiko Shibatou, Katsumi Sako<br />

457<br />

Improved Survival of Hepatocellular Carcinoma (HCC)<br />

in HIV/Hepatitis B Virus (HBV)-Coinfected Patients Who<br />

Are Diagnosed Through HCC Screening<br />

Maaz B. Badshah 1 , Meritxell Ventura-Cots 2 , Caitlin C. Citti 3 , Luciana<br />

Kikuchi 4 , Sonja Marcus 5 , Beatriz Minguez 6 , Ting-Yi Chen 7 ,<br />

Maria D. Hernandez 8 , Judith Aberg 3 , Myron E. Schwartz 9 , Douglas<br />

Dieterich 3 , Norbert Bräu 5,3 ; 1 Indiana University School of<br />

Medicine, Indianapolis, IN; 2 Hospital Universitari Vall ‘Hebron,<br />

Barcelona, Spain; 3 Dept. of Medicine, Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 4 Universidade de São Paulo, São<br />

Paulo-SP, Brazil; 5 James J. Peters VA Medical Center, Bronx, NY;<br />

6 Hospital Universitari Vall d’Hebron, Barcelona, Spain; 7 University<br />

of Texas Southwestern Medical Center, Dallas, TX; 8 University of<br />

Miami School of Medicine, Miami, FL; 9 Dept. of Surgery, Icahn<br />

School of Medicine at Mount Sinai, New York, NY<br />

BACKGROUND: Current recommendations for HCC screening<br />

in patients chronically infected with the hepatitis B virus<br />

(HBV) are based on a randomized controlled trial from China.<br />

However, no data are available on the effectiveness of HCC<br />

screening in HIV/HBV-coinfected patients. METHODS: HIV/<br />

HBV-coinfected patients with HCC were retrospectively identified<br />

from 1992-2013 in 38 centers in 8 countries. Patients<br />

were considered screened if they presented with an abnormal<br />

alpha-fetoprotein level or imaging study, and not screened<br />

if they presented with symptoms. RESULTS: Among 74 HIV/<br />

HBV-coinfected patients with HCC, 40 (60%) were screened.<br />

Compared to 34 unscreened patients, screened patients had<br />

similar mean age (49.4 vs. 50.7 years). However, screened<br />

patients had a lower mean Child-Pugh score (5.8 vs. 7.4,<br />

p=0.002), had a smaller median tumor size (3.6 vs. 8.7 cm,<br />

p=0.001), had a lower median alpha-fetoprotein level (107<br />

vs. 1,422 ng/ml, p=0.010), and more commonly met Milan<br />

criteria for liver transplantation (44% vs. 7%, p=0.001). They<br />

presented less frequently with portal vein thrombosis (23% vs.<br />

47%, p=0.031) and with advanced Barcelona-Clinic-Liver-Cancer<br />

stages C+D (36% vs. 78%, p=0.001), and more often<br />

received effective HCC therapy (80% vs. 27%, p< 0.001). With<br />

adjustment for lead time of 10.2 months, screened patients had<br />

a longer median survival (89 vs. 3.7 months, p=0.010, log<br />

rank). The actuarial 1-year survival rate was 70% in screened<br />

patients and 27% in unscreened patients. In multi-variable Cox<br />

proportional hazard analysis, screening was an independent<br />

predictor of survival (hazard ratio for death, 0.42; 95% confidence<br />

interval, 0.19 – 0.93; p=0.033), together with effective<br />

HCC therapy and Child-Pugh score. CONCLUSIONS: In<br />

HIV/HBV-coinfected patients with HCC, a diagnosis through<br />

screening was associated with earlier HCC stages, more HCC<br />

therapy, and independently predicted better survival.


440A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

5.34, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 441A<br />

respectively. All searches and data extraction were performed<br />

independently by 2 authors. Included <strong>studies</strong> were published in<br />

2000 or after and had ≥100 patients undergoing surveillance<br />

as defined above. Analysis was via random-effects models.<br />

Results: A total of 14 <strong>studies</strong> with 15,429 (12,828 CHB; 2,601<br />

cirrhosis) met inclusion criteria. Overall adherence was 61%<br />

(95% CI 56-67%). Cirrhosis patients had significantly higher<br />

adherence than CHB patients, 65% (95% CI 61-69%) vs. 58%<br />

(95% CI 49-67%), p


442A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

461<br />

Etiological difference in body composition of patients<br />

with hepatocellular carcinoma<br />

Ryosuke Tateishi 1 , Naoto Fujiwara 1 , Hayato Nakagawa 1 , Yotaro<br />

Kudo 1 , Ryo Nakagomi 1 , Mayuko Kondo 1 , Takuma Nakatsuka 1 ,<br />

Tatsuya Minami 1 , Masaya Sato 1 , Koji Uchino 1 , Kenichiro Enooku 1 ,<br />

Yuji Kondo 1 , Yoshinari Asaoka 1 , Yasuo Tanaka 1 , Kyoji Moriya 1 ,<br />

Shuichiro Shiina 2 , Kazuhiko Koike 1 ; 1 Department of Gastroenterology,<br />

The University of Tokyo Hospital, Tokyo, Japan; 2 Department<br />

of Gastroenterology, Juntendo University, Tokyo, Japan<br />

Background: We have reported that sarcopenia, intramuscular<br />

fat deposition, and visceral adiposity indicate poor prognosis<br />

of patients with hepatocellular carcinoma (HCC). To clarify the<br />

etiological difference in body composition of HCC patients,<br />

we compared visceral adipose tissue index (VATI), subcutaneous<br />

adipose tissue index (SATI), and visceral to subcutaneous<br />

adipose tissue ratio (VSR) among etiological groups in the<br />

cohort. Methods: We enrolled naïve HCC patients diagnosed<br />

from January 2004 to December 2009 in our department.<br />

Patients were divided into 3 groups according to chronic viral<br />

infection: hepatitis C virus (HCV), hepatitis B virus (HBV) and<br />

non-viral etiology (NBNC). Those who were infected with both<br />

hepatitis viruses were excluded. To minimize the tumor effect<br />

on body composition, those with tumor larger than 5 cm were<br />

also excluded. We assumed that visceral fat accumulation<br />

increases inversely proportion to impact of virus related-risk of<br />

HCC development (i.e., HCV>HBV>NBNC). Each component<br />

of body composition was compared between two groups using<br />

HCV group as a control. Results: Baseline characteristics of<br />

enrolled patients were as follows: HCV group (N = 534, M/F<br />

= 311/223, median age = 71), HBV group (N = 84, M/F<br />

= 67/17, age = 61), and NBNC group (N = 126, M/F =<br />

97/29, age = 71). Median (interquartile range [IQR]) tumor<br />

size was 2.3 (1.8-2.9) cm. Number of tumor nodule was 1 in<br />

404, 2-3 in 248, and more than 3 in 92 patients. Child-Pugh<br />

class was A in 574, B in 164, C in 6. Significant difference<br />

was observed in VATI between HBV and HCV groups in males<br />

(P = 0.02) as well as between NBNC and HCV groups in<br />

males (P 0.05).<br />

11 patients were treated with CE + sorafenib. Advanced BCLC<br />

stage disease, cirrhosis, and increasing MELD were statistically<br />

significant predictors of decreased survival in univariate models.<br />

In multivariate models the only statistically significant predictors<br />

of improved survival included hepatitis B etiology and<br />

treatment with resection or ablation (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 443A<br />

463<br />

Progranulin Autoantibodies and Genetic Variation in<br />

Patients with Cholangiocarcinoma<br />

Vincent Zimmer 1 , Lorenz Thurner 2 , Monica Acalovschi 3 , Michael<br />

Pfreundschuh 2 , Frank Lammert 1 ; 1 Department for Medicine II, Saarland<br />

University Medical Center, Homburg, Germany; 2 Department<br />

of Medicine I, Saarland University Medical Center, Homburg, Germany;<br />

3 Department of Medicine III, University Iuliu Hatieganu,<br />

Cluj-Napoca, Romania<br />

Background: Recently, the novel growth factor and TNF signalling<br />

adaptor progranulin (PGRN) has been implicated in<br />

the inflammation-driven pathogenesis of cholangiocarcinoma<br />

(CCA).(1,2) Neutralizing PGRN autoantibodies have been<br />

described in autoimmune diseases, including inflammatory<br />

bowel disease (IBD).(3,4) Therefore, our aim now was to<br />

assess PGRN autoantibody incidence and potential effects of<br />

the common PGRN gene variant rs5848 (known to be associated<br />

with reduced PGRN serum concentrations) on susceptibility<br />

and PGRN autoantibody formation in a large European<br />

CCA population.(5) Patients & Methods: Overall, 221 individuals<br />

with CCA (males n = 131, age 66 ± 11 years) originating<br />

from Germany (n = 164) and Romania (n = 57) and 295<br />

CCA-free control individuals were genotyped. The common<br />

non-synonymous single nucleotide polymorphism (SNP) rs5846<br />

was genotyped by PCR-based assays with 5’-nuclease and<br />

fluorescence detection (TaqMan). PGRN autoantibodies were<br />

determined by ELISA.(4) Results: rs5848 genotype frequencies<br />

in the control group were consistent with the Hardy-Weinberg<br />

equilibrium (HWE), with a call rate > 99 % indicating robust<br />

genotyping. The association tests did not provide evidence<br />

for genetic CCA risk modulation by the PGRN variant (Odds<br />

ratio = 0.92; 95 % confidence interval = 0.71-1.19; P>0.05).<br />

Seven of 61 patients (11.5%) tested positive for PGRN autoantibodies,<br />

with no significant sex differences (3/32 males, 4/29<br />

females). Considering rs5848 variation in the subgroup with<br />

available PGRN autoantibody status, no significant differences<br />

in allele and genotype distribution were observed. Conclusions:<br />

Our preliminary data indicate PGRN autoantibodies as a novel<br />

potential tumor antigen in a defined subset of CCA patients.<br />

Extension of the current study in larger sample sizes and delineation<br />

of potential implications in terms of CCA subtypes and<br />

prognostic relevance of PGRN autoimmunity is warranted. References:<br />

1. Frampton et al. Gut 2012 2. Frampton et al. Am<br />

J Physiol Gastrointest Liver Physiol 2012 3. Thurner et al. J<br />

Autoimmun 2013 4. Thurner et al. Dig Dis Sci 2014 5. Hsiung<br />

et al. J Neurol Sci 2011<br />

Disclosures:<br />

The following authors have nothing to disclose: Vincent Zimmer, Lorenz Thurner,<br />

Monica Acalovschi, Michael Pfreundschuh, Frank Lammert<br />

464<br />

The effect of additional transarterial chemoembolization<br />

after combined chemoradiation therapy for locally<br />

advanced hepatocellular carcinoma with portal vein<br />

thrombosis<br />

Ja Kyung Kim, Jung Il Lee, Jun Won Kim, Ik Jae Lee, Seung-Moon<br />

Joo, Kwang-Hun Lee, Eun-Suk Cho, Tae Joo Jeon, Jae Keun Kim,<br />

Nomi Park, Ah-Young Kang, Kwan Sik Lee; Yonsei University College<br />

of Medicine, Seoul, Korea (the Republic of)<br />

Backgrounds: Combined chemoradiation therapy (CCRT) followed<br />

by intraarterial chemotherapy (IAC) has been reported<br />

to be beneficial for locally advanced hepatocellular carcinoma<br />

(HCC) with portal vein thrombosis (PVT). However, hepatic<br />

arterial chemoembolization (TACE) was generally not accepted<br />

for this situation. The aim of this study is to evaluate the role of<br />

adding TACE instead of IAC after CCRT. Methods: Thirty-seven<br />

patients were treated by CCRT as the initial treatment for HCC<br />

with PVT between 2009 and 2014. IAC through hepatic arterial<br />

chemoport and intermittent TACE (Group A) followed in<br />

18 patients and only IAC were repeated in 19 patients (Group<br />

B). We excluded the patients who had extra-hepatic metastasis,<br />

active peptic ulcer, and inadequate hepato-renal function.<br />

For CCRT, infusion of 5-FU (500 mg/day) via chemoport<br />

was given during the first and last five days of external radiation<br />

therapy (RTx) for 5 weeks. For IAC, infusion of cisplatin<br />

(70~100mg/m2 for one day) with 5-FU (750~1000mg for<br />

3 days) was repeated monthly. Results: Mean age was 55<br />

year; male to female was 32 to 5. Underlying liver diseases<br />

were hepatitis B, hepatitis C and non-B/C in 26, 2 and 10<br />

patients, respectively. Mean follow-up duration was 14 months<br />

(3-65 months). The objective response (CR+PR) rates in tumor<br />

size after 1 month after CCRT were 45.9%. The mean level<br />

of αFP after 1 month after CCRT were decreased in 81% of<br />

patients. During follow up, complete response was noted in<br />

3 patients. One patient underwent resection for cure. Median<br />

overall survival duration was 12.4 months. Overall survival of<br />

group A was significantly better than that of group B (mean<br />

42.5 vs. 12.2 months, p=0.001). Conclusion: Adding intermittent<br />

transarterial chemoembolization contributed survival gain<br />

following combined chemoradiation therapy for advanced<br />

hepatocellular carcinoma with portal vein thrombosis. Further<br />

prospective randomized study comparing TACE with TAC after<br />

CCRT is warranted.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ja Kyung Kim, Jung Il Lee, Jun<br />

Won Kim, Ik Jae Lee, Seung-Moon Joo, Kwang-Hun Lee, Eun-Suk Cho, Tae Joo<br />

Jeon, Jae Keun Kim, Nomi Park, Ah-Young Kang, Kwan Sik Lee<br />

465<br />

Clinicopathologic analysis of patients with HBV-positive,<br />

HCV-positive and non-B non-C hepatocellular carcinomas<br />

Jun Akiba 1 , Osamu Nakashima 2 , Yoshiki Naito 1 , Hironori<br />

Kusano 1 , Reiichiro Kondo 1 , Hirohisa Yano 1 ; 1 Department of<br />

Pathology, Kurume University School of Medicine, Kurume, Japan;<br />

2 Department of Clinical Laboratory Medicine, Kurume University<br />

Hospital, Kurume, Japan<br />

[aim] The chronic infection of hepatitis B virus (HBV) and hepatitis<br />

C virus (HCV) is strongly associated with HCC. However,<br />

recent developments of anti-viral treatments for HBV and HCV<br />

contribute to reduce virus-associated HCC. On the other hand,<br />

HCCs, which are negative for both hepatitis B surface antigen<br />

(HBs-Ag) and anti-hepatitis C antibody (HCV-Ab), are increasing<br />

in number in developed nations. The aim of this study is<br />

to clarify the clinicopathologic features of surgically resected<br />

HCCs with different etiologies. [Materials and methods] Surgically<br />

resected 1430 cases (1598 nodules) at our hospital from<br />

1991 to 2013 were enrolled in this study. All cases did not<br />

have any previous treatments for HCC. HBV-related HCC (HBV-<br />

HCC) and HCV-related HCC (HCV-HCC) were defined as positive<br />

for serum HBs-Ag and serum HCV-Ab, respectively. Non<br />

HBV and non HCV-related HCC (NBNC-HCC) was defined<br />

as negative for both HBs-Ag and HCV-Ab. The cases which<br />

were positive for both HBs-Ag and HCV-Ab were excluded.<br />

Clinicopathologic features were compared. [Results] Out of<br />

1430, 253, 949 and 228 cases were HBV-HCC, HCV-HCC<br />

and NBNC-HCC, respectively. Patient age of HBV-HCC (55.8)<br />

was significantly younger than those of HCV-HCC (67.6) and<br />

NBNC-HCC (68.5). Tumor size was the smallest in HCV-HCC


444A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(31.5 mm), followed by HBV-HCC (40.0 mm) and NBNC-HCC<br />

(46.4 mm). Serum AFP and DCP were significantly higher in<br />

HBV-HCC than in HCV-HCC. Macroscopically, the percentage<br />

of HCC with indistinct margin was significantly higher<br />

in HCV-HCC (10.7%) compared with HBV-HCC (3.2%) and<br />

NBNC-HCC (3.6%). On the other hand, HBV-HCC (20.7%) significantly<br />

more frequently showed confluent multinodular type<br />

compared with HCV-HCC (11.2%) and NBNC-HCC (8.7%).<br />

The number of well-differentiated HCC and poorly differentiated<br />

HCC were significantly higher in HCV-HCC (11.1%) and<br />

in HBV-HCC (22.1%), respectively. Portal vein invasion was<br />

significantly more frequent in HBV-HCC (67.4%) compared<br />

with HCV-HCC (46.1%) and NBNC-HCC (56.5%). Patients<br />

with NBNC-HCC significantly had higher rate of a history of<br />

diabetes and alcohol intake. Out of 228 NBNC-HCC patients,<br />

21.3% showed steatohepatitis in the background liver. About<br />

half of them (11.2%) had a history of alcohol intake. The<br />

residual cases (10.1%) were defined as non-alcoholic steatohepatitis.<br />

Also, 19.7% of NBNC-HCC showed positivity of at<br />

least one of serum HBV-related markers except HBs-Ag. [conclusions]<br />

HBV-HCC, HCV-HCC and NBNC-HCC had different<br />

clinicopathologic features. HBV-HCC showed aggressive biologic<br />

features. NBNC-HCC was closely associated with not<br />

only diabetes and alcohol intake but also HBV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jun Akiba, Osamu Nakashima,<br />

Yoshiki Naito, Hironori Kusano, Reiichiro Kondo, Hirohisa Yano<br />

466<br />

The presence of histological abnormalities even after the<br />

elimination of hepatitis C virus<br />

Nobuhiro Aizawa 1 , Kyohei Kishino 1 , Yoshihiro Shimono 1 , Chikage<br />

Nakano 1 , Kunihiro Hasegawa 1 , Ryo Takata 1 , Kazunori<br />

You 1 , Akio Ishii 1 , Tomoyuki Takashima 1 , Yoshiyuki Sakai 1 , Naoto<br />

Ikeda 1 , Takashi Nishimura 1 , Hiroki Nishikawa 1 , Yoshinori Iwata 1 ,<br />

Hirayuki Enomoto 1 , Hiroko Iijima 1 , Seikan Hai 2 , Jiro Fujimoto 2 ,<br />

Shuhei Nishiguchi 1 ; 1 Division of Hepatobiliary and Pancreatic Diseases,<br />

Department of Internal Medicine, Hyogo collage of medicine,<br />

Nishinomiya, Japan; 2 Division of surgery, Hyogo college of<br />

medicine, Nishinomiya, Japan<br />

Background: Despite the improved histological findings in the<br />

liver after hepatitis C virus (HCV) eradication, hepatocellular<br />

carcinoma (HCC) occurs in some patients with a sustained<br />

virological response (SVR) to interferon therapy. However, the<br />

mechanism of carcinogenesis in virus-eliminated liver tissues<br />

remains unclear. The aim of this study is to clarify the characteristics<br />

of SVR-related hepatocarcinogenesis. Subjects and<br />

Methods: A total of 43 patients with SVR were enrolled. Fifteen<br />

patients had HCC (Group A), and the remaining 28 patients<br />

did not (Group B). We pathologically examined the non-cancerous<br />

hepatic tissues of Group A patients, and compared those<br />

tissues with the liver tissues of Group B patients by electron<br />

microscopy (EM). We evaluated the degree of the dilatation of<br />

vesicular endoplasmic reticulum on a scale of 0 to 3. We also<br />

evaluated the mitochondrial alterations on a scale of 0 to 10<br />

according to the following points: (1) Dense granules, (2) Paracrystalline<br />

inclusions, (3) Vacuole formation, (4) Irregularity<br />

of mitochondrial shapes, and (5) Lack of mitochondrial membrane.<br />

This study was conducted in accordance with the Helsinki<br />

declaration and written informed consents were obtained<br />

from all patients before the study. Results: Regarding dilatation<br />

of vesicular endoplasmic reticulum, elevated scores (> 2 points)<br />

were found in 14.3% (4/28) of the Group B patients and<br />

in 92.9% (13/15) of the Group A patients. Regarding the<br />

morphological mitochondrial alterations, elevated scores (><br />

4 points) were found in 3.6% (1/28) of the Group B patients<br />

and in 80.0% (12/15) of the Group A patients. Scores of<br />

the alterations regarding the mitchondria and the vesicular<br />

endoplasmic reticulum were higher in the HCC Group than in<br />

the non-HCC Group. The morphological abnormalities were<br />

not associated with the histological findings, including activity<br />

grade and fibrosis stage. Conclusion: The EM findings<br />

revealed the presence of histological abnormalities even after<br />

the elimination of HCV. In particular, remarkable morphological<br />

abnormalities were observed in non-cancerous hepatic<br />

tissues of HCC patients. Their histological alterations after SVR<br />

should be correlated with hepatocarcinogenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nobuhiro Aizawa, Kyohei<br />

Kishino, Yoshihiro Shimono, Chikage Nakano, Kunihiro Hasegawa, Ryo Takata,<br />

Kazunori You, Akio Ishii, Tomoyuki Takashima, Yoshiyuki Sakai, Naoto Ikeda,<br />

Takashi Nishimura, Hiroki Nishikawa, Yoshinori Iwata, Hirayuki Enomoto,<br />

Hiroko Iijima, Seikan Hai, Jiro Fujimoto, Shuhei Nishiguchi<br />

467<br />

Prognostic Score predicting overall survival of Patients<br />

with Intermediate Stage of Hepatocellular Carcinoma<br />

after Transarterial Chemoembolization<br />

Saharat Jarupongprapa, Supot Nimanong, Siwaporn Chainuvati,<br />

Phunchai Charatcharoenwitthaya, Tawesak Tanwandee, Watcharasak<br />

Chotiyaputta; Gastroenterology, Internal Medicine, Siriraj<br />

Hospital, Mahidol University, Bangkok, Thailand<br />

Background: Transarterial chemoembolization (TACE) is the<br />

standard of care for patients (pts) with intermediate stage<br />

of hepatocellular carcinoma (HCC). Multiple TACE sessions<br />

are common in treatment of HCC which can result in shorten<br />

survival. Several baseline characteristics and dynamic tumor<br />

response after the first TACE were studied to select the patients<br />

who are not beneficial form repeated TACE. Objectives: To<br />

determine factors associated with overall survival (OS) after first<br />

TACE and to develop a prognostic score to identify candidates<br />

who have benefit form repeated TACE. Methods: A retrospective<br />

study between 2007 and 2012, all HCC pts underwent<br />

at least two consecutive TACE within 90 days. Pts with Child-<br />

Pugh (CTP) C, Barcelona Clinic Liver Cancer (BCLC) stage C<br />

and ruptured HCC at presentation were excluded. Baseline<br />

characteristics, liver function and dynamic tumor response after<br />

first TACE were analyzed on the median OS. Results: 216<br />

pts were included in this study. 75% were male, mean age<br />

was 59.2 years, and 74.5% were viral associated HCC. At<br />

baseline, mean tumor size was 7.9 cm, 62.5% was unilobar,<br />

and 60.6% was multifocal HCC. The median OS was 17.6<br />

months (15.3-19.9). 89.8% of pts (n=194) died during study<br />

period. On multivariate analysis, five parameters were significant<br />

factors to predict OS including baseline CTP B [HR1.77,<br />

p=0.04], BCLC stage B [HR1.95, p=0.001], increased CTP<br />

score after TACE [+1 (HR2.29), +2 (HR11.74), p 200 IU/L with < 50% reduction after TACE<br />

[HR2.07, p=0.003], and radiologic response (stable or progressive<br />

disease on mRECIST) [HR1.56, p=0.01]. A prognostic<br />

score using five parameters was developed and divided into<br />

two groups (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 445A<br />

and high major complications. Retreatment with TACE should<br />

be avoided.<br />

Figure 1. Overall survival analysis according to SMART<br />

Disclosures:<br />

Tawesak Tanwandee - Grant/Research Support: MSD, BMS, Fibrogen, Biotron,<br />

Celcion<br />

The following authors have nothing to disclose: Saharat Jarupongprapa, Supot<br />

Nimanong, Siwaporn Chainuvati, Phunchai Charatcharoenwitthaya, Watcharasak<br />

Chotiyaputta<br />

468<br />

Combination of modified Response Evaluation Criteria<br />

in Solid Tumors and hand-foot-skin reaction: a new<br />

prognostic evaluation for HCC patients treated with<br />

sorafenib and transarterial chemoembolization<br />

Wenjun Wang 1 , Yan Zhao 1 , Wei Bai 1 , Lei Liu 1 , Man Yang 1 , Hongwei<br />

Cai 3 , Lei Zhang 1 , Zhanxin Yin 1 , Zhuoli Zhang 1 , Daiming<br />

Fan 1 , Jielai Xia 2 , Guohong Han 1 ; 1 Xijing Hospital of Digestive<br />

Diseases, Xijing Hospital, Fourth Military Medical University, Xi’an,<br />

China; 2 Department of Medical Statistics, Fourth Military Medical<br />

University, Xi’an, China; 3 Information Center, School of Stomatology,<br />

Fourth Military Medical University, Xi’an, China<br />

The aim of this study was to combine the modified Response<br />

Evaluation Criteria in Solid Tumors (mRECIST) and hand-footskin<br />

reaction (HFSR) to establish a new prognostic evaluation<br />

for hepatocellular carcinoma (HCC) patients undergoing<br />

sorafenib and transarterial chemoembolization (TACE). 176<br />

consecutive intermediate-advanced HCC patients treated with<br />

combination therapy were enrolled. The therapeutic responses<br />

were assessed by mRECIST and HFSR criteria at 1, 2 and 3<br />

months, respectively. Uni/multivariate Cox regressions were<br />

used to investigate the earliest time when treatment responses<br />

could be accurately assessed. Then according to the mRECIST<br />

and HFSR assessed at that time, SMART (Sorafenib with Modified<br />

RECIST Assessment plus hand-foot-skin Reaction in TACE)<br />

prognostic evaluation was developed: SMART A, responders<br />

on both assessments; SMART B, responders on either of assessment<br />

and SMART C, non-responders on both assessments.<br />

Moreover, we explore the prognostic value of SMART for<br />

predicting overall survival(OS) in comparison with mRECIST<br />

and HFSR about likelihood ratio, Akaike information criterion<br />

(AIC) and C-index respectively. The earliest time at which the<br />

responses of mRECIST and HFSR correlated with the survival<br />

was 2 months after therapy. The SMART stratified patients with<br />

three different prognosis; the SMART A had the longest median<br />

OS, followed by SMART B and SMART C (30.5, 17.4, and<br />

8.3 months, respectively; P


446A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

6% and 1% patients, respectively. In explant analysis, tumor<br />

was uninodular in 45% and moderately differentiated in the<br />

majority of cases (66%). Median HCC size was 28mm. Vascular<br />

invasion and satellite nodules were observed in 24.5%<br />

and 25% of patients, respectively. In 818 patients that survived<br />

beyond the immediate post-transplant period, mean follow-up<br />

was 27,7 months (±23,8), an overall survival was 70% in 5<br />

years. Recurrence occurred in 8/818 (8%) cases, at a mean<br />

time of 15 months (1.5-76m). Sites of recurrence were 40% in<br />

liver, extrahepatic in 46% and both hepatic and extrahepatic<br />

in 14%. Vascular invasion and alpha-fetoprotein (AFP) level<br />

before liver transplantation were risk factors for tumor recurrence.<br />

The presence of HCC recurrence was directly related<br />

to poor survival. Female gender, need for re-transplantation,<br />

vascular invasion and explant outside the Milan criteria were<br />

also predictors of poor survival. Conclusion: Liver transplantation<br />

for hepatocellular carcinoma in Brazil was associated with<br />

an overall survival of 70% in 5 years. HCC recurrence ocurred<br />

in 8% of patients. The presence of vascular invasion and AFP<br />

before liver transplantation were associated with increased<br />

risk of tumor recurrence. Female, vascular invasion, explant<br />

outside the Milan criteria and HCC recurrence were related to<br />

poor survival.<br />

Disclosures:<br />

The following authors have nothing to disclose: Aline Chagas, Luciana Kikuchi,<br />

Guilherme Felga, Angelo A. Mattos, Renato F. Silva, Rita de Cássia M. da Silva,<br />

Fernanda Branco, Marcio D. Almeida, Ilka F. Boin, José H. Garcia, Luiz C. D’Albuquerque,<br />

Flair J. Carrilho<br />

470<br />

MESIAH score is an effective subclassification tool in<br />

Barcelona Clinic Liver Cancer stage B: comparison with<br />

other proposed methods<br />

Jeong-Ju Yoo 1 , Su Jong Yu 1 , Eun Ju Cho 1 , Jeong-Hoon Lee 1 , June<br />

Sung Lee 2 , Yoon Jun Kim 1 , Jung-Hwan Yoon 1 ; 1 Department of<br />

Internal Medicine and Liver Research Institute, Seoul National<br />

University College of Medicine, Seoul, Korea (the Republic of);<br />

2 Department of Internal Medicine, Ilsan Paik Hospital, Inje University<br />

College of Medicine, Goyang, Korea (the Republic of)<br />

Background: The intermediate stage of hepatocellular carcinoma<br />

(HCC) is a highly heterogeneous population, therefore<br />

many models to predict survival of patients has been proposed.<br />

The aim of this study is to evaluate the prognostic performance<br />

in intermediate HCC among four models; Model to Estimate<br />

Survival in Ambulatory Patient (MESIAH), original Barcelona<br />

Clinic Liver Cancer (BCLC) B subclassification, modified model<br />

A and modified model B. Methods: From January 2005 to<br />

December 2006, a total of 1184 HCC patients who initially<br />

treated with TACE were enrolled. Among them, 235 (19.8%)<br />

patients were classified as BCLC stage B. Four subclassification<br />

systems were tested; MESIAH, original BCLC B subclassification<br />

(B1, B2, B3, B4), modified model A (B1, B2, B3+B4), and<br />

modified model B (B1, B2+B3, B4). The predictive accuracies<br />

of the four systems were compared based on c-statistic (c-index)<br />

together with its confidence interval (CI). Results: Patients in<br />

the cohort had a median age of 57 years and 80.9% were<br />

men. 81.7% had a Child–Pugh class A and Hepatitis B virus<br />

infection was present in 71.1% and the median overall survival<br />

was 57.8 months. The MESIAH score had a highest degree<br />

of discrimination, with a C-statistic of 0.617 [95% confidence<br />

interval (CI), 0.564–0.670, hazard ratio 2.06 (1.72-2.48),<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 447A<br />

*all from bilirubin increase<br />

**AFP reduction from baseline of ≥20% AFP or ≥50% AFP in<br />

patients with baseline value ≥20ng/mL<br />

Disclosures:<br />

Robin K. Kelley - Advisory Committees or Review Panels: Acceleron Inc.; Consulting:<br />

Eli Lilly & Co., Acceleron Inc.; Grant/Research Support: Exelixis, Inc.,<br />

Celgene Corp., Eli Lilly & Co., Regeneron Pharmaceuticals, Tekmira Pharmaceuticals,<br />

Novartis, Sanofi<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

The following authors have nothing to disclose: Varun Saxena, Neil Mehta, K.<br />

Pallav Kolli, Robert Kerlan, Francis Y. Yao, Albert Chang<br />

two groups (P = 0.666). Multivariate analysis identified ICG<br />

R and solitary tumor as independent factors associated with<br />

disease-free survival. Conclusion: The two treatment modalities<br />

attained similar survival benefits in the management of initial<br />

recurrent HCC after curative treatment. In addition to host and<br />

tumor factors, time from primary HCC development to recurrence<br />

was a prognostic factor for recurrence of HCC.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Asuka, Bayer, Daiichi Sankyo, Dainippon Sumitomo, Eisai, Janssen,<br />

Kowa, Mitsubishi Tanabe, MSD, Eli Lily, Nippon Kayaku, Nippon Shinyaku,<br />

Otsuka, Roche, Takeda, Toray, Torii, Tsumura, Zeria; Speaking and Teaching:<br />

Eisai, Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Asuka, Bayer, BMS,<br />

Chugai, Daiichi Sankyo, Dainippon Sumitomo, J&J, Jimro, Miyarisan, MSD,<br />

Nihon Kayaku, Olympus<br />

The following authors have nothing to disclose: Takayuki Fukuhara, Hiroshi<br />

Aikata, Fumi Shinohara, Norihito Nakano, Yuki Nakamura, Masahiro Hatooka,<br />

Kei Morio, Tomoki Kobayashi, Yuuko Nagaoki, Tomokazu Kawaoka, Masataka<br />

Tsuge, Akira Hiramatsu, Michio Imamura, Yoshiiku Kawakami<br />

472<br />

Survival analysis of second hepatectomy versus radiofrequency<br />

ablation for initial recurrent small hepatocellular<br />

carcinoma after curative treatment<br />

Takayuki Fukuhara, Hiroshi Aikata, Fumi Shinohara, Norihito<br />

Nakano, Yuki Nakamura, Masahiro Hatooka, Kei Morio, Tomoki<br />

Kobayashi, Yuuko Nagaoki, Tomokazu Kawaoka, Masataka<br />

Tsuge, Akira Hiramatsu, Michio Imamura, Yoshiiku Kawakami,<br />

Kazuaki Chayama; Hiroshima university, Hiroshima, Japan<br />

Background and Aim: Japanese and American clinical practice<br />

guidelines recommend hepatectomy and radiofrequency<br />

ablation (RFA) as curative treatments for small hepatocellular<br />

carcinoma (HCC; ≤ 3 nodules and each ≤ 3 cm in diameter).<br />

However, the algorithms recommended in the above-mentioned<br />

guidelines are for primary HCC, not recurrent HCC. The treatment<br />

algorithm for recurrent HCC is not yet reach consensus.<br />

The aim of this retrospective study is to compare the efficacy<br />

of hepatectomy and RFA as the treatment of initial recurrence<br />

after curative treatment. Methods: The following inclusion criteria<br />

were used: 1) initial recurrent HCC after curative treatment,<br />

2) ≤ 3 nodules and each ≤ 3 cm in diameter, 3) no vascular<br />

invasion and no extrahepatic metastasis, and 4) Child-Pugh<br />

class A or B. From January 2001 to December 2014, one<br />

hundred and ninety two patients who met the inclusion criteria<br />

were evaluated. Patients were divided into 2 groups: second<br />

hepatectomy group (n = 120) and RFA group (n = 72). To overcome<br />

bias due to the different distribution of covariates for the<br />

two groups, a one-to-one match was created using propensity<br />

score analysis. After matching, we compared overall survival<br />

(OS), disease-free survival (DFS) between these groups. Results:<br />

1) After one-to-one matching, there was no significant difference<br />

in clinical backgrounds between second hepatectomy<br />

group (n = 62) and RFA group (n = 62). 2) The rate of OS at<br />

3, 5 and 10 years was 77.1%, 59.0% and 38.9% in the second<br />

hepatectomy group and 85.3%, 58.6% and 35.1% in the<br />

RFA group, respectively. There was no significant difference in<br />

OS between these two groups (P = 0.986). Multivariate analysis<br />

identified several host and tumor factors (age


448A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Armando Carvalho - Advisory Committees or Review Panels: Gilead Sciences,<br />

Bristol-Myers Squibb, Roche, Janssen, Bayer, AbbVie; Grant/Research Support:<br />

Boehringer-Ingelheim, AbbVie<br />

The following authors have nothing to disclose: Adélia Simão, Miriam Santos,<br />

Raquel S. Silva, Pedro Correia, Lurdes Correia, Pedro Henriques Abreu, Alberto<br />

Cardoso, José N. Costa<br />

474<br />

Long-term Survival Analysis of Re-resection Versus<br />

Re-radiofrequency Ablation for Intrahepatic Recurrence<br />

After Initial Hepatectomy or Radiofrequency Ablation<br />

for Hepatocellular Carcinoma: A 10-year Observational<br />

Study<br />

Minjong Lee 1 , Sohee Oh 2 , YOUNG CHANG 1 , Joon Yeul Nam 1 ,<br />

Young Youn Cho 1 , Jeong-Ju Yoo 1 , Yuri Cho 1 , Donghyeon Lee 1 ,<br />

Eun Ju Cho 1 , Jeong-Hoon Lee 1 , Su Jong Yu 1 , Nam-Joon Yi 3 , Jeong<br />

Min Lee 4 , Kwang-Woong Lee 3 , Jung-Hwan Yoon 1 , Kyung-Suk<br />

Suh 3 , Yoon Jun Kim 1 ; 1 Department of Internal Medicine and Liver<br />

Research Institute, Seoul National University College of Medicine,<br />

Seoul, Korea (the Republic of); 2 Department of Biostatistics, Seoul<br />

Metropolitan Government Seoul National University Boramae<br />

Medical Center, Seoul, Korea (the Republic of); 3 Department of<br />

Surgery, Seoul National University College of Medicine, Seoul,<br />

Korea (the Republic of); 4 Department of Radiology, Seoul National<br />

University College of Medicine, Seoul, Korea (the Republic of)<br />

Background/Aim: Survival outcomes of resection (OP) and<br />

radiofrequency ablation (RFA) were not significantly different<br />

in HCC patients with the Barcelona Clinic Liver Cancer (BCLC)<br />

stage 0/A. However, in patients with recurrent HCC, survival<br />

rates of patients treated with re-OP have not been compared<br />

with those with re-RFA. We aimed to compare survival of patient<br />

treated with re-OP to those with re-RFA in patients with BCLC<br />

0/A. Methods: This retrospective study included 225 patients<br />

with recurred BCLC 0/A HCC in Child-Pugh class A/B. According<br />

to initial treatment strategies, patients were divided into two<br />

groups; 148 patients treated with RFA; 77 patients with OP.<br />

For treating recurrent HCC, patients were also divided into two<br />

treatments: 181 patients treated with RFA; 44 patients with OP.<br />

Survival rates comparing each group according to treatment<br />

strategies were analyzed after matching of propensity scores<br />

after matching of propensity scores according to established<br />

liver cirrhosis or HCC risk factors. Results: The median tumor<br />

size 2.4/1.5 cm; single HCC 80.9/84.9% in patients with<br />

initial/recurrent HCC, respectively. The 5-year overall survival<br />

rates were 85.3%, 89.5%, 93.0%, and 95.0% in patients<br />

treated with RFA-RFA, OP-RFA, RFA-OP, and OP-OP for initial<br />

and recurrent HCC, respectively. Patients treated with re-OP<br />

for initial and recurrent HCCs showed significantly lower mortality<br />

than those with re-RFA; hazard ratio (HR)=0.143 (95%<br />

confidence interval[CI], 0.021–0.991, P=0.04). Patients who<br />

received resection at least for initial or recurrent HCC showed<br />

lower mortality than those who only treated with re-RFA;<br />

HR=0.414 (95% CI, 0.187–0.913, P=0.029). Conclusions: In<br />

patients with initial and recurrent BCLC stage 0/A, and early<br />

cirrhosis, who recurred HCC with BCLC stage 0/A, re-hepatectomy<br />

improves the patient survival compared to those treated<br />

with re-RFA treatment. Treatments including resection for initial<br />

or recurrent HCC also improves the survival when compared<br />

to RFA only treatment. Therefore, for selected patients, re-hepatectomy<br />

should be considered as the first treatment option for<br />

initial and recurrent HCC.<br />

Disclosures:<br />

Jeong Min Lee - Advisory Committees or Review Panels: Bayer healthcare<br />

Kwang-Woong Lee - Grant/Research Support: ChongGeunDang, Astellas,<br />

GreenCross<br />

Yoon Jun Kim - Grant/Research Support: Bristol-Myers Squibb, Roche, JW Creagene,<br />

Bukwang Pharmaceuticals, Handok Pharmaceuticals, Hanmi Pharmaceuticals,<br />

Yuhan Pharmaceuticals; Speaking and Teaching: Bayer HealthCare<br />

Pharmaceuticals, Gilead Science, MSD Korea, Yuhan Pharmaceuticals, Samil<br />

Pharmaceuticals, CJ Pharmaceuticals, Bukwang Pharmaceuticals, Handok Pharmaceuticals<br />

The following authors have nothing to disclose: Minjong Lee, Sohee Oh, YOUNG<br />

CHANG, Joon Yeul Nam, Young Youn Cho, Jeong-Ju Yoo, Yuri Cho, Donghyeon<br />

Lee, Eun Ju Cho, Jeong-Hoon Lee, Su Jong Yu, Nam-Joon Yi, Jung-Hwan Yoon,<br />

Kyung-Suk Suh<br />

475<br />

Simple serological tests and Transient Elastography as<br />

screening tools for clinically significant portal hypertension<br />

in resectable hepatocellular carcinoma<br />

Simona Bota, Florian Hucke, Mattias Mandorfer, Remy Schwarzer,<br />

Philipp Schwabl, Thomas Reiberger, Arnulf Ferlitsch, Michael<br />

Trauner, Wolfgang Sieghart, Markus Peck-Radosavljevic; Gastroenterology<br />

and Hepatology, Medical University of Vienna, Vienna,<br />

Austria<br />

Background:Hepatocellular carcinoma (HCC) resection is recommended<br />

by the EASL Guideline as first-line treatment option<br />

for patients with solitary HCC and very well-preserved liver<br />

function, defined as normal bilirubin with either hepatic venous<br />

pressure gradient (HVPG)≤10 mmHg.However HVPG measurements<br />

are available only in highly specialized Hepatology centers.<br />

Aim:to assess the diagnostic value of serological scores<br />

(initially developed for fibrosis evaluation) and transient elastrography<br />

(TE) as screening tool for prediction of HVPG>10<br />

mmHg in patients with potentially resectable HCC. Methods:<br />

Our study included 46 consecutive patients with chronic liver<br />

disease and potentially resectable (Child-Pugh A,normal bilirubin,<br />

platelet count≥100000cells/mm 3 and no esophageal<br />

varices) evaluated in our Hemodynamic Labor.We performed<br />

HVPG measurements, serological tests and TE in a subgroup<br />

of patients.Simple serological scores were calculated:Fibrosis<br />

Index (FI),Forns score and Lok score. According to the newly<br />

proposed TE quality criteria, ten valid measurements with a<br />

median liver stiffness value ≥7.1kPa and IQR/median>30%<br />

were considered poorly reliable and not included for analysis.<br />

TE cases without 10 valid measurements were also excluded.<br />

Results:The etiology of liver disease was : alcholic -41.3%,<br />

viral -39.1%, others etiologies – 19.6%. The presence of<br />

HVPG>10mmHg was diagnosed in 10/46 patients (21.7%).<br />

Liver stiffness measurements by TE were available in 67.3%of<br />

patients. Reliable liver stiffness measurements were obtained<br />

in 87% of patients. Transient Elastography had a very good<br />

performace (93.7%) to exclude the presence of CSPH, while<br />

from the serological tests Fibrosis Index (FI) had the best performance<br />

(NPV 84.7%)(Table) Conclusions:TE is a very good<br />

tool to exclude the presence of HVPG>10mmHg in patients<br />

with potentially resectable HCC.Simple serological tests were<br />

less accurate and could correctly exclude CSPH in 80-85% of<br />

cases, but were feasible in all patients.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 449A<br />

Disclosures:<br />

Simona Bota - Speaking and Teaching: Janssen Pharmaceutica, Bristol-Myers<br />

Squibb<br />

Mattias Mandorfer - Consulting: Janssen; Speaking and Teaching: AbbVie, Gilead,<br />

Janssen, Boehringer Ingelheim, Bristol-Myers Squibb, Roche<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Wolfgang Sieghart - Grant/Research Support: Bayer Schering Pharma, Bayer<br />

Schering Pharma, Bayer Schering Pharma, Bayer Schering Pharma; Speaking<br />

and Teaching: Bayer Schering Pharma, Bayer Schering Pharma, Bayer Schering<br />

Pharma, Bayer Schering Pharma<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

The following authors have nothing to disclose: Florian Hucke, Remy Schwarzer,<br />

Philipp Schwabl, Arnulf Ferlitsch<br />

476<br />

Pro-angiogenic Tie2-expressing monocytes/TEMs as a<br />

biomarker of the response to sorafenib in patients with<br />

advanced hepatocellular carcinoma<br />

Hirotaka Shoji 1,2 , Sachiyo Yoshio 1 , Yohei Mano 1 , Masaya Sugiyama<br />

1 , Yoshihiko Aoki 1 , Norio Itokawa 3 , Masanori Atsukawa 3 ,<br />

Yosuke Osawa 4 , Kiminori Kimura 4 , Akinobu Taketomi 2 , Masashi<br />

Mizokami 1 , Tatsuya Kanto 1 ; 1 The Research Center for Hepatitis &<br />

Immunology, National Center for Global Health and Medicine,<br />

Ichikawa, Japan; 2 Department of Gastroenterological Surgery<br />

1, Hokkaido University Graduate School of Medicine, Sapporo,<br />

Japan; 3 Division of Gastroenterology, Department of Internal Medicine,<br />

Nippon Medical School Chiba Hokusoh Hospital, Inzai,<br />

Japan; 4 Division of Hepatology, Tokyo Metropolitan Cancer and<br />

Infectious Diseases Center Komagome Hospital, Tokyo, Japan<br />

Background Sorafenib is a multi-kinase inhibitor that has been<br />

used for the patients with advanced stages of HCC. The survival<br />

benefit of sorafenib has been clinically proven, however,<br />

majority of patients are forced to discontinue the therapy due to<br />

severe adverse effects. It is anticipated to discover the biomarkers<br />

for the prediction of sorafenib response, in order to establish<br />

the tailored therapy by selecting patients promising for the<br />

outcome. We reported that Tie2-expressing monocytes (TEMs)<br />

were increased in patients with HCC, the frequency of which<br />

was correlated with the degree of angiogenesis in the liver.<br />

We sought to examine the possibility of TEMs as the biomarker<br />

for the drug response in patients with HCC. Furthermore, in<br />

order to discover more feasible markers reflecting the response<br />

to sorafenib, we explore the factors involving in the development<br />

of Tie2 expression on myeloid lineage cells. Method We<br />

enrolled 25 patients with advanced HCC treated with sorafenib<br />

in this study (13 Child-Pugh A and 12 CP-B). The anti-tumor<br />

effect of sorafenib was evaluated after a month and thereafter<br />

by dynamic CT and MRI with the criteria of modified RECIST.<br />

We analyzed the frequency of TEMs (CD14+CD16+TIE2+<br />

cells) in the periphery of the patients and its correlation with<br />

clinical and radiological response to sorafenib. We examined<br />

41 serum factors by multiplexed analysis in 71 HCC patients<br />

and healthy donors. Finally, we examined the impact of such<br />

factors on the differentiation of TEMs from monocytes in vitro.<br />

Results Among the subjects, 9 patients tolerated the continuation<br />

of sorafenib therapy. They were categorized into the<br />

responder group (one partial response and 4 stable disease)<br />

and non-responder group (4 progressive disease) judged by<br />

the modified RECIST. In the responders, the pre-treatment frequency<br />

of TEMs tended to be higher than those in the non-responders<br />

(6.1±2.2% vs. 3.5±0.9%, P=0.06). Furthermore, the<br />

frequency of TEMs in the responders was dropped from 6.1%<br />

to 3.5% after one month. In contrast, TEMs in the non-responders<br />

did not change after the therapy. In patients with high TEMs<br />

frequency (>2.75% cut-off), the levels of M-CSF, HGF, Osteopontin,<br />

and Follistatin were significantly higher than those in<br />

patients with lesser TEMs (p


450A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

m 2 for the men and 77.9 ± 44.3 cm 2 /m 2 for the women. The<br />

median survival times for the men and women were 12.4 and<br />

20.1 months, respectively. In the multivariate Cox regression<br />

analysis, only L3-SMI < 43 cm 2 /m 2 (hazard ratio [HR], 5.17;<br />

95% confidence interval [CI], 1.81–14.79; p = 0.002) and<br />

Child-Pugh score (HR, 1.95; 95% CI, 1.19–3.21; p = 0.008)<br />

were independently associated with mortality in the men only.<br />

Conclusions: Skeletal muscle depletion was a poor prognostic<br />

factor in the treatment with sorafenib for advanced HCC. Prevention<br />

of skeletal muscle depletion may be important for the<br />

improvement of the prognosis in advanced HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takashi Hoshino, Atsushi<br />

Naganuma, Yuhei Suzuki, Daisuke Uehara, Takahiro Miyoshi, Ken Sato, Satoru<br />

Kakizaki, Masanobu Yamada, Hitoshi Takagi<br />

478<br />

The role of cytokines and adhesion molecules in hepatocellular<br />

carcinoma to predict survival<br />

Filiz Akyuz 1 , Raim Iliaz 1 , Umit Akyuz 2 , Derya Duranyildiz 1 , Murat<br />

Serilmez 1 , Cetin Karaca 1 , Kadir Demir 1 , Fatih Besisik 1 , Sabahattin<br />

Kaymakoglu 1 ; 1 Istanbul University, Istanbul, Turkey; 2 Fatih Sultan<br />

Mehmet Educational and research center, Istanbul, Turkey<br />

Introduction: Hepatocellular carcinoma (HCC) is one of the<br />

most common cancers worldwide. The prognosis is related to<br />

stage of tumor and predictors of prognosis may determine the<br />

management of HCC. Some cytokines and growth factors are<br />

increased in HCC and these may be related with prognosis of<br />

the cancer. We aimed to investigate the effects of interleukin<br />

(IL)-32, IL-1 beta, IL- 18, vascular cellular adhesion molecule-1<br />

(VCAM-1) and epithelial cell adhesion molecule (EpCAM) on<br />

the prognosis of patients with HCC. Material and Methods: Fifty<br />

HCC patients and 19 healthy volunteers as a control group<br />

were enrolled in this prospective study. All demographic and<br />

clinical data were analyzed. Serum samples were obtained<br />

on first admition before any adjuvant and metastatic treatment<br />

was given. Serum IL-32, IL-1 beta, IL-18, VCAM-1 and EpCAM<br />

were determined by using ELISA kits. Results: All patients had<br />

cirrhosis and Child-Pugh stage were as follows: %70 Child<br />

A, 28% Child B and 2% Child C (78% HBV, 2% HCV, others<br />

cryptogenic). Fifty-eight percent of HCC patients were died<br />

in median 7.3 months. The mean age of the patients was<br />

60.2±9.4 years and 53.4±8.2 years for the controls. Mean<br />

serum level of IL-32 was higher in patients with HCC than controls<br />

(65.1 vs 13.7 pg/mL) (p0.05). Conclusion:<br />

IL-32, IL-18, VCAM-1 and EpCAM were significantly higher in<br />

HCC patients. Although, cytokines can be a diagnostic marker<br />

for HCC, they had no prognostic value in HCC patients. On<br />

the other hand EpCAM can be use to the determine prognosis<br />

of HCC and the management of treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Filiz Akyuz, Raim Iliaz, Umit<br />

Akyuz, Derya Duranyildiz, Murat Serilmez, Cetin Karaca, Kadir Demir, Fatih<br />

Besisik, Sabahattin Kaymakoglu<br />

479<br />

Reappraisal of portal vein tumor thrombosis as a prognostic<br />

factor for patients with hepatocellular carcinoma:<br />

tumor factor matters<br />

YOUNG CHANG 1 , Minjong Lee 1 , Joon Yeul Nam 1 , Hongkeun<br />

Ahn 1 , Hyeki Cho 1 , Yuri Cho 1 , Jeong-Ju Yoo 1 , Donghyeon Lee 1 ,<br />

Young Youn Cho 1 , Eun Ju Cho 1 , Jeong-Hoon Lee 1 , Yoon Jun Kim 1 ,<br />

June Sung Lee 2 , Jung-Hwan Yoon 1 , Su Jong Yu 1 ; 1 Department<br />

of Internal Medicine and Liver Research Institute, Seoul National<br />

University College of Medicine, Seoul, Korea (the Republic of);<br />

2 Department of Internal Medicine, Ilsan Paik Hospital, Inje University<br />

College of Medicine, Goyang, Korea (the Republic of)<br />

Background/Aim: The presence of portal vein tumor thrombosis<br />

(PVTT) is a poor prognostic factor for patients with hepatocellular<br />

carcinomas (HCC). HCC patients with PVTT are considered<br />

advanced stage regardless of the tumor factor and only<br />

sorafenib is recommended for those patients by current guidelines.<br />

This study was aimed to assess whether HCC patients<br />

with PVTT could have better prognosis with proper treatment if<br />

the tumor factors were favorable. Methods: This single-center<br />

retrospective study involved 1,184 patients diagnosed HCC<br />

from January 2005 to December 2006, before sorafenib<br />

era. Overall survival (OS) was estimated by the Kaplan-Meier<br />

method and Cox proportional hazard model. To control the<br />

different distributions of covariates of the two groups with or<br />

without PVTT, matched patients were selected by propensity<br />

score analysis. Results: Among the patients with PVTT, the proportion<br />

of diffuse infiltrative type was significantly higher than<br />

nodular type (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 451A<br />

480<br />

Value of PIVKA II and AFP serum level kinetics for treatment<br />

monitoring in patients with hepatocellular carcinoma<br />

(HCC): a French pilot study<br />

Audrey Payance 1 , Mohamed Bouattour 1 , Mohamed Achahboun 2 ,<br />

Miguel Albuquerque 2 , Laurent Castera 1 , Pierre Bedossa 2 , Olivier<br />

Soubrane 3 , Francois Durand 1 , Valerie Paradis 2 ; 1 HEPATOLOGY,<br />

BEAUJON HOSPITAL, Clichy, France; 2 Pathology, BEAUJON<br />

HOSPITAL, Clichy, France; 3 HPB Surgery and Liver Transplan,<br />

BEAUJON HOSPITAL, Clichy, France<br />

Background and Aims: In Asia, prothrombin induced by vitamin<br />

K absence-II (PIVKA-II) is largely used as a diagnostic and<br />

prognostic biomarker for hepatocellular carcinoma (HCC).<br />

We already confirmed the diagnostic value of PIVKA-II for<br />

European patients with early HCC. Herein, our aim was to<br />

study the dynamics of PIVKA-II serum level in the follow-up of<br />

European patients with HCC, according to treatment monitoring.<br />

Material and Methods: Patients with histologically proven<br />

HCC diagnosed in our center from January 2013 to August<br />

2014 were included. For each patient, serum PIVKA-II and<br />

afetoprotein (AFP) levels were assessed with an enzyme-linked<br />

immunosorbent assay at two different times. We defined two<br />

groups of patients: 1) “treated patients” group with first serum<br />

sample obtained within two months before treatment (D0) and<br />

the second within 2-3 months after treatment initiation (M2-3),<br />

2) “non-treated patients” as a control group, with two serum<br />

samples, at diagnosis (T1) and 2-3 months subsequently (T2).<br />

Baseline serum levels of AFP and PIVKA-II and dynamics during<br />

follow-up were compared between the two groups. Results are<br />

presented as median (interquartile range). Results: A total of<br />

32 patients with HCC developed on chronic liver disease were<br />

included (men, 85 %; median age, 59 years; BCLC stage: A<br />

47% - B 28%- C 19% - D 6%). Aetiologies of underlying liver<br />

disease were alcohol consumption, viral hepatitis B, viral hepatitis<br />

C, non alcoholic steato-hepatitis and genetic hemochromatosis<br />

in 14 (44%) ; 8 (25%), 13 (41%), 4 (12%), 2 (6%) of<br />

patients respectively. In 23 patients, HCC treatment was performed<br />

including chemoembolization (n=14), radiofrequency<br />

ablation (n=3), surgical resection (n=1), sorafenib (n=4), and<br />

radioembolisation (n=1). For all patients, median baseline AFP<br />

and PIVKA II serum levels were 11.6 ng/ml (4.8 - 139.8 ng/<br />

ml) and 416 mAU/ml (100.5 - 5894.25 mAU/ml), respectively.<br />

In the “treated patients” group, PIVKA II serum levels<br />

decreased statistically significantly from D0 to M2-3 (D0: 811<br />

mAU/ml (100.5-11559.75) vs M2-3: 292 mAU/ml (46.25<br />

- 7391); p = 0.031)). However, no significant decrease was<br />

observed regarding AFP serum levels between D0 and M2-3<br />

(D0: 18.9 ng/ml (4.8-392) vs M2-3: 16.3 ng/ml (4.5-103);<br />

p=0.831). Finally, no changes of AFP and PIVKA II serum levels<br />

from baseline T1 to T2 were observed in the control group.<br />

Conclusions: Significant decrease of PIVKA II serum levels following<br />

treatment of HCC was observed. These results suggest<br />

that this biomarker, in contrast to AFP, could be useful for monitoring<br />

treatment in patients with HCC. Further investigations are<br />

needed to validate these results.<br />

Disclosures:<br />

Mohamed Bouattour - Advisory Committees or Review Panels: Bayer Schering<br />

Pharma; Speaking and Teaching: Bayer Schering Pharma<br />

Laurent Castera - Speaking and Teaching: Gilead, BMS, Janssen, Echosens,<br />

Abbvie, Biopredictive<br />

Francois Durand - Advisory Committees or Review Panels: Astellas, Novartis,<br />

BMS; Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Audrey Payance, Mohamed<br />

Achahboun, Miguel Albuquerque, Pierre Bedossa, Olivier Soubrane, Valerie<br />

Paradis<br />

481<br />

Liver-directed (LD) Therapy for Intermediate Stage<br />

Hepatocellular Carcinoma: Survival Across Treatment<br />

Strategies<br />

Alvaro E. Castillo 1 , Willscott E. Naugler 2 , Kenneth J. Kolbeck 3 ,<br />

Khashayar Farsad 3 , Susan L. Orloff 1 , Kevin Billingsley 4 , C. Kristian<br />

Enestvedt 1 ; 1 Surgery, Oregon Health and Science University,<br />

Portland, OR; 2 Internal Medicine, Oregon Health and Science<br />

University, Portland, OR; 3 Dotter Interventional Institute, Oregon<br />

Health and Science University, Portland, OR; 4 Division of Surgical<br />

Oncology, Oregon Health and Science University, Portland, OR<br />

Treatment strategies favor LD options for BCLC intermediate<br />

stage HCC. Evaluate the outcomes of different multidisciplinary<br />

treatment strategies in BCLC B patients. A retrospective cohort<br />

analysis of HCC patients referred to our institutional Multidisciplinary<br />

Liver Tumor Board. All with Barcelona Clinic Liver<br />

Cancer (BCLC) stage B who underwent either trans-arterial<br />

chemoembolization (TACE), radiation therapy (XRT)/TACE,<br />

Yttrium-90 (Y-90) alone and Y-90/TACE were selected. Resection,<br />

ablation or transplantation was excluded. Univariate analysis<br />

was used to compare demographics, tumor characteristics,<br />

and treatment modalities. Multivariate analysis of clinical factors<br />

associated with survival was performed using Cox proportional<br />

hazards modeling. Survival was compared using the<br />

Kaplan Meier log-rank method. Eighty eight BCLC B patients<br />

were included. Comparison by treatment groups showed<br />

no difference in severity of disease or demographics except<br />

for bilirubin which was higher for the TACE and TACE/XRT<br />

groups when compared to Y-90 and Y-90/TACE (p=0.024).<br />

Those who received Y-90 had larger tumors (median 6.5 cm,<br />

p=


452A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Alvaro E. Castillo, Willscott<br />

E. Naugler, Kenneth J. Kolbeck, Susan L. Orloff, Kevin Billingsley, C. Kristian<br />

Enestvedt<br />

482<br />

Analyses of progressive disease patients with advanced<br />

hepatocellular carcinoma treated with sorafenib: prognosis,<br />

treatment, and progression pattern<br />

Keiko Yamane, Takahide Nakazawa, Wasaburo Koizumi; Gastroenterology,<br />

Kitasato University School of Medicine, Sagamihara,<br />

Japan<br />

Purpose: Progressive disease (PD) is highly observed in patients<br />

with advanced hepatocellular carcinoma (HCC) treated with<br />

sorafenib. Here, we examined PD patients about progression<br />

pattern, treatment for PD, and the prognosis. Methods: PD<br />

patients radiologically assessed using RECIST 1.1. were retrospectively<br />

examined. Overall survival (OS), Time to progression<br />

(TTP), post-progression survival (PPS), progression pattern,<br />

and prognostic factors were analyzed. Progression pattern was<br />

classified as follows (i-v): (i) intrahepatic increase in tumor size<br />

(ii) extrahepatic increase in tumor size (iii) new intrahepatic<br />

lesion (iv) new extrahepatic lesion (v) new intra/extrahepatic<br />

lesion. Result: One-hundred and eight patients were treated<br />

with sorafenib, and 102 were radiologically assessed. PD was<br />

observed in 56 patients (median age 69 year old, hepatitis<br />

C virus 50%, Child-Pugh [C-P] A 98%, and BCLC-C 61%).<br />

Twenty-four patients of PD (43%) had invasion to intrahepatic<br />

major vessels. The median OS, TTP, and PPS were 7, 1.9,<br />

and 5 months, respectively. Progression pattern was (i) in 12<br />

patients, (ii) in 1, (iii) in 30, (iv) in 6, (v) in 7. The median OS<br />

by progression pattern was 7.5 months in (i), 15 in (ii), 7.1 in<br />

(iii), 6.9 in (iv), and 2.6 in (v). The pattern of new intra/extrahepatic<br />

lesion showed significantly poor prognosis. C-P score<br />

of 5 point (P = 0.004, OR 0.402, 95% CI 0.214-0.753) and<br />

invasion to intrahepatic major vessels (P = 0.017 OR 2.119,<br />

95% CI: 1.142-3.932), at the beginning of sorafenib treatment<br />

were independent prognositic factors of OS by a Cox proportion<br />

hazard model. While, C-P A (P = 0.036, OR 0.450, 95%<br />

CI 0.213-0.947), invasion to vessels (P = 0.008 OR 2.605,<br />

95% CI: 1.278-5.306) and new intra/extrahepatic lesion (P =<br />

0.004, OR 4.517, 95% CI: 1.618-12.608) were independent<br />

factors of PPS by a Cox model. Second line treatment for PD<br />

determined the good prognosis. Median PPS was 6.2 months<br />

in PD patients with 2nd line treatment, while median PPS of 2.5<br />

months without any treatment. C-P A of hepatic function was<br />

statistically observed in PD patients with 2nd line treatment by<br />

Logistic-regression analysis (P = 0.046, OR 4.317, 95% CI<br />

0.215-18.183). Conclusion: Hepatic function and progression<br />

pattern were important factors to determine the prognosis in PD<br />

patients. Second line treatment based on these findings should<br />

be considered.<br />

Disclosures:<br />

The following authors have nothing to disclose: Keiko Yamane, Takahide Nakazawa,<br />

Wasaburo Koizumi<br />

483<br />

Impact of Viral Hepatitis Etiology on Survival Outcomes<br />

in Patients with Hepatocellular Carcinoma: A Large Multicentre<br />

Study<br />

Sara Mgaieth 1 , William W. Kemp 1 , Paul Gow 2 , Michael A. Fink 3 ,<br />

Virginia H. Knight 4 , Anouk T. Dev 4 , Amanda J. Nicoll 5 , John<br />

Lubel 5 , Sally Bell 6 , Thai Hong 6 , Marno C. Ryan 6 , Siddharth Sood 7 ,<br />

Stuart K. Roberts 1 ; 1 Gastroenterology, The Alfred Hospital, Melbourne,<br />

VIC, Australia; 2 Liver Transplant Unit, The Austin Hospital,<br />

Melbourne, VIC, Australia; 3 Department of Surgery, The Austin<br />

Hospital, Melbourne, VIC, Australia; 4 Gastroenterology, Monash<br />

Medical Center, Melbourne, VIC, Australia; 5 Gastroenterology,<br />

Eastern Health, Melbourne, VIC, Australia; 6 Gastroenterology, St<br />

Vincent’s Hospital, Melbourne, VIC, Australia; 7 Gastroenterology,<br />

The Royal Melbourne Hospital, Melbourne, VIC, Australia<br />

BACKGROUND: Hepatocellular carcinoma is the fifth most common<br />

cancer worldwide in men, and the second leading cause of<br />

death. The incidence of HCC is increasing, with HBV and HCV<br />

infections accounting for more than 50% of cases. Although<br />

HBV and HCV are known risk factors for HCC it is unclear if<br />

there is an overall survival difference between patients with<br />

viral and non-viral hepatitis related HCC. AIM: The aim of the<br />

study is to compare the overall survival and recurrence free survival<br />

of patients with viral and non-viral hepatitis related HCC<br />

and between those with viral hepatitis B and C related HCC.<br />

METHODS: A multi-centre study was conducted on HCC cases<br />

from 6 large tertiary hospitals from 1994-2014. Patient demographics,<br />

severity of liver disease, tumour characteristics, and<br />

patient outcomes were retrieved from the HCC database and<br />

computer records. Survival outcomes were compared between<br />

the three cohorts of HBV, HCV and non-viral related HCC and<br />

predictors of survival were determined using Cox proportional<br />

hazards regression. RESULTS: There were a total of 1551 cases<br />

of HCC between 1994-2014. 17% (265) had HBV, 35%<br />

(541) had HCV, 30 (3%) co-infected. There was a total of 50%<br />

(776) with viral hepatitis related HCC, and 43% (661) with<br />

non-viral hepatitis related HCC, and 7%(114) unknown. On<br />

univariate analysis, patients with viral hepatitis related HCC<br />

had better overall survival compared to those with non-viral<br />

hepatitis related HCC (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 453A<br />

The following authors have nothing to disclose: Sara Mgaieth, William W.<br />

Kemp, Paul Gow, Michael A. Fink, Virginia H. Knight, Anouk T. Dev, John Lubel,<br />

Thai Hong, Marno C. Ryan<br />

484<br />

Stereotactic body radiation therapy for recurrent hepatocellular<br />

carcinoma as a local salvage treatment<br />

Baek Gyu Jun 1 , Sae Hwan Lee 1 , Hong Soo Kim 1 , Sang Gyune<br />

Kim 3 , Young Seok Kim 3 , Boo Sung Kim 3 , Soung Won Jeong 4 ,<br />

Jae Young Jang 4 , Young Don Kim 2 , Gab Jin Cheon 2 ; 1 Department<br />

of Internal Medicine, Soonchunhyang University College of<br />

Medicine Cheonan Hospital, Cheonan-si, Korea (the Republic of);<br />

2 Internal Medicine, University of Ulsan College of Medicine, Gangneung,<br />

Korea (the Republic of); 3 Department of Internal Medicine,<br />

Soonchunhyang University College of Medicine Bucheon Hospital,<br />

Bucheon, Korea (the Republic of); 4 Department of Internal Medicine,<br />

Soonchunhyang University College of Medicine Seoul Hospita,<br />

Seoul, Korea (the Republic of)<br />

BACKGROUND: The aim of this study was to evaluate the<br />

efficacy and toxicity of stereotactic body radiation therapy<br />

(SBRT) as a local salvage treatment for small, recurrent hepatocellular<br />

carcinoma(HCC). METHODS: Between March 2011<br />

and February 2014, 32 patients with recurrent HCC(diameter<br />


454A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

486<br />

Clinical Features and Treatment Modalities of Long-Term<br />

Survivors with Advanced Hepatocellular Carcinoma: A<br />

Single Institution Cohort Study<br />

Kyung Hee Kim 1 , Joong-Won Park 1,2 , Bo Hyun Kim 1 , Chang-Min<br />

Kim 1 , Byung Ho Nam 2 ; 1 Center for Liver Cancer, National cancer<br />

center, Go yang si, Korea (the Democratic People’s Republic of);<br />

2 Department of Cancer Control and Policy, Graduate School of<br />

Cancer Science and Policy, National Cancer Center, Go yang si,<br />

Korea (the Democratic People’s Republic of)<br />

Background and Aim: The prognosis of patients with advanced<br />

hepatocellular carcinoma (HCC) is poor despite systemic or<br />

local treatments. We analyzed the characteristics of long-term<br />

survivors (LTS) with advanced HCC to determine the prognostic<br />

and treatment factors responsible for long-term survival. Methods:<br />

Between January 2001 and December 2010, 543 patients<br />

were first diagnosed as having modified International Union<br />

Against Cancer (mUICC) stage IV HCC, with Child-Pugh (CP)<br />

class A liver function at the National Cancer Center, Korea.<br />

The authors retrospectively reviewed all medical records and<br />

radiographs. Among the patients, we compared clinical features<br />

and treatment modalities of the LTS group (survival > 24<br />

months; n = 66) with those of the control (C) group (survival < 6<br />

months; n = 333). Results: For 543 patients, the median overall<br />

survival (OS) was 4.6 months (95% CI: 4.2-5.0): 37.7 months<br />

(95% CI: 34.4-41.0) in the LTS group and 3.1 months (95%<br />

CI: 2.9-3.3) in the C group. Compared to the C group, LTS<br />

group patients were significantly more likely to be older and<br />

to have non-hepatitis B virus etiology, lower tumor numbers,<br />

smaller tumor size, nodular-type tumors, absence of portal vein<br />

invasion, and better liver function. Ill-defined tumor type, portal<br />

vein invasion and CP A6 proved to be independent prognostic<br />

variables (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 455A<br />

Patients characteristics<br />

fatty liver disease. Overall, the percentage of non-cirrhotic,<br />

non-HBV HCC was 2.93%, non-cirrhotic, non-HBV HCC with<br />

hepatitis C was 1.66%, and non-cirrhotic, non-HBV HCC with<br />

fatty liver disease was 6.6%. Conclusions Non-HBV HCC can<br />

arise in the setting of chronic hepatitis C and fatty liver disease<br />

without advanced fibrosis or cirrhosis. Our prevalence rates<br />

may underestimate the true prevalence owing to the exclusion<br />

of patients without adequate histology available.<br />

Disclosures:<br />

Steven-Huy B. Han - Grant/Research Support: Bristol Myers Squibb, Gilead<br />

The following authors have nothing to disclose: Vivian Ng, Jennifer Phan, Alan<br />

J. Sheinbaum<br />

Disclosures:<br />

Rene Adam - Grant/Research Support: Merck Serono, Roche, Sanofi aventis,<br />

astellas, novartis, Merck Serono, Roche, Sanofi aventis, astellas, novartis, Merck<br />

Serono, Roche, Sanofi aventis, astellas, novartis, Merck Serono, Roche, Sanofi<br />

aventis, astellas, novartis<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

The following authors have nothing to disclose: Eleonora De Martin, Michel<br />

Rayar, Damien Bergeat, Luis Filipe Abreu de Carvalho, Maximiliano Gelli, Denis<br />

X. Castaing, Daniel Cherqui, Antonio Sa-Cunha, Emmanuel Boleslawski, Karim<br />

Boudjema, Eric Vibert<br />

488<br />

Non-Cirrhotic Hepatocellular Carcinoma in the Absence<br />

of Hepatitis B<br />

Vivian Ng 1 , Jennifer Phan 1 , Alan J. Sheinbaum 2 , Steven-Huy B.<br />

Han 1 ; 1 Pfleger Liver Institute, University of California, Los Angeles,<br />

Los Angeles, CA; 2 Gastroenterology, VA West Los Angeles Healthcare<br />

Center, Los Angeles, CA<br />

Purpose Hepatocellular carcinoma (HCC) accounts for approximately<br />

90% of primary hepatic malignancies worldwide.<br />

In hepatitis B virus (HBV) infection, HCC can develop in the<br />

absence of cirrhosis. However, it is thought that in chronic<br />

hepatitis C virus infection and fatty liver disease, development<br />

of cirrhosis is necessary prior to formation of HCC. To date,<br />

there have only been a small number of case reports showing<br />

that HCC can develop in patients with chronic hepatitis C and<br />

in fatty liver disease without evidence of underlying cirrhosis.<br />

We sought to evaluate whether or not non-cirrhotic HCC exists<br />

and assess underlying risk factors contributing to its development.<br />

Methods We performed a single-center, retrospective<br />

chart review study at a large academic tertiary care hospital.<br />

All patients seen in both the liver cancer clinic and orthotopic<br />

liver transplantation (OLT) clinic for liver cancer in the last 10<br />

years were evaluated. We determined how many patients<br />

had non-cirrhotic HCC based on available pathology reports,<br />

such as explanted liver pathology, liver resection pathology,<br />

and liver biopsy pathology. Patients with hepatitis B, non-<br />

HCC tumors, and patients without both tumor and non-tumor<br />

liver histology available were excluded. Demographic information,<br />

pertinent lab values, and co-morbid conditions were<br />

recorded. Results 1925 unique patients were identified. 1382<br />

patients were excluded from the study, including 303 patients<br />

with hepatitis B, 510 patients with non-HCC tumors, and 569<br />

patients without available tumor and non-tumor histology. Of<br />

the remaining patients, 491 patients underwent (OLT), 70<br />

underwent local ablative therapy, 42 underwent resection of<br />

the HCC, and 17 underwent systemic therapy. All 491 OLT<br />

patients had underlying cirrhosis. Of the 134 patients who<br />

did not undergo (OLT), 35 patients (26.1%) were identified as<br />

having non-cirrhotic HCC. 13 patients (37.1%) had an underlying<br />

diagnosis of hepatitis C, and 6 patients (17.1%) had<br />

489<br />

Albi score predicts survival in patients with BCLC 0/A<br />

stage Hepatocellular Carcinoma (HCC) independently of<br />

Child Pugh (CP) score and treatment allocation<br />

Simona Onali 1 , Aileen Marshall 1 , Dinesh Sharma 1 , Pam O’Donoghue<br />

1 , Emily Dannhorn 1 , Phillip Johnson 2 , James O’Beirne 1 ;<br />

1 Sheila Sherlock Liver Centre, Royal Free Hospital, London, United<br />

Kingdom; 2 Department of Molecular and Clinical Cancer Medicine,<br />

University of Liverpool, Liverpool, United Kingdom<br />

Introduction: BCLC 0/A stage HCC patients are considered<br />

candidates for curative treatments. Within the BCLC classification,<br />

liver function is assessed by CP score despite this score<br />

not being derived from patients with HCC. ALBI is a novel<br />

evidence based score derived from a large cohort of HCC<br />

patients and has been shown to define different prognostic<br />

groups even within CP A. 1 We studied the performance of ALBI<br />

score in a well characterised cohort of patients with BCLC 0/A<br />

stage HCC to determine if ALBI score impacted on prognosis.<br />

Methods: Consecutive patients with BCLC 0/A HCC, evaluated<br />

between 2010-2014 at the Royal Free Hospital were included.<br />

All patients had a performance score of 0 and were undergoing<br />

evaluation for potential resection. Patients were considered<br />

resectable if they had HVPG≤10mmHg or HVPG≥10mmHg<br />

with good liver function as assessed by ICG clearance. ALBI<br />

score was calculated at the time of HVPG and ICG clearance<br />

measurement and divided into 3 grades as per published cutoffs.<br />

Kaplan Meier and Cox regression analysis were used to<br />

assess the impact of ALBI on survival. Results: 67 patients were<br />

included: 64 (95.5%) were CP A, 3 (4.5%) were CP B. 60<br />

(89%) had uninodular tumour and 7 (10.4%) multifocal (≤3<br />

tumours ≤30mm). 46 (69%) underwent resection, 4 (6%) RFA,<br />

9 (14%) TAE, 3 (5%) OLT, 3 (5%) either no treatment or undergoing<br />

evaluation. Median follow up was 9 months (13-54).<br />

10 (15%) patients died during follow up. Cause of death was<br />

HCC-related n= 4, liver failure n=4, non-liver related n=2. In<br />

Kaplan Meier analysis CP score was not associated with difference<br />

in survival. Cox regression analysis revealed ALBI score as<br />

the only significant factor related to survival and was independent<br />

of CP score and treatment allocation (p=0.002, OR=15,<br />

95%CI=3-80). Of patients with ALBI grade 1, 44 (83%) were<br />

considered suitable and underwent resection vs only 2 (14%)<br />

with ALBI 2. One-year survival was 88% in patients with ALBI<br />

grade 1 and 43% in patients with ALBI grade 2 (p=0.002).<br />

Conclusion: In patients with BCLC 0/A stage HCC ALBI score<br />

clearly defines patients with different prognosis regardless<br />

of Child Pugh score and treatment allocation. ALBI grade 2<br />

patients are very rarely suitable candidates for resection. ALBI<br />

score is more informative than Child Pugh score in patients with<br />

early stage HCC and may be useful in determining treatment<br />

allocation References: 1 . Johnson PJ et al. Assessment of Liver<br />

Function in Patients With Hepatocellular Carcinoma: A New


456A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Evidence-Based Approach-The ALBI Grade. J Clin Oncol. 2015<br />

Feb 20;33(6):550-8<br />

Disclosures:<br />

The following authors have nothing to disclose: Simona Onali, Aileen Marshall,<br />

Dinesh Sharma, Pam O’Donoghue, Emily Dannhorn, Phillip Johnson, James<br />

O’Beirne<br />

490<br />

Albi score predicts survival independently of hepatic<br />

venous pressure gradient (HVPG) and indocyanine<br />

green (ICG) clearance in HCC patients undergoing resection<br />

Aileen Marshall 1 , Simona Onali 1 , Dinesh Sharma 1 , Pam O’Donoghue<br />

1 , Emily Dannhorn 1 , Phillip Johnson 2 , James O’Beirne 1 ;<br />

1 Sheila Sherlock Liver Centre, Royal Free Hospital, London, United<br />

Kingdom; 2 Department of Molecular and Clinical Cancer Medicine,<br />

University of Liverpool, Liverpool, United Kingdom<br />

Introduction: A new evidence based model, the Albumin-Bilirubin<br />

(ALBI) grade † , has been recently proposed for assessing<br />

liver function in patients with hepatocellular carcinoma<br />

(HCC). We aimed to evaluate the impact of ALBI in predicting<br />

decompensation and survival after liver resection (LR) for<br />

HCC. Method: Consecutive patients undergoing resection for<br />

HCC between 2011-2014 at the Royal Free Hospital were<br />

evaluated. Demographic, clinical data and histopathological<br />

features of resected tumour were collected. Patients underwent<br />

HVPG and ICG clearance measurement to evaluate potential<br />

resectability. ALBI score was calculated pre and post-operatively.<br />

Patients were divided into 3 categories (ALBI 1,2,3)<br />

according to published ALBI score cut-offs. Cox regression was<br />

used to identify predictors of decompensation and survival post<br />

LR. Results: 48 patients were included with a median post LR<br />

follow up 17 months (1-54): male 42(87.5%), mean age 63<br />

years (28-83). Median HVPG 6 mmHg(2-15), median ICG PDR<br />

18.2 (6.1-29.4), median ICG R15 6.5 (1.2-19.2). All patients<br />

had a Child Pugh score A5. 46 (96%) had a pre-LR ALBI grade<br />

1 and 2 (4%) had an ALBI grade 2. Clinically significant portal<br />

hypertension (HVPG ≥10mmHg) was found in 11 (23%)<br />

patients. Eight (19%) patients had an ICG PDR 15. Thirty (62,5%) patients underwent<br />

an anatomical resection, while 18 (37,5%) had a wedge<br />

resection. Post-LR decompensation was observed in 5(10%)<br />

patients: ascites n=4, encephalopathy n=1 at a median of 10<br />

days post-LR. 12(25%) patients had HCC recurrence after a<br />

mean time of 11 months (5-28). 7 (15%) died after a mean<br />

follow up of 15 months (1-40), 3 of them due to tumour recurrence.<br />

Patients with 1 HCC


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 457A<br />

Inherited severe unconjugated hyperbilirubinemia is known as<br />

Crigler-Najjar syndrome (CN). CN is characteristically caused<br />

by deficiency of UDP-glucuronosyl transferase 1A1 (UGT1A1),<br />

encoded by the UGT1A1 gene. This enzyme is responsible<br />

for glucuronidation of bilirubin needed to excrete this toxic<br />

compound via the bile. The severity of inherited unconjugated<br />

hyperbilirubinemia depends on the residual UGT1A1 activity<br />

and can be described as a spectrum from complete absence<br />

(CN type 1) or some activity (CN type 2) to 30% of residual<br />

activity as seen in the benign Gilbert syndrome. In CN<br />

patients with residual UGT1A1 activity a reduction of serum<br />

bilirubin levels can be achieved by phenobarbital administration<br />

that results in transcriptional induction of UGT1A1 via<br />

activation of the Constitutive Androstane Receptor (CAR). We<br />

analyzed the UGT1A1 promoter sequence of a patient with<br />

severe unconjugated hyperbilirubinemia suspected for CN type<br />

2, but unresponsive to phenobarbital treatment and without<br />

clarifying mutation in the UGT1A1 coding region. Besides an<br />

extra TA repeat in the TATA box, we disclosed a homozygous<br />

three nucleotide (3nt) insertion on position -85 to -83 of the<br />

proximal promoter, corresponding with the HNF1α binding<br />

site. We generated UGT1A1 promoters with all different mutations<br />

and cloned them upstream of a luciferase reporter gene<br />

to perform functional <strong>studies</strong>. The insertion of this 3nt in the<br />

HNF1α binding site resulted in a >95% reduction of promoter<br />

activity, independent of the extra TA-repeat in the TATA-box. In<br />

addition, this insertion rendered the promoter unresponsive to<br />

the induction by phenobarbital via CAR and rifampicin via the<br />

Pregnane X Receptor (PXR), thereby explaining the intermediate<br />

CN phenotype of this patient. In conclusion, disruption of the<br />

HNF1α binding site in the promoter of a liver specific gene<br />

such as UGT1A1 is for the first time demonstrated to have a<br />

major impact on gene expression in man and may result in<br />

severe liver disorder.<br />

Disclosures:<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

Ulrich Beuers - Consulting: Intercept via University of Amsterdam, Novartis via<br />

University of Amsterdam; Grant/Research Support: Falk, Zambon; Speaking and<br />

Teaching: Falk Foundation, Gilead, Roche, Shire<br />

The following authors have nothing to disclose: Remco van Dijk, Isabel<br />

Mayayo-Peralta, Sem J. Aronson, Anja A. Kattentidt-Mouravieva, Vincent van<br />

der Mark, Nevin Oruç, Piter J. Bosma<br />

493<br />

MicroRNA-122 (miR-122) Regulates Polyploidy in the<br />

Liver<br />

Shu-hao Hsu 1 , Evan Delgado 1 , P. A. Otero 1 , Kolin Meehan 1 , Justin<br />

B. Moroney 1 , Kun-Yu Teng 2 , Kalpana Ghoshal 2 , Andrew W.<br />

Duncan 1 ; 1 Department of Pathology, McGowan Institute for Regenerative<br />

Medicine, University of Pittsburgh, Pittsburgh, PA; 2 Department<br />

of Pathology, Comprehensive Cancer Center, The Ohio State<br />

University, Columbus, OH<br />

A defining feature of the mammalian liver is polyploidy, a<br />

numerical change in the entire complement of chromosomes.<br />

Hepatocytes are either mononucleate or binucleate, and<br />

ploidy is determined by the number of nuclei/cell, as well as<br />

the ploidy of each nucleus. The first step of polyploidization<br />

involves cell division with failed cytokinesis. Although polyploidy<br />

is common, affecting ~90% of hepatocytes in mice and<br />

50% in humans, the specialized role played by polyploid cells<br />

in liver homeostasis and disease remains poorly understood.<br />

The goal of this study was to identify novel signals that regulate<br />

polyploidization, and we focused on microRNAs (miR-<br />

NAs). First, to test whether miRNAs could regulate hepatic<br />

polyploidy we examined livers from Dicer1 knockout mice,<br />

which are devoid of mature miRNAs. Loss of miRNAs resulted<br />

in a 3-fold reduction in binucleate hepatocytes, indicating that<br />

miRNAs could indeed regulate polyploidization. Secondly, we<br />

surveyed age-dependent expression of >500 miRNAs (NanoString)<br />

in wild type mice and identified a subset of miRNAs,<br />

including miR-122, differentially expressed at 2-3 weeks, a<br />

period when extensive polyploidization occurs. Thirdly, we<br />

examined Mir122 knockout mice and observed profound, lifelong<br />

depletion of polyploid hepatocytes, proving that miR-122<br />

is required for complete hepatic polyploidization. Next, we<br />

identified direct targets of miR-122, Cux1, Iqgap1, Mapre1,<br />

Nedd4l and Slc25a34, that regulate cytokinesis. Inhibition of<br />

each target induced cytokinesis failure and promoted hepatic<br />

binucleation. Finally, to validate the relevance of these targets<br />

to liver disease, we examined expression in a subset of human<br />

hepatocellular carcinomas (HCC) with reduced miR-122. Consistent<br />

with the mouse data, target expression was inversely<br />

proportional to miR-122. In summary, our data suggest a novel<br />

regulatory role for miR-122 in liver polyploidization. Moreover,<br />

differential expression of miR-122 targets in HCC provides new<br />

insights into miR-122-mediated tumorigenesis. These <strong>studies</strong><br />

will serve as the foundation for future work investigating miR-<br />

122 and polyploidy in lipid metabolism, viral infection, tumorigenesis<br />

and regeneration.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shu-hao Hsu, Evan Delgado, P.<br />

A. Otero, Kolin Meehan, Justin B. Moroney, Kun-Yu Teng, Kalpana Ghoshal,<br />

Andrew W. Duncan<br />

494<br />

Overexpression Screening to Discover Gene-Gene Interactions<br />

that Drive Hepatocellular Carcinoma<br />

Monica Teta-Bissett 1 , Kirk J. Wangensteen 1,2 , Klaus H. Kaestner 1 ;<br />

1 Genetics, University of Pennsylvania, Philadelphia, PA; 2 Gastroenterology,<br />

University of Pennsylvania, Philadelphia, PA<br />

Hepatocellular carcinoma (HCC) is the third leading cause<br />

of cancer death worldwide, and one of the top ten in the<br />

United States according to the Global Cancer Statistics and<br />

the US CDC, respectively. Sorafenib, a small molecule targeting<br />

receptor tyrosine kinases, is the only drug approved for<br />

treatment of advanced HCC in the United States. In order to<br />

determine the relative contribution of HCC candidate genes in<br />

HCC formation and growth, we developed an overexpression<br />

model to screen multiple gene combinations simultaneously<br />

during liver tumor formation. A pooled plasmid library of 34<br />

constructs, each designed to express the FAH gene plus a different<br />

cancer-related gene of interest, was delivered into the<br />

liver of Fah-null mice by hydrodynamic tail vein injection. The<br />

constructs allowed for constitutive expression of FAH, rescuing<br />

the tyrosinemia of Fah-null mice, and creating repopulation<br />

nodules overexpressing random combinations of the genes<br />

within the plasmids. The mouse livers were riddled with tumors<br />

at four and six months following library injection. More than<br />

20 unique tumors were dissected and the tumor phenotypes<br />

were characterized using hematoxylin and eosin staining and<br />

immunohistochemistry for Osteopontin (Opn) on serial sections.<br />

DNA was extracted from unique tumors and, by high-throughput<br />

sequencing of unique 5-nucleotide barcodes attached to<br />

each of the expression plasmids, the profiles of the exogenous<br />

DNA constructs was compiled. We found that c-myc<br />

was linked to all of the tumors that were tested, indicating<br />

that it was a consistent driver of tumor growth in this system.<br />

TGFα, Foxa3, Akt, Met and Nr1h3 were among other factors<br />

detected to drive tumor growth. Interestingly, the combination


458A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of four genes, c-myc, AKT, MET and Nr1h3, correlated with a<br />

subset of Opn-positive HCC tumors. This innovative paradigm<br />

allows for prospective testing of millions of combinations of<br />

overexpressed factors to determine the interaction of known<br />

and suspected liver oncogenes and tumor suppressors in a<br />

systematic fashion.<br />

Disclosures:<br />

The following authors have nothing to disclose: Monica Teta-Bissett, Kirk J. Wangensteen,<br />

Klaus H. Kaestner<br />

495<br />

Peretinoin, an Acyclic Retinoid, Inhibits Hepatocarcinogenesis<br />

through Suppression of Sphingosine Kinase 1<br />

Expression In vitro and In vivo<br />

Masaya Funaki 1 , Tetsuro Shimakami 1 , Tsuguhito Ota 2 , Masao<br />

Honda 1 , Takayoshi Shirasaki 1 , Shuichi Kaneko 1 ; 1 Disease Control<br />

and Homeostasis, Kanazawa university, Kanazawa, Japan;<br />

2 Brain/Liver Interface Medicine Research Center, Kanazawa University,<br />

Kanazawa, Japan<br />

[Objective] A sphingolipid, sphingosine-1-phospate (S1P) is<br />

a potent bioactive lipid which can regulate carcinogenesis<br />

and cancer progression. Both sphingosine kinase 1 (SPHK1)<br />

and SPHK2 are the essential kinases that produce S1P. Therefore,<br />

SPHK can be a therapeutic target by crucially regulating<br />

sphingolipid metabolism in several kinds of cancer. Peretinoin<br />

(NIK-333), an acyclic retinoid, was reported to inhibit the post<br />

therapeutic recurrence of hepatocellular carcinoma (HCC) in<br />

patients with chronic hepatitis C. However, the mechanism of<br />

its inhibitory effects against recurrent HCC remains unclear.<br />

We hypothesized that peretinoin could prevent hepatocarcinogenesis<br />

by modifying a SPHK-S1P axis. In the present study,<br />

we assessed the effect of peretinoin on SPHK activation and<br />

development of liver cancer in vivo and in vitro. [Method] We<br />

examined the effect of peretinoin on the expression and the<br />

enzymatic activity of SPHK1 in a human hepatoma cell line,<br />

Huh-7 cells. Next, using a liver cancer mice model with atherogenic<br />

high fat (AHF) diet-induced NASH we administrated<br />

an AHF diet with and without peretinoin (0.03%) for 48 weeks<br />

and examined the effect of peretinoin on hepatocarcinogenesis<br />

and expression of SPHK1. Finally, we clarified the effect of<br />

SPHK1 on hepatocarcinogenesis induced by Diethylnitrosoamine<br />

(DEN) using SPHK1 knockout mice. [Results] [In vitro] The<br />

treatment with peretinoin (10 to 40 μM) reduced the mRNA<br />

and protein expression of SPHK1 in Huh-7 cells in a time- and<br />

dose- dependent manner. However, peretinoin did not change<br />

the expression of SPHK2 in Huh-7 cells. Furthermore, peretinoin<br />

markedly reduced the enzymatic activity of SPHK1 assessed by<br />

in vitro 32P labeled SPHK activity assay. Next, we performed<br />

reporter gene assays by using constructs containing the SPHK1<br />

promoter region. Peretinoin reduced the promoter activity of<br />

SPHK1, and overexpression of SP1 restored promoter activity.<br />

In addition, three deletion constructs with and without SP1<br />

binding sites were created. The deletion constructs without SP1<br />

binding sites abolished the promoter activity. Interestingly, we<br />

previously reported that peretinoin suppresses the expression of<br />

SP1. Collectively, peretinoin reduced the expression of SPHK1<br />

mRNA via SP1. [In vivo] At week 48 week of treatment, peretinoin<br />

down-regulated SPHK1 mRNA expression and prevented<br />

AHF diet-induced hepatocarcinogenesis. The SPHK1 knockout<br />

mice developed significantly less numbers of HCC induced by<br />

DEN than did wild-type mice. [Conclusion] Our data indicate<br />

that peretinoin prevents hepatocarcinogenesis at least partly<br />

through reduction of expression and activation of SPHK1.<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Masaya Funaki, Tetsuro Shimakami,<br />

Tsuguhito Ota, Masao Honda, Takayoshi Shirasaki<br />

496<br />

Therapeutic overexpression of miR-122 protects mice<br />

from chronic alcoholic liver injury through regulation of<br />

hypoxia-inducible factor-1α<br />

Abhishek Satishchandran 1 , Aditya Ambade 2 , Banishree Saha 2 ,<br />

Benedek Gyongyosi 2 , Patrick Lowe 2 , Nicita Mehta 2 , James V. Zatsiorsky<br />

2 , Arvin Iracheta-Vellve 2 , Jia Li 3,4 , Donna Catalano 2 , Karen<br />

Kodys 2 , Li Zhong 3,4 , Jun Xie 3,5 , Shashi Bala 2 , Guangping Gao 3,5 ,<br />

Gyongyi Szabo 2 ; 1 University of Massachusetts Medical School,<br />

Worcester, MA; 2 Medicine, UMass Medical School, Worcester,<br />

MA; 3 Gene Therapy Center, UMass Medical School, Worcester,<br />

MA; 4 Pediatrics, UMass Medical School, Worcester, MA; 5 Microbiology<br />

and Physiology Systems, UMass Medical School, Worcester,<br />

MA<br />

Introduction: miR-122 regulates essential pathways in alcoholic<br />

liver disease (ALD). We have observed that chronic alcohol<br />

administration reduces miR-122 levels in murine hepatocytes.<br />

Given its essential role in hepatic homeostasis, we hypothesized<br />

that loss of miR-122 contributes to ALD and can be a therapeutic<br />

target. In this study, our goals were to asses: first, the<br />

role of miR-122 in the pathogenesis of ALD and the therapeutic<br />

potential of miR-122 restoration; second, the role of Hypoxia<br />

Inducible Factor 1-α (HIF-1α), a putative miR-122 target; and<br />

third, the mechanisms by which alcohol downregulates miR-<br />

122. Methods: Wild-type 8-week-old, C57Bl/6 or hepatocyte<br />

specific HIF-1α knockout (hepHIFKO) mice were injected intravenously<br />

with hepatocyte-tropic AAV (AAV8) containing antimiR-122<br />

TuD (TuD), or scrambled (Scr) vector. After 14 days,<br />

mice were started on a Lieber-DeCarli (LDC) alcohol diet (Et) or<br />

calorie-matched control diet (PF) for 4 weeks. Some WT mice<br />

were treated with AAV8 pri-miR-122 (OX) on day 7 of the<br />

LDC diet. Results: Anti-miR-122 TuD treatment alone resulted in<br />

significant increases in liver injury (ALT), steatosis, inflammation<br />

(TNFα, IL1-β, MCP-1), and fibrosis (Acta2, pro-col1α) in PF<br />

mice compared to Scr-treated controls. The co-administration of<br />

miR-122 TuD and alcohol resulted in a synergistic effect, further<br />

increasing liver injury, inflammation and fibrosis. Restoration of<br />

miR-122 in hepatocytes of alcohol-fed mice by treatment with<br />

scAAV8 OX resulted in a significant improvement in serum ALT,<br />

inflammation and fibrosis suggesting a protective role for miR-<br />

122 in ALD. The hepatic expression and DNA-binding activity<br />

of HIF-1α, a miR-122 target, was increased in TuD+PF mice<br />

equivalent to that of alcohol feeding alone. HIF1-α activity was<br />

the highest in TuD+Et mice compared to all other groups. Hep-<br />

HIF deficiency protected mice from liver damage, inflammation<br />

and fibrosis induced either by alcohol or miR-122 knockdown.<br />

We discovered that alcohol specifically inhibits expression<br />

of the precursor miR-122 transcript (pri-miR-122) at the transcriptional<br />

level. Furthermore, grainyhead-like 2 (GRHL2), an<br />

inhibitor of miR-122 transcription, was significantly increased<br />

in hepatocytes of alcohol fed mice compared to controls. Conclusions:<br />

Our findings dissect a novel mechanistic regulatory<br />

axis of miR-122 and demonstrate the therapeutic potential of<br />

miR-122 restoration in ALD. miR-122 inhibition alone mimics<br />

and together with alcohol augments the steatosis, inflammation<br />

and early fibrosis seen in ALD through its downstream target,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 459A<br />

HIF-1α. Alcohol inhibits miR-122 via transcriptional repression<br />

mediated by GRHL2 upregulation.<br />

Disclosures:<br />

Guangping Gao - Stock Shareholder: VoyagerTherapeutics<br />

The following authors have nothing to disclose: Abhishek Satishchandran, Aditya<br />

Ambade, Banishree Saha, Benedek Gyongyosi, Patrick Lowe, Nicita Mehta,<br />

James V. Zatsiorsky, Arvin Iracheta-Vellve, Jia Li, Donna Catalano, Karen Kodys,<br />

Li Zhong, Jun Xie, Shashi Bala, Gyongyi Szabo<br />

497<br />

Niclosamide ethanolamine inhibits growth of patient-derived<br />

hepatocellular carcinoma xenografts<br />

Bin Chen 1 , Wei Wei 2 , Mei-Sze Chua 2 , Atul Butte 1 , Samuel K. So 2 ;<br />

1 Institute for Computational Health Sciences, UCSF, San Francisco,<br />

CA; 2 Asian Liver Center, Stanford University, Stanford, CA<br />

Background and Aim Hepatocellular carcinoma (HCC) is a fatal<br />

malignancy with a dismal prognosis. Sorafenib is the only drug<br />

approved for the treatment of advance HCC patients in the last<br />

decade; it is associated with only a modest response rate. The<br />

recent advances in genomics and transcriptomics have led to<br />

large volumes of molecular data for HCC, providing an unprecedented<br />

opportunity to translate these data into more effective<br />

therapeutics. To address the unmet clinical need in HCC, we<br />

leveraged the public large datasets on disease and drug gene<br />

expression signatures to identify drug candidates that may<br />

have yet unidentified anti-tumor efficacy in HCC. Materials and<br />

Methods We created a HCC gene expression signature by<br />

analyzing 250 patient samples from The Cancer Genome Atlas<br />

and validated it using 1624 patient samples from five independent<br />

cohorts. We further identified drug candidates that<br />

can reverse the HCC disease gene expression, from a library<br />

consisting of over 1000 FDA-approved drugs. We identified<br />

niclosamide and other antihelminthics as potential drug candidates<br />

for the treatment of HCC, and evaluated niclosamide and<br />

niclosamide ethanolamine (NEN, a water soluble niclosamide<br />

salt), for their in vitro and in vivo anti-tumor efficacy against<br />

HCC cells and patient-derived HCC xenografts (PDHX). Results<br />

In silico analysis indicated niclosamide could reverse the gene<br />

expression of HCC patients regardless of their etiology and<br />

tumor grade (FDR


460A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

cedure. A novel surgical approach, ALPPS (Associating Liver<br />

Partition and Portal Vein Ligation for Staged Hepatectomy),<br />

which combines PVL with parenchymal transition, may be used<br />

for unresectable disease due to its ability to markedly accelerate<br />

liver regeneration via unknown mechanisms. We hypothesize<br />

that transection is sufficient to activate plasma proteins<br />

necessary for accelerated liver regeneration, and that portal<br />

vein ligation (PVL) is necessary to stimulate liver growth. Aim<br />

Using a novel mouse model, we aim at studying the molecular<br />

mechanisms underlying the accelerated liver growth associated<br />

with ALPPS. Methods Liver tissue was collected at early time<br />

points after ALPPS (30min, 1h, 4h, 8h, 12h) and analysed by<br />

RNA deep sequencing. Plasma samples were collected 30min<br />

after ALPPS Step 1 for proteomics analysis by protein enrichment,<br />

albumin depletion, and mass spectrometry. Each protein<br />

fraction is confirmed for activity by inducing accelerated regeneration<br />

upon injection into mice treated with PVL only. Results<br />

Injection of plasma, derived from mice, subjected to ALPPS,<br />

into PVL alone mice is sufficient to accelerate liver regeneration<br />

as observed following ALPPS surgery. Likewise, a comparable<br />

response is induced in PVL alone mice by plasma from<br />

mice subjected to transection only. Plasma proteins appear<br />

to chiefly account for the accelerated liver growth. Transcript<br />

analysis shows that differential gene expression occurs 4 hours<br />

post ALPPS. The accelerated liver growth is characterized by<br />

a hyperplastic reaction reflected in an increased number of<br />

hepatocytes entering the cell cycle as early as 8 hours after<br />

surgery. Conclusions ALPPS is conducive to accelerated liver<br />

regeneration through proteins circulating in the plasma as early<br />

as 30 minutes after the transection step. ALPPS-specific gene<br />

expression changes are observed in liver at early times after<br />

surgery and are consistent with the induction of regenerative<br />

pathways. Identification of plasma proteins stimulating liver<br />

growth and associated regenerative pathways may have a<br />

high therapeutic value in expanding the surgical cure of liver<br />

disease previously deemed unresectable.<br />

Disclosures:<br />

The following authors have nothing to disclose: Magda Langiewicz, Andrea<br />

Schlegel, Bostjan Humar, Rolf Graf, Pierre-Alain Clavien<br />

500<br />

Identifying novel therapeutic targets in HCC through<br />

an integrated transcriptomics and pharmacogenomics<br />

approach<br />

Bin Chen 1 , Li Ma 2 , Mei-Sze Chua 2 , Atul Butte 1 , Samuel K. So 2 ;<br />

1 Institute for Computational Health Sciences, UCSF, San Francisco,<br />

CA; 2 Asian Liver Center, Stanford University, Stanford, CA<br />

Background Large-scale gene expression profiling of disease<br />

systems perturbed by chemical agents in preclinical models<br />

allows the understanding of underlying cellular responses<br />

induced by the chemicals. Such information may guide the discovery<br />

of new therapeutic drugs and targets in specific disease<br />

systems. Using hepatocellular carcinoma (HCC) as our disease<br />

model, we hypothesized that genes whose expressions are specifically<br />

reversed by anti-cancer drugs may be potentially novel<br />

therapeutic targets in HCC. Methods and Materials We compiled<br />

data on IC 50<br />

s of anti-cancer drugs from ChEMBL (a chemical<br />

database of bioactive molecules with drug-like properties),<br />

and drug-induced gene expression signatures from LINCS (a<br />

database containing perturbation profiles of 1,000 genes in<br />

HCC cell lines) and from CMap (a database containing whole<br />

genome perturbation profiles in all cancer cell types). Additionally,<br />

we computed gene expression signatures of HCC and<br />

non-tumor liver samples from TCGA. We compared the HCC<br />

gene expression signature with the drug signatures, and computed<br />

the ability of each drug to reverse the HCC signature.<br />

Each drug was associated with a Reversing Gene Expression<br />

Score (RGES), and the correlation between the IC 50<br />

s and RGES<br />

of 36 drugs was examined. We further identified genes that<br />

are specifically reversed by anti-cancer drugs with IC 50<br />

s


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 461A<br />

tion in HCC cell lines and enhanced the nuclear translocation<br />

of TFEB-GFP and p62 clearance in all HCC cell lines. Western<br />

blot analysis showed impaired expression of lysosomal<br />

hydrolases cathepsin D in all HCC cell lines except HepG2<br />

cells. Using real-time RT-PCR assay, we found that mRNA levels<br />

of number of lysosomal genes (TFEB, cathepsin D) were significantly<br />

reduced in all HCC cell lines compared to primary<br />

human hepatocytes. Furthermore, we investigated whether<br />

autophagy inhibitor (hydroxychloroquine) or inducer (Torin 1)<br />

could be used as a therapeutic strategy for HCC. Interestingly,<br />

long-term culture of HCC cells with autophagy inhibitor leads<br />

to complete inhibition of HCC growth by MTT and cell colony<br />

assay. Conclusion: Our results show that increased expression<br />

of lysosomal efflux protein (p62) in HCC cell lines is due to<br />

defects in mTOR-TFEB signaling. Our results also show that the<br />

autophagy inhibitor hydroxychloroquine can be a potential<br />

agent to inhibit HCC growth.<br />

Disclosures:<br />

The following authors have nothing to disclose: Rajesh Panigrahi, Partha K. Chandra,<br />

Pauline Ferraris, Fatma Aboulnasr, Srinivas Chava, Tong Wu, Srikanta Dash<br />

502<br />

Chimeric hepatitis E virus-like particles as a tool of oral<br />

vaccination against hepatitis A virus<br />

Eun Byul Lee 1 , Jung-Hee Kim 1 , Jung Eun Choi 1 , Seung Kew<br />

Yoon 1,2 ; 1 The Catholic University Liver Research Center & WHO<br />

Collaborating Center of Viral Hepatitis, Seoul, Korea (the Republic<br />

of); 2 College of Medicine, The Catholic University of Korea, Seoul,<br />

Korea (the Republic of)<br />

Background: Virus-like particles (VLPs) has been shown the possibility<br />

of new tools for oral vaccination because VLPs can be<br />

protected against harsh environment of the human digestive<br />

tract. Recent <strong>studies</strong> have demonstrated that hepatitis E virus<br />

(HEV)-LPs can induce both mucosal and systemic immunity after<br />

oral administration. Currently, commercial vaccine against<br />

hepatitis A virus (HAV) are in the intravenous form, but those<br />

for oral administration are yet to be developed. In this study,<br />

we investigated whether HEV-LPs containing HAV antigens<br />

(HEV-LP-HAVag) could be useful delivery tool for oral vaccination<br />

against HAV. Methods: To produce HEV-LPs from mammalian<br />

cells, Huh7 cells were infected with the recombinant<br />

baculovirus containing CMV promoter derived-Nt-ORF2 gene<br />

(Bac-Nt-ORF2). Nt-ORF2 expression in Huh7 cells was confirmed<br />

by western blot analysis, and HEV-LPs produced from<br />

Huh7 cells were purified by sucrose gradient centrifugation.<br />

The morphology of purified HEV-LPs was observed by electron<br />

microscopy (EM). Expression vectors of HAV antigens (VP1,<br />

VP3, VP1-P2A)-derived from HM175 strain were constructed<br />

by insertion of gene encoding each HAV antigen into pcDNA3<br />

vector in frame. To establish HEV-LPs packing system carrying<br />

HAV antigen expression vectors, HEV-LPs were disassembled<br />

using biochemical buffer containing DTT and low concentration<br />

of CaCl 2,<br />

and then HAV antigen expression vectors were<br />

packed into the re-assembled HEV-LPs by increasing concentration<br />

of CaCl 2.<br />

Delivery of HAV antigen expression vectors and<br />

their expression were confirmed in Huh7 cells infected HEV-<br />

LP-HAVag by western blot analysis. Results: Nt-ORF2 was successfully<br />

expressed in Huh7 cells infected with Bac-Nt-ORF2.<br />

Moreover, Nt-ORF2 expressed in Huh7 cells formed HEV-LPs<br />

of 25 nm in diameter through their self-assembly property. Furthermore,<br />

HEV-LP-HAVags were efficiently build up by in vitro<br />

disassembly/reassembly systems and, their morphology was<br />

well preserved compared to that of empty HEV-LPs. By infection<br />

with HEV-LP-HAVags, HAV antigens expression vectors were<br />

delivered to Huh7 cells, and subsequently 3 kinds of HAV antigens<br />

were efficiently expressed in the cells. Conclusion: We<br />

established an antigen-encapsulation system with HEV-LPs and<br />

successfully delivered the antigen expression vectors using the<br />

system. This might be useful tool for development of vaccine as<br />

well as vesicles for gene therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Eun Byul Lee, Jung-Hee Kim, Jung<br />

Eun Choi, Seung Kew Yoon<br />

503<br />

High-throughput sequencing to define the miRNA cargo<br />

of LSEC-derived, anti-viral exosomes<br />

Michael Kriss 1 , Michael Edwards 3 , Lucy Golden-Mason 1 , Colin<br />

Larson 4 , Katrina Diener 4 , Hugo R. Rosen 1,2 ; 1 Division of Gastroenterology<br />

& Hepatology, University of Colorado Anschutz Medical<br />

Campus, Aurora, CO; 2 Denver VA Medical Center, Denver,<br />

CO; 3 Division of Pulmonary Sciences and Critical Care Medicine,<br />

University of Colorado Anschutz Medical Campus, Aurora, CO;<br />

4 Genomics and Microarray Core, University of Colorado Anschutz<br />

Medical Campus, Aurora, CO<br />

Aim: To characterize the miRNA content of LSEC-derived exosomes<br />

to identify novel anti-viral mediators. Methods: Exosomes<br />

were isolated from supernatants of an immortalized LSEC cell<br />

line cultured for 48h ± rhIFNα (100ng/ml) or ± IFNλ1/2/3<br />

(100ng/ml each) using ExoQuick-TC (SBI). Total RNA was isolated<br />

using SeraMir RNA extraction kit (SBI). Small RNA library<br />

was generated using Illumina TruSeq Small RNA sample prep<br />

kit (1μg input RNA) and validated on Agilent Bioanalyzer.<br />

Sequencing was performed using Illmina HiSEQ 2500 Platform<br />

using V4 Chemistry Single Read 1x50. The sequenced reads<br />

were aligned with TopHat2 against both the human mirBASE<br />

and hg38 databases separately, and quantitated using Partek<br />

Genomics Suite v6.6. The average amount of reads produced<br />

from the miRNA and genome alignments were approximately<br />

3.3 and 9.2 million reads respectively per sample. Results:<br />

Exosomes derived from LSECs stimulated with either rhIFNα<br />

or IFNλ1/2/3 suppressed HCV replication in vitro and led<br />

to induction of ISGs in HCV-infected hepatocytes, consistent<br />

with prior reports. Electrophoresis of isolated RNA revealed<br />

absence of ribosomal RNA (18s and 28s) compared to RNA<br />

isolated from derivative cells. High throughput sequencing was<br />

performed with a base call accuracy of 99.96% (Phred Q score<br />

33.813±0.257). miRNA fold changes relative to mock were<br />

calculated based on read counts for all targets with greater<br />

than 100 total read counts. The top ten fold changes for each<br />

treatment group are included in table 1. miR-3168 was the<br />

only miRNA uniquely induced in the IFNλ-treatment group.<br />

miR-122 was present in all samples with modest induction in<br />

both IFNλ (1.31-fold) and IFNα (1.75-fold) treatment groups.<br />

Additional alignment with hg38 database revealed presence<br />

of miR-138 in only the IFNλ and IFNα treated groups. Conclusions:<br />

LSEC-derived exosomes differentially regulate miRNA in<br />

the setting of type I and type III interferons. This may contribute<br />

to the anti-viral effect of these exosomes and provide novel<br />

insights into therapeutic mechanisms moving forward.


462A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Michael Kriss, Michael Edwards,<br />

Lucy Golden-Mason, Colin Larson, Katrina Diener, Hugo R. Rosen<br />

504<br />

The Roles of Kras Isoforms in Liver Cancer<br />

Hyuk Moon 1,2 , Sook In Chung 1,2 , Kyung Joo Cho 1,2 , Hye Lim Ju 1 ,<br />

Dayoung Kim 1 , Do Young Kim 1,3 , Sang Hoon Ahn 1,3 , Kwang-<br />

Hyub Han 1,3 , Weonsang S. Ro 1,3 ; 1 Liver Center, Yonsei University<br />

College of Medicine, Seoul, Korea (the Republic of); 2 Brain Korea<br />

21 PLUS Project for Medical Science College of Medicine, Yonsei<br />

University, Seoul, Korea (the Republic of); 3 Department of Internal<br />

Medicine, Yonsei University College of Medicine, Seoul, Korea<br />

(the Republic of)<br />

Background/Aims: Activating mutation in Kras is frequently<br />

observed in human carcinogenesis. Alternative splicing of Kras<br />

transcripts gives rise to two isoforms, Kras4A and Kras4B. In<br />

this study, we compared the hepatocarcinogenic potentials of<br />

the KRAS splice variants carrying the same activating mutation.<br />

Methods: Transgenic liver cancer mouse models were<br />

developed using a hydrodynamic injection method and the<br />

Sleeping Beauty Transposon System. Transposons encoding<br />

Kras4AG12V or Kras4BG12V were mixed with transposons<br />

expressing short hairpin RNA downregulating p53 (shp53) and<br />

plasmids encoding the Sleeping Beauty transposases and then<br />

hydrodynamically delivered to the liver of 6-week-old C57BL/6<br />

mice. Livers were harvested and analyzed at 4 weeks post-hydrodynamic<br />

injection. Results: Co-expression of KRas4BG12V<br />

and shp53 resulted in massive abdominal enlargement within<br />

4 weeks after injection. Numerous nodular lesions emerged<br />

from the liver parenchyma and occupied most of the liver<br />

surface. In contrast, mice transfected with KRas4AG12V and<br />

shp53 had much smaller and fewer nodules in the liver. The<br />

ratio of liver weight/body weight in the Kras4BG12V group<br />

was significantly higher than that in KRas4AG12V group (p <<br />

0.001). Survival analysis also showed a significantly shorter<br />

life span for Kras4BG12V mice compared to Kras4AG12V<br />

mice (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 463A<br />

50% cell viability varies significantly among HCC cell lines.<br />

We have isolated HCC cell lines that were totally resistant<br />

to sorafenib and continuously grow in growth media supplemented<br />

with sorafenib. The mechanism of sorafenib resistance<br />

relates to the impaired uptake due to the reduced expression of<br />

organic cation transporter-1 and 3 (OCT1& OCT3). The uptake<br />

of doxorubicin is not altered in HCC cell lines. We found that<br />

OCT1 mRNA and protein expression is reduced in all HCC<br />

cell lines compared to primary human hepatocytes. Stable<br />

expression of full-length OCT1 in the resistant cell line induced<br />

cellular cytotoxicity to sorafenib and overcome the resistance.<br />

Conversely, we showed that inhibiting OCT1 expression by<br />

small molecule inhibitor (quinine hydrochloride) in a sensitive<br />

cell line reduced sorafenib uptake and induces resistance.<br />

The expression of OCT1 and OCT3 is significantly reduced<br />

in human hepatocellular carcinoma tissues as compared to the<br />

surrounding non-tumorous hepatocytes in the cirrhotic nodules.<br />

Conclusions: Our results indicate that there are intrinsic differences<br />

among HCC cell clones that affect sorafenib sensitivity.<br />

Expression of OCT1 and OCT3 down regulation in human<br />

samples is associated with sorafenib sensitivity. We propose<br />

that a detailed understanding of the sorafenib sensitivity and<br />

resistance mechanisms should allow for novel treatment options<br />

to improve sorafenib chemotherapy response in patients with<br />

liver cancer.<br />

Disclosures:<br />

The following authors have nothing to disclose: Srinivas Chava, Partha K. Chandra,<br />

Rajesh Panigrahi, Pauline Ferraris, Fatma Aboulnasr, James Liu, Jose J.<br />

Marin, Swan N. Thung, Tong Wu, Srikanta Dash<br />

507<br />

Stabilin-mediated endocytosis of antisense oligonucleotides<br />

for modulation of liver gene expression<br />

Colton M. Miller 1 , Aaron J. Donner 2 , Emma E. Blank 1 , Punit Seth 2 ,<br />

Edward N. Harris 1 ; 1 Biochemistry, University of Nebraska, Lincoln,<br />

NE; 2 ISIS Pharmaceuticals, Inc., Carlsbad, CA<br />

Introduction: Antisense oligonucleotides (ASOs) are short<br />

(5000-7400 Da) chemically-modified oligonucleotides that<br />

bind to cellular RNAs and modulate their intermediary metabolism<br />

to produce a pharmacological effect. DNA-based ASOs<br />

down-regulate gene-expression by promoting RNase H mediated<br />

degradation of the targeted mRNA. There are currently<br />

over 40 ASOs in clinical development targeting genes that are<br />

involved with metabolic diseases, cancer and inherited genetic<br />

diseases. An important difference between natural DNA and<br />

the ASO is that the phosphodiester (PO) backbone of DNA has<br />

been partially or fully substituted with a phosphorothioate (PS)<br />

backbone. The replacement of the negatively charged oxygen<br />

for the negatively charged sulfur atom covalently bound to<br />

phosphorus renders the short polymer both high stability in<br />

biological fluids and a propensity to adhere to cell surfaces.<br />

Some ASOs are further modified by 2’-O-methoxyethyl (MOE)<br />

or 2’-O-Methyl RNA nucleotides to further enhance metabolic<br />

stability and RNA-binding affinity. Injected ASOs accumulate<br />

in the liver due to the scavenging function of the organ. Previous<br />

work has shown that the liver sinusoidal endothelial cells<br />

(SECs) and Kupffer cells (KC) internalize a much greater proportion<br />

of ASO than hepatocytes. We have discovered that the<br />

non-DNA binding Stabilin-2/HARE receptor specifically binds<br />

to and internalizes several classes of ASO as this scavenger<br />

receptor is enriched in these liver cells. Methods: In our system,<br />

we used stable cell lines expressing recombinant human Stabilin-1<br />

and -2 to assess internalization rates and wildtype CD-1<br />

mice to determine 125 I-ASOs in organ distribution and uptake<br />

in hepatic cells. Results: Cell lines expressing either Stabilin-1<br />

or Stabilin-2 or the parental (non-Stabilin) empty vector (EV) cell<br />

line were incubated with 125 I-ASO that was fully modified with<br />

PS at 0°C (binding) and 37°C (endocytosis) and determined<br />

that the amount of ASO bound did not differ, though the Stabilin-2<br />

cells internalized 3-fold more than the Stabilin-1 cells and<br />

7-fold more than the EV cells. Twenty percent of the retro-orbitally<br />

injected ASO accumulated in liver with an amount of 48%,<br />

30%, and 22% in SECs, KCs and hepatocytes, respectively.<br />

Despite progress in the use of therapeutic oligonucleotides, the<br />

pathways by which negatively charged ASOs enter cells and<br />

modulate gene expression remain poorly understood. Conclusion:<br />

In this context, our work provides fundamental insights<br />

into cell-surface proteins which mediate ASO uptake into cells<br />

and provide opportunities to leverage these findings to further<br />

enhance the therapeutic properties of oligonucleotide drugs.<br />

Disclosures:<br />

Aaron J. Donner - Employment: Isis Pharmaceuticals; Stock Shareholder: Isis<br />

Pharmaceuticals<br />

Punit Seth - Employment: Isis Pharmaceuticals<br />

The following authors have nothing to disclose: Colton M. Miller, Emma E. Blank,<br />

Edward N. Harris<br />

508<br />

Shifting gut microbiota and bile acid profiles in retinoic<br />

acid-primed mice that exhibit accelerated liver regeneration<br />

Hui-Xin Liu, Ying Hu, Lili Sheng, Yu-Jui Yvonne Wan; Medical<br />

Pathology and Laboratory Medicine, UC DAVIS, Sacramento, CA<br />

Abstract Similar to bile acid (BA), all-trans retinoic acid (RA)<br />

facilitates 2/3 partial hepatectomy (PHx)-induced liver regeneration.<br />

RA can activate BA receptor farnesoid x receptor (FXR) to<br />

regulate BA signaling because retinoid x receptor α (RXRα) is<br />

a permissive partner of FXR. Our published data also showed<br />

that RA regulates hepatic lipid homeostasis by shifting hepatic<br />

BA profile. Since the gut microbiota plays a pivotal role in regulating<br />

BA homeostasis and BA composition, we examined the<br />

effect of RA in altering gut microbial and BA composition and<br />

their relationship with increased hepatocyte proliferation in<br />

response to liver resection. C57BL/6 mice were primed with a<br />

single dose of RA (25 μg/g) followed by PHx. RA-primed mice<br />

exhibited accelerated hepatocyte proliferation as indicated by<br />

higher number of Ki67-positive cells and earlier induction of<br />

cell cycle genes compared to untreated mice. Firmicutes and<br />

Bacteroidetes phyla, which affect the efficiency of host energy<br />

and are linked with adiposity in both mice and human, dominated<br />

the gut microbial community (>85%) in both control and<br />

RA-primed mice after PHx. RA reduced the ratio of Firmicutes<br />

to Bacteroidetes, which is associated with a lean phenotype.<br />

Consistently, RA-primed mice lacked transient lipid accumulation,<br />

an energy source for proliferating hepatocytes, normally<br />

found in regenerating mouse livers. This finding suggested that<br />

RA-primed mice might have increased lipid metabolism. In<br />

addition, RA altered BA signaling in the regenerating livers.<br />

RA activated the FXR-SHP-FGF15 pathway, reduced CYP7A1<br />

and CYP8B1 expression to inhibit BA synthesis, and increased<br />

BA transport through Oatp2, Asbt, and Ibabp induction in the<br />

ileum. These data suggested enhanced enterohepatic recirculation<br />

of BAs from RA-priming. RA treatment also shifted the<br />

BA profile by increasing the hydrophilic/hydrophobic ratio.<br />

The ratio of taurine-conjugated cholic acid (CA) to free CA<br />

and taurine-conjugated-β-murine CA to free β-murine CA were<br />

elevated in RA-treated regenerating livers suggesting increased<br />

solubility and lipid emulsification capability. Accordingly, fibroblast<br />

growth factor 21 (FGF21) and Sirtuin1 (SIRT1), master<br />

metabolic regulators, and their downstream targets AMPK and


464A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ERK1/2 were more robustly activated in RA-primed regenerating<br />

livers than controls. Conclusion: Priming mice with RA<br />

resulted in a lean microbiota composition and hydrophilic bile<br />

acid profile, which are associated with facilitated metabolism<br />

and enhanced cell proliferation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hui-Xin Liu, Ying Hu, Lili Sheng,<br />

Yu-Jui Yvonne Wan<br />

509<br />

Pre-operative plasma glypican-3 levels detected by a<br />

novel ELISA system predict the risk of post-operative<br />

recurrence in patients with stage I hepatocellular carcinoma<br />

Kazuya Ofuji 1,2 , Keigo Saito 1 , Yasunari Nakamoto 2 , Tetsuya<br />

Nakatsura 1 ; 1 Division of Cancer Immunotherapy, Exploratory<br />

Oncology Research and Clinical Trial Center, National Cancer<br />

Center, Kashiwa, Japan; 2 Second Department of Internal Medicine,<br />

University of Fukui, Fukui, Japan<br />

Background: Prognosis of post-operative hepatocellular carcinoma<br />

(HCC) patients remains unsatisfactory due to the high<br />

risk of recurrence. Glypican-3 (GPC3) is specifically overexpressed<br />

in HCC and has been recognized as a novel prognostic<br />

factor after curative resection of HCC. Recent <strong>studies</strong><br />

have shown the usefulness of GPC3 as a biomarker for the<br />

diagnosis of early HCC. However, the diagnostic value of<br />

GPC3 as a predictive marker for recurrence of postoperative<br />

early-stage HCC remains unclear. Here, this study was undertaken<br />

to evaluate the detection of plasma GPC3 using a novel<br />

sandwich enzyme-linked immunosorbent assay (ELISA) system<br />

in early-stage HCC patients after surgical resection. Methods:<br />

Plasma samples were collected from 25 patients with stage<br />

I HCC who underwent for surgical resection between 2008<br />

and 2010. The ELISA system for detection of plasma GPC3<br />

levels was established, by use of a newly-generated anti-GPC3<br />

mouse monoclonal capture antibody and a biotinylated detection<br />

antibody and recognized with streptavidin-conjugated<br />

horseradish peroxidase. GPC3 expression of surgical specimens<br />

was analyzed by immunohistochemical staining. Results:<br />

HCC recurrence was observed in 14 cases after surgical resection<br />

(51.8%) during follow-up period (median, 738 days).<br />

In the immunohistochemical analysis, GPC3 positive rate of<br />

resected specimen was 53% (13/23). There was a tendency<br />

towards a shorter recurrence-free survival (RFS) in GPC3 positive<br />

cases than in GPC3 negative cases (p 0.233). In the ELISA<br />

system, the mean ± standard deviation (SD) of plasma GPC3<br />

concentrations was 110.12 ± 37.70 ng/ml in control subjects,<br />

and cut-off value was determined to be 132 ng/ml (mean +<br />

1.5 SD). In stage I HCC, the sensitivity and specificity were<br />

40.7% and 92%, respectively. GPC3-positive rate at the time<br />

of recurrence was 64.3%. In the recurrence group, pre-operative<br />

plasma GPC3 levels were significantly higher than in the<br />

non-recurrence group (median, 191.7 ng/ml vs 29.7 ng/ml,<br />

p = 0.029). There was a tendency of shorter RFS in the pre-operative<br />

plasma GPC3-positive patients group compared with<br />

the negative group (median RFS, 769 days vs. not reached,<br />

p = 0.147). However, there was no significant difference in<br />

overall survival. The recurrence rate of patients with post-operative<br />

plasma GPC3 positive was significantly higher than<br />

GPC3 negative patients (p = 0.012). Conclusions: We have<br />

established a novel ELISA system for detecting plasma GPC3<br />

levels and evaluated the utility of GPC3 as a diagnostic marker<br />

of HCC. This study suggests pre-operative plasma GPC3 levels<br />

may help in identification of stage I HCC patients at high risk<br />

of post-operative recurrence.<br />

Disclosures:<br />

Tetsuya Nakatsura - Consulting: Ono Pharmaceutical co.Ltd.; Grant/Research<br />

Support: Ono Pharmaceutical co.Ltd., MEDINET co.Ltd., Sysmex co. Ltd.<br />

The following authors have nothing to disclose: Kazuya Ofuji, Keigo Saito, Yasunari<br />

Nakamoto<br />

510<br />

Evaluation of Glyoxylate and Hydroxyproline Metabolic<br />

Pathways Through the Use of Dicer-Substrate Small<br />

Interfering RNA<br />

Nicole Avitahl-Curtis 1 , Benjamin Holmes 1 , Luciano Apponi 2 , Rohan<br />

Diwanji 1 , Marita S. Larsson Cohen 2 , Chaitali Dutta 1 , Natalie W.<br />

Pursell 1 , Dongyu Chen 2 , Xue Shui 2 , Purva Pandya 2 , Utsav H.<br />

Saxena 2 , Martin Koser 2 , Abrams Marc 2 , Weimin Wang 2 , Hank<br />

Dudek 2 , Chengjung Lai 1 , Bob D. Brown 2 ; 1 Translational Biology,<br />

Dicerna Pharmaceuticals, Inc., Cambridge, MA; 2 Dicerna Pharmaceuticals,<br />

Inc., Cambridge, MA<br />

Primary Hyperoxaluria (PH) results from genetic mutations<br />

which affect glyoxylate metabolism, causing overproduction<br />

of oxalate. Three types of Primary Hyperoxaluria have been<br />

described to date, and the genetic lesions underlying disease<br />

have been identified. PH1, PH2 and PH3 are caused by mutations<br />

in the genes, AGXT, GRHPR and HOGA1, respectively,<br />

which encode the metabolic enzymes alanine: glyoxylate<br />

aminotransferase (AGT), glyoxylate reductase (GRHPR), and<br />

4-hydroxy-2-oxoglutarate adolase (HOGA). The roles of AGT<br />

and GRHPR in glyoxylate metabolism have been explored utilizing<br />

mouse strains harboring a germline disruption of the<br />

Agxt or Grhpr gene. Agxt- and Grhpr-deficient mice develop<br />

hyperoxaluria, consistent with the proposed roles of AGT and<br />

GRHPR in glyoxylate metabolism. Dicer-substrate siRNAs (DsiR-<br />

NAs) are potent and specific RNAi triggers that inhibit the<br />

expression of disease-relevant targets at the mRNA level, and<br />

are currently in clinical development. In this work, we describe<br />

the use of DsiRNAs to silence the expression of Agxt and Grhpr<br />

in mice to generate PH1 and PH2 animal models that phenotypically<br />

mimic Agxt- and Grhpr-deficient mice. DsiRNA-induced<br />

PH disease models were further explored for their utility<br />

to identify effective therapeutics. This approach was also used<br />

to investigate the dysregulation of the glyoxylate and hydroxyproline<br />

metabolic pathways occurring in each type of primary<br />

hyperoxaluria. Finally, we describe the utility of siRNA as an<br />

efficient and powerful method for generating models of liver<br />

disease in rodents and, potentially, other species.<br />

Disclosures:<br />

Nicole Avitahl-Curtis - Employment: Dicerna Pharmaceuticals, Inc.<br />

Benjamin Holmes - Employment: Dicerna Pharmaceuticals; Stock Shareholder:<br />

Dicerna Pharmaceuticals<br />

Luciano Apponi - Employment: Dicerna Pharmaceuticals<br />

Rohan Diwanji - Employment: Dicerna Pharmceuticals; Stock Shareholder:<br />

Dicerna Pharmceuticals<br />

Chaitali Dutta - Employment: Dicerna Pharmaceuticals<br />

Utsav H. Saxena - Employment: Dicerna Pharmaceuticals<br />

Martin Koser - Employment: Dicerna Pharmaceuticals<br />

Abrams Marc - Employment: Dicerna Pharmaceuticals<br />

Hank Dudek - Employment: Dicerna<br />

Chengjung Lai - Employment: Dicerna Pharmaceuticals<br />

Bob D. Brown - Employment: Dicerna Pharmaceuticals<br />

The following authors have nothing to disclose: Marita S. Larsson Cohen, Natalie<br />

W. Pursell, Dongyu Chen, Xue Shui, Purva Pandya, Weimin Wang


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 465A<br />

511<br />

Cell-Type-Specific Expression Profile of Repopulating<br />

Hepatocytes by Translating Ribosome Affinity Purification<br />

(TRAP)<br />

Kirk J. Wangensteen 1,2 , Amber W. Wang 2 , Monica Teta-Bissett 2 ,<br />

Klaus H. Kaestner 2 ; 1 Gastroenterology, University of Pennsylvania,<br />

Philadelphia, PA; 2 Genetics, University of Pennsylvania, Philadelphia,<br />

PA<br />

Understanding the genetic regulation of liver regeneration<br />

could help in the development of new treatments for liver disease.<br />

Much of the research on liver regeneration has focused<br />

on the partial hepatectomy model in rodents; however, the lack<br />

of cell necrosis and immune response calls for the development<br />

of novel paradigms that recapitulate human liver diseases. Fah -<br />

/-<br />

mice, a model of hereditary tyrosinemia, are suitable for the<br />

study of liver repopulation after injury because repopulating<br />

hepatocytes can be genetically traced. We first injected Fah -<br />

/-<br />

mice with plasmids that co-express Fah and GFP, and found<br />

that we could isolate GFP-positive hepatocytes by fluorescence<br />

activated cell sorting, but met challenges in isolating pure populations<br />

of repopulating hepatocytes in the setting of injury. We<br />

then employed a different approach, the Translating Ribosome<br />

Affinity Purification (TRAP) system, to isolate mRNA specifically<br />

from repopulating hepatocytes. We injected Fah -/- mice with<br />

a plasmid that co-expresses Fah and a fusion of GFP and the<br />

ribosomal subunit L10a, and then induced liver injury for 1 or<br />

4 weeks. Liver tissue harvested from these mice was immediately<br />

lysed and polysomes were extracted. Anti-GFP antibodies<br />

were then used to affinity-purify the mRNAs bound to the GFP-<br />

L10a fusion protein expressed within repopulating hepatocytes<br />

(Figure). No RNA was obtained from wildtype control liver,<br />

as expected. We performed high throughput mRNA sequencing<br />

(RNA-seq) to identify differentially expressed genes during<br />

repopulation as compared to quiescent hepatocytes. Gene<br />

ontology analysis showed that the genes were enriched in metabolic<br />

pathways and in c-Myc-responsive genes. In conclusion,<br />

this represents the first analysis of the differential expression<br />

of repopulating hepatocytes in an injury model. Intriguingly,<br />

liver repopulation involves differential regulation of metabolic<br />

genes, likely to maintain metabolic homeostasis in the setting<br />

of widespread liver injury.<br />

Bioanalyzer tracings of affinity-purified RNA from mice treated<br />

with (Green and Red), or without (Blue) the TRAP vector, illustrating<br />

the specificity of the system.<br />

512<br />

In vivo distribution of exosome packaged extracellular<br />

miRNA-155 to the liver, hepatocytes and Kupffer cells<br />

and its functional effects<br />

Shashi Bala 1 , Timea Csak 1 , Fatemah Momen-Heravi 1 , Dora Lippai<br />

1 , Karen Kodys 1 , Donna Catalano 1 , Abhishek Satishchandran 1 ,<br />

Victor Ambros 2 , Gyongyi Szabo 1 ; 1 Medicine, UMass Medical<br />

School, Worcester, MA; 2 Molecular Medicine, UMass Medical<br />

School, Worcester, MA<br />

Purpose: Circulating miRNAs are gaining increasing interest<br />

as new biomarkers of diseases and means of intercellular communication.<br />

Previously, we showed induction of miR-155 in the<br />

plasma and found that miR-155 was associated with extracellular<br />

vesicles (EVs). However, little is known about the half-life<br />

and biodistribution of EVs associated miRNAs in vivo. Here, we<br />

established an exosome-based miR-155 mimic delivery system<br />

to study the biodistribution and half-life of EV associated miR-<br />

155 using a miR-155 knock out (KO) mouse model. Methods:<br />

Female wild type (WT) (C57/BL6J) or miR-155 KO mice were<br />

used. Exosomes were isolated from murine B cells using filtration<br />

method and characterized by electron microscopy, Nanosight<br />

and Western blot analyses. Electroporation was used<br />

to introduce miR-155 mimic into exosomes. Purified miR-155<br />

or control mimic loaded exosomes were administered intravenously<br />

to miR-155 KO mice. Mice were perfused to eliminate<br />

blood contamination. The Mann-Whitney test was employed<br />

for statistical analysis. Results: Our results indicate that administration<br />

of exosomes loaded with miR-155 mimic resulted in a<br />

rapid accumulation of miR-155 in the plasma of recipient miR-<br />

155 KO mice with subsequent distribution in the liver, adipose<br />

tissue, lung, muscle and kidney (highest to lowest, respectively).<br />

At the cellular level, miR-155 was detected in hepatocytes and<br />

liver mononuclear cells of recipient miR-155 KO mice in vivo,<br />

suggesting successful uptake of exosomal miR-155 by various<br />

cell populations of the liver. Functionally, exosomes-mediated<br />

restoration of miR-155 in Kupffer cells isolated from miR-155<br />

KO mice resulted in the induction of pro-inflammatory cytokines<br />

(MIP2 and MCP1) after LPS challenge in an in vitro co-culture<br />

system. To mimic the half-life and distribution of miRNA naturally<br />

present in the plasma, we administered (iv) WT plasma<br />

(donor) to miR-155 KO (recipient) mice. WT plasma was generated<br />

from CpG+LPS treated WT mice and was enriched in miR-<br />

155. Our results indicate detection of donor miR-155 in plasma<br />

of recipient miR-155 KO mice and distribution to the liver and<br />

adipose tissue. Induction of pro-inflammatory cytokines was<br />

found in recipient miR-155 KO mice suggesting some biological<br />

activity. In summary, our results demonstrate tissue biodistribution<br />

and biologic function of EV-associated miR-155.<br />

Conclusion: Exosome packaged miR-155 is readily distributed<br />

to the liver in a cell–specific manner resulting in KC activation.<br />

These observations will aid in developing new exosome-mediated<br />

miRNA delivery and investigating biological function of<br />

other extracellular miRNAs in vivo.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shashi Bala, Timea Csak,<br />

Fatemah Momen-Heravi, Dora Lippai, Karen Kodys, Donna Catalano, Abhishek<br />

Satishchandran, Victor Ambros, Gyongyi Szabo<br />

Disclosures:<br />

The following authors have nothing to disclose: Kirk J. Wangensteen, Amber W.<br />

Wang, Monica Teta-Bissett, Klaus H. Kaestner


466A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

513<br />

Integrated proteomics analyses identified a<br />

phosphoprotein relevant to apoptosis resistance in<br />

human hepatocellular carcinoma<br />

Hideaki Naoe 1 , Yohmei Hoshida 1 , Seiichiro Yokote 1 , Satomi<br />

Fujie 1 , Takehisa Watanabe 1 , Katsuya Nagaoka 1 , Taiji Yamazoe<br />

1 , Hiroko Setoyama 1 , Motohiko Tanaka 1 , Jiro Fujimoto 3 , Norie<br />

Araki 2 , Yutaka Sasaki 1 ; 1 gastroenterology and hepatology, Kumamoto<br />

University, Kumamoto, Japan; 2 Tumor Genetics and Biology,<br />

Kumamoto University, Kumamoto, Japan; 3 Surgery, Hyogo College<br />

of Medicine, Nishinomiya, Japan<br />

Aim: This study was aimed to identify the key molecules which<br />

account for apoptosis resistance of hepatocellular carcinoma<br />

(HCC) through integrated comprehensive analyses not only<br />

of gene and protein expression but also of post-translational<br />

modification which strictly regulate protein function and biology.<br />

Materials & Methods: Human HCC cells were stimulated<br />

with or without H 2<br />

O 2<br />

, an apoptosis-inducing stimulation. Protein<br />

and mRNA samples were extracted from these cells, and<br />

subjected to two-dimensional fluorescence differential gel electrophoresis<br />

(2-D DIGE) and cDNA microarray, respectively.<br />

After these analyses, we integrated obtained data to identify<br />

differentially expressed genes and proteins. Proteins that<br />

showed a significant change in phosphorylation after H 2<br />

O 2<br />

stimulation were further characterized by mass spectrometry.<br />

To block the phosphorylation of the target protein, we replaced<br />

a major phosphorylation site threonine by an alanine and stably<br />

expressed the wild-type protein and its TA mutant in HCC<br />

cells. MiRNA extracted in the same manner described above<br />

were subjected to miRNA array analysis. Obtained data were<br />

reprocessed for network analysis focused on the target protein.<br />

Finally, expression and phosphorylation of the protein was<br />

determined in surgically resected human HCCs and in adjacent<br />

liver tissues. Results: Broad range of around 30 proteins,<br />

including several classes of cytoskeletal proteins or molecular<br />

chaperons exhibited a significant phosphorylation change in<br />

response to apoptosis stimulation. Based on pathway analysis,<br />

we focused on one of the most prominently changed protein,<br />

nucleophosmin (NPM). Sensitivity to apoptosis stimulation varied<br />

among the cell lines, and decrease in NPM phosphorylation<br />

was more prominent in apoptosis-sensitive cell line than in<br />

apoptosis-refractory cell line. Knockdown of NPM using NPM<br />

specific siRNA in HCC cells enhanced cell death when treated<br />

with H 2<br />

O 2<br />

. Functional disruption of NPM phosphorylation in<br />

HCC cells reduced their apoptosis resistance. Network analysis<br />

indicated that miRNA X is associated with NPM phosphorylation<br />

via proteins such as PTEN and Cdk. Expression and<br />

phosphorylation of NPM were significantly enhanced in human<br />

HCC tissues compared to the adjacent liver tissues (n=23).<br />

Recurrence-free survival after surgical resection is significantly<br />

longer in HCC with lower expression of NPM phosphorylation<br />

group (n=14) than higher expression group (n=9). Conclusion:<br />

Our integrated proteomics strategy demonstrates that apoptosis<br />

resistance of human HCCs is, in part, accounted for by phosphorylation<br />

of NPM and downstream signaling, which may be<br />

a novel therapeutic target for human HCC.<br />

Disclosures:<br />

Yutaka Sasaki - Grant/Research Support: Chugai Pharmaceutical Co. JAPAN,<br />

MSD Co. JAPAN<br />

The following authors have nothing to disclose: Hideaki Naoe, Yohmei Hoshida,<br />

Seiichiro Yokote, Satomi Fujie, Takehisa Watanabe, Katsuya Nagaoka, Taiji<br />

Yamazoe, Hiroko Setoyama, Motohiko Tanaka, Jiro Fujimoto, Norie Araki<br />

514<br />

YAP Promotes Cyclin D1 Expression in Cholangiocytes<br />

and Hepatocytes<br />

Quy P. Nguyen, Ying Wan, Fanyin Meng, Haibo Bai; BaylorScott&White<br />

Hospital, Temple, TX<br />

Background: YAP is a critical regulator of bile duct specification,<br />

cholangiocyte and hepatocyte proliferation and survival.<br />

Elevated YAP activity stimulated bile duct development, cholangiocyte<br />

and hepatocyte proliferation and survival. In agreement,<br />

Yap deficiency results in bile duct paucity and reduced<br />

cholangiocyte/hepatocyte proliferation and liver survival.<br />

However, how YAP regulates these events is less understood.<br />

We aim to identify YAP’s downstream target in the liver. Methods:<br />

Immunohistochemistry of Cyclin D1 was performed in liver<br />

sections. Bile ducts and hepatocytes were isolated from H&E<br />

stained frozen liver sections with laser capture microscope.<br />

mRNA levels of Cyclin D1 were measured by real-time PCR.<br />

Results: In both developing livers and adult livers, Cyclin D1 is<br />

highly expressed in cholangiocytes. When knocking out Yap<br />

in the developing liver with Alb-Cre;Yap flox/flox mice, Cyclin<br />

D1-positive biliary cells on ductal plates are significantly less<br />

than WT controls. When increasing YAP activity through ablating<br />

upstream negative regulator NF2 with Alb-Cre;Nf2 flox/flox<br />

mice, the number of Cyclin D1-positive cells on ductal plates<br />

increases dramatically. In adult liver, when inducing YAP<br />

overexpression in hepatocytes of Yap transgenic mice, there<br />

is a compensatory loss of YAP expression in cholangiocytes.<br />

Hepatocyte YAP overexpression induces Cyclin D1 expression<br />

in hepatocytes. The loss of YAP expression in cholangiocytes<br />

results in corresponding loss of Cyclin D1 expression in cholangiocytes.<br />

mRNA analysis of bile ducts and hepatocytes isolated<br />

from Yap Tg mice with laser capture microscope revealed that<br />

YAP did not regulate Cyclin D1 mRNA transcription. Conclusions:<br />

Because Cyclin D1 is known to promote proliferation<br />

and survival, YAP may regulate hepatocyte and cholangiocyte<br />

proliferation and survival through promoting Cyclin D1 activity.<br />

Cyclin D1 inhibitor represent another treatment strategy for<br />

YAP overexpression liver cancer.<br />

Disclosures:<br />

The following authors have nothing to disclose: Quy P. Nguyen, Ying Wan,<br />

Fanyin Meng, Haibo Bai<br />

515<br />

Therapeutic Potential in Toxic Liver Injury of Exosomes<br />

Isolated from Healthy or Diseased Donors Depends on<br />

their Content with Capacity to Beneficially Regulate Cellular<br />

Gene Expression<br />

Yogeshwar Sharma 1 , Tatyana Tchaikovskaya 1 , Preeti Viswanathan<br />

2 , Charles E. Rogler 1,3 , David B. Rhee 4 , Pilib Ó Broin 4 , Wilber<br />

Quispe 4 , Alexander Y. Maslov 4 , Aaron Golden 4 , Sanjeev<br />

Gupta 1,5 ; 1 Medicine, Albert Einstein College of Medicine, Bronx,<br />

NY; 2 Division of Pediatric Gastroenterology, Children’s Hospital<br />

at Montefiore Medical Center, Bronx, NY; 3 Department of Immunology<br />

& Microbiology, Albert Einstein College of Medicine and<br />

Montefiore Medical Center, Bronx, NY; 4 Department of Computational<br />

Sciences, Albert Einstein College of Medicine and Montefiore<br />

Medical Center, Bronx, NY; 5 Department of Pathology,<br />

Albert Einstein College of Medicine and Montefiore Medical Center,<br />

Bronx, NY<br />

The biological role and therapeutic potential of exosomes<br />

recently gathered much interest. Exosomes contain a variety<br />

of materials, including cellular proteins, mRNA and microRNA<br />

species, lipids, small biomolecules, etc. When isolated exosomes<br />

are injected into subjects, these are efficiently targeted to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 467A<br />

liver with incorporation in hepatocytes. To elucidate the ability<br />

of exosomes to alter pathophysiological events and processes<br />

in recipients and to determine donor-dependent potential of<br />

exosomes, we studied healthy C57BL/6 mice and mice given<br />

single LD50 dose of acetaminophen (APAP) with liver injury<br />

characterized by abnormal liver tests and hepatic necrosis.<br />

Exosomes were isolated from mouse sera and medium from<br />

hepatocytes in cell culture. Also, exosomes were isolated from<br />

human sera. Administration of exosomes with reporters showed<br />

these were efficiently incorporated in cultured hepatocytes and<br />

intact mouse liver. In mouse or human hepatocyte cell culture<br />

assays with MTT for APAP cytotoxicity, exosomes from healthy<br />

mouse or human sera and culture medium from healthy hepatocytes,<br />

but not APAP-treated cells or mice, were cytoprotective.<br />

Administration of healthy exosomes after APAP in C57BL/6<br />

mice decreased mortality significantly, p


468A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of HSC. We are now carrying out the molecular analyses of<br />

CYGB promoter activation stimulated by phosphorylated JNK.<br />

Disclosures:<br />

Massimo Pinzani - Advisory Committees or Review Panels: Intercept Pharmaceutical,<br />

Silence Therapeutic, Abbot; Consulting: UCB; Speaking and Teaching:<br />

Gilead, BMS<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

The following authors have nothing to disclose: Misako Sato, Matsubara, Tsutomu<br />

Matsubara, Atsuko Daikoku, Krista Rombouts, Kazuo Ikeda<br />

518<br />

Impaired skeletal muscle mitochondrial respiration<br />

during hyperammonemia of cirrhosis contributes to sarcopenia<br />

Julie H. Rennison 2 , Avinash Kumar 2 , Gangarao Davuluri 2 , Allawy<br />

Allawy 2,4 , Rafaella Nascimento e Silva 2 , Dharmvir Singh 2 , David<br />

R. Van Wagoner 5 , Hoppel Charles 3 , Srinivasan Dasarathy 1 ;<br />

1 Department Of Gastroenterology and Hepatology, Cleveland<br />

Clinic, Cleveland, OH; 2 Pathobiology, Cleveland Clinic, Cleveland,<br />

OH; 3 Pharmacology and Medicine, Case Western Reserve<br />

University, Cleveland, OH; 4 Department Of Internal Medicne,<br />

Cleveland Clinic, Cleveland, OH; 5 Molecular Cardiology, Cleveland<br />

Clinic, Cleveland, OH<br />

Hyperammonemia is a consistent abnormality in cirrhosis.<br />

Skeletal muscle metabolic disposal of ammonia in cirrhosis<br />

has been shown to contribute to sarcopenia or loss of skeletal<br />

muscle mass. Ammonia disposal occurs in muscle mitochondria<br />

and increased fragmentation and impaired ATP content occur<br />

in the muscle during hyperammonemia of cirrhosis. We tested<br />

the hypothesis that mitochondrial oxidation is decreased during<br />

hyperammonemia and identified the specific sites on the electron<br />

transport chain complexes (ETCC) that contribute to the<br />

impaired mitochondrial respiration. We evaluated the impact<br />

of hyperammonemia (AmAc) on mitochondrial respiration<br />

in differentiated C2C12 murine myotubes by high-resolution<br />

respirometry in the basal state in intact cells and in response<br />

to specific substrates for the ETCC (pyruvate for complex I,<br />

succinate for complex II, TMPD for complex III) and inhibitors<br />

of each of the ETCC (rotenone for complex I, oligomycin for<br />

ATP synthetase, FCCP for uncoupling oxidative phosphorylation,<br />

antimycin A for Complex III). Control and AmAc treated<br />

cells were were assessed in separate chambers simultaneously.<br />

Hyperammonemia decreased basal respiration in intact cells<br />

by ~35% (Control, 67.0±14.1; 24hr AmAc 42.3±7.9 pmol<br />

O 2<br />

s -1 10 -6 ) (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 469A<br />

way, thereby also significantly decreasing the expression of the<br />

FXR-inducible target genes OSTα/β and OATP1B3. We show<br />

a new miRNA-dependent mechanism of FXR regulation, which<br />

could affect the expression of FXR target genes as shown here<br />

for colon and liver carcinoma derived cell lines and tissue.<br />

Disclosures:<br />

The following authors have nothing to disclose: Regina Krattinger, Jessica<br />

Mwinyi, Gerd A. Kullak-Ublick, Helgi B. Schiöth, Adrian Boström<br />

520<br />

Chenodeoxycholic acid induced changes in the expression<br />

of microRNAs and genes involved in lipid, bile acid<br />

and drug metabolism in human hepatocytes<br />

Gerd A. Kullak-Ublick 2 , Regina Krattinger 2 , Serene M. Lee 3 ,<br />

Wolfgang E. Thasler 3 , Jessica Mwinyi 1,2 ; 1 Division of Functional<br />

Pharmacology, Department of Neuroscience, Uppsala University,<br />

Uppsala, Sweden; 2 Department of Clinical Pharmacology and Toxicology,<br />

University Hospital Zurich, Zurich, Switzerland; 3 Department<br />

of General, Visceral, Transplantation, Vascular and Thoracic<br />

Surgery, Grosshadern Hospital, Ludwig Maximilians University,<br />

Munich, Munich, Germany<br />

Background: Bile acids (BAs) are important for the absorption<br />

of dietary lipids, lipid soluble vitamins and xenobiotics. As<br />

ligands for different transcription factors, e.g. the farnesoid X<br />

receptor, they can act as gastrointestinal signaling hormones<br />

that influence lipid, glucose, and energy homeostasis. Both<br />

hepatic disease states as well as drugs can lead to cholestasis,<br />

which is characterized by elevated intrahepatic BA levels.<br />

Here, we studied to which extent BAs modulate the mRNAome<br />

and miRNAome in human hepatocytes. Methods: Five batches<br />

of primary human hepatocytes were treated with 50 μmol/L<br />

chenodeoxycholic acid (CDCA) for 24 or 48 hours. Total RNA<br />

was extracted, size fractionated and subjected to Next Generation<br />

Sequencing to generate comprehensive mRNA and<br />

miRNA profiles. Results: Expression of 738 mRNAs and 52<br />

miRNAs was significantly decreased, whereas 1566 mRNAs<br />

and 29 miRNAs were significantly increased in CDCA treated<br />

hepatocytes. Several mRNAs from within gene clusters that<br />

control both BA and lipid homeostasis (FGF(R), APO and FABP<br />

family members, HMGCS2) and drug metabolism (CYP, UGT<br />

and SULT family members) showed a modulated expression in<br />

response to CDCA. Furthermore, CDCA was able to modulate<br />

the expression of several microRNAs such as e.g. miR-149,<br />

-505, -885 and -452. Of note, miR-34a showed an inverse<br />

expression pattern in relation to a distinct cluster of mRNAs<br />

consisting of ABCG5 and ABCG8, SLCO1B1, SLC22A7<br />

and FGF19. Conclusions: CDCA alters intracellular levels of<br />

regulatory miRNAs such as miR-34a, thereby influencing the<br />

expression of genes that are regulated by these miRNAs. Gene<br />

clusters thus controlled by CDCA are involved in bile acid<br />

homeostasis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gerd A. Kullak-Ublick, Regina<br />

Krattinger, Serene M. Lee, Wolfgang E. Thasler, Jessica Mwinyi<br />

521<br />

A Regional Multidisciplinary Liver Tumor Board<br />

improves Access to Hepatocellular Carcinoma Treatment<br />

for Patients Geographically Distant from Tertiary Medical<br />

Center<br />

E. M. Egert 1 , Rena D. Johnson 2 , Micah M. Watts 3 , Jordan Booty 3 ,<br />

Nevena Damjanov 3 , Rosemary Mohr 3 , Jenifer Griffiths 4 , Major<br />

Lee 3 , Haroon A. Shaikh 4 , Richard C. Hawley 4 , Martha S. Ghosh 5 ,<br />

Nicole S. Mungiole 3 , David E. Kaplan 3 ; 1 Coatesville VA Medical<br />

Center, Cochranville, PA; 2 Wilmington VA Medical Center, Wilmington,<br />

DE; 3 Philadelphia VA Medical Center, Philadelphia, PA;<br />

4 Lebanon VA Medical Center, Lebanon, PA; 5 Wilkes-Barre VA<br />

Medical Center, Wilkes-Barre, PA<br />

Introduction: Optimal care for hepatocellular carcinoma (HCC)<br />

requires multiple disciplines with expertise not always available<br />

at non-tertiary medical centers. Distance from tertiary centers<br />

has previously been shown to negatively impact access to liver-related<br />

services such as transplantation in the VA (Goldberg<br />

DS, et al. JAMA 2014). The goal of the VISN4 HCC Initiative<br />

was to improve access to HCC care for patients diagnosed<br />

at distant hospitals. Methods: A regional liver cancer team<br />

was established in eastern Pennsylvania and Delaware with<br />

clinicians, radiologists and coordinators from 4 spokes centers<br />

(Coatesville, Lebanon, Wilkes-Barre and Wilmington) and one<br />

tertiary hub (Philadelphia). Telehealth infrastructure was created<br />

to foster face-to-face participation of referring providers<br />

with the central multidisciplinary liver tumor board (LTB) for<br />

weekly meetings starting July 2014. Telehealth infrastructure<br />

was also used clinically to facilitate patient-to-specialist consultation<br />

from remote sites. To assess the impact of this program,<br />

we collected the date of first radiologic or serologic (abnormal<br />

AFP) abnormality, the first cross-sectional contrast enhanced<br />

imaging (CT/MRI), the first consultation order to the tertiary<br />

center, the first cancer plan, and the first cancer-related intervention<br />

for patients diagnosed with HCC in all 5 centers in<br />

the 12 months preceding and 10 months following the first<br />

meeting of the LTB. Date differences for the pre- and post-initiation<br />

time frames were compared using Wilcoxon Rank Sum<br />

tests stratified by Spoke/Hub status. Results: 124 patients<br />

with HCC were diagnosed: 61 pre-LTB (spokes 37, hub 28)<br />

and 63 post-LTB (spokes 26, hub 37). Median age was 64;<br />

55% were white, 44% were black, and 1% other. The number<br />

of cases diagnosed by liver biopsy as opposed to imaging<br />

declined from 15/61 (24%) to 4/63 (6%) (p=0.006). Time<br />

between the first abnormality and evaluation by a specialist/<br />

LTB declined from 44d (18-75) to 21d (11-35) (p=0.0018)<br />

primarily for cases at spokes (55d to 23d, p=0.0004). For<br />

patients diagnosed at a spoke facility, time from first abnormality<br />

to first treatment declined from 79d (62-124) to 54d (34-83)<br />

(p=0.025). The number of resections performed increased from<br />

0 pre-LTB to 6 post-LTB (p=0.015).Conclusion: The creation of<br />

a LCTB improved access to HCC care, reduced unnecessary<br />

biopsies, shortened time from identification to treatment primarily<br />

due to improvement from spokes sites. These data mimic<br />

HCC care access improvements achieved with a similar program<br />

in VISN3 suggesting that this model warrants exploration<br />

for further implementation with in the VA system and possibly<br />

in private settings.<br />

Disclosures:<br />

Nevena Damjanov - Grant/Research Support: ArQule; Speaking and Teaching:<br />

SIRTex, Bayer, Amgen<br />

David E. Kaplan - Grant/Research Support: Bayer Healthcare Inc, Gilead, Inovio<br />

Pharmaceuticals<br />

The following authors have nothing to disclose: E. M. Egert, Rena D. Johnson,<br />

Micah M. Watts, Jordan Booty, Rosemary Mohr, Jenifer Griffiths, Major Lee,<br />

Haroon A. Shaikh, Richard C. Hawley, Martha S. Ghosh, Nicole S. Mungiole


470A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

522<br />

Diminished Health-Related Quality of Life is an Independent<br />

Predictor of Hospitalization in Patients with<br />

Cirrhosis<br />

Sujit V. Janardhan 1 , Jason Morris 1 , David G. Beiser 2 , Nancy<br />

Reau 1 ; 1 Medicine, Section of Gastroenterology, Hepatology and<br />

Nutrition, Center for Liver Diseases, University of Chicago Medicine,<br />

Chicago, IL; 2 Medicine, Section of Emergency Medicine,<br />

University of Chicago Medicine, Chicago, IL<br />

INTRODUCTION: Decreased health-related quality of life (QOL)<br />

has been shown to be predictive of adverse clinical outcomes<br />

in many disease states. Patients with cirrhosis have diminished<br />

QOL that worsens with disease progression. However, it is not<br />

known whether diminished QOL is an independent predictor<br />

of clinical outcomes, such as hospitalization. HYPOTHESIS: We<br />

hypothesized that decreased QOL is an independent predictor<br />

of hospitalization in patients with cirrhosis. METHODS: We performed<br />

a retrospective cohort analysis of patients with cirrhosis<br />

who were enrolled in the chronic liver disease (CLD) patient<br />

database at the University of Chicago between 11/2012 and<br />

11/2014. QOL scores (as measured by patient responses to<br />

the Chronic Liver Disease Questionnaire) and clinical data (e.g.<br />

compensated vs. decompensated cirrhosis) were collected at<br />

the time of enrollment into the database. Patients were assessed<br />

for the number and length of hospitalizations occurring from<br />

the time of enrollment until the end of available follow up. QOL<br />

scores were analyzed for both overall score and QOL quartile<br />

(first quartile = highest 25% of HR-QOL scores). Univariate<br />

(ANOVA) and multivariate (linear regression) analysis was<br />

performed to find independent predictors of hospitalization.<br />

RESULTS: 100 patients with cirrhosis were identified from a<br />

cohort of 579 patients enrolled in the CLD database including<br />

67 patients with compensated cirrhosis (Childs-Turcotte-Pugh<br />

(CTP) class A) and 33 patients with decompensated cirrhosis<br />

(CTP B = 31, 2 CTP C = 2). The average length of follow-up<br />

was 19.61 +/- 8.82 months. 40 patients had at least one hospitalization<br />

during the follow-up period, with an average length<br />

of stay of 8.95 +/- 8.66 days (range 1-37 days). In univariate<br />

analysis, the average number of days spent hospitalized varied<br />

significantly between patients with compensated cirrhosis vs.<br />

decompensated cirrhosis (2.07 +/- 3.93 vs. 6.64 +/- 10.26<br />

days, respectively; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 471A<br />

524<br />

Did improvement of access to healthcare change the<br />

clinical features and long-term outcomes of hepatocellular<br />

carcinoma in a suburban area in Japan?<br />

Teru Kumagi 2,1 , Takahide Uehara 2 , Masaki Ohmoto 2 , Atsushi<br />

Hiraoka 2 , Miyake Teruki 2,1 , Kazuhiro Tange 2 , Akiko Toshimori 2 ,<br />

Norio Horiike 2 , Morikazu Onji 2 ; 1 Gastroenterology and Metabology,<br />

Ehime University Graduate School of Medicine, To-on, Japan;<br />

2 Saiseikai Imabari Hospital, Imabari, Japan<br />

Background and aim: Access to healthcare is an important<br />

factor that may affect the management of any disease. The<br />

aim of this study is to identify the change of clinical features at<br />

diagnosis and long-term outcomes of HCC, focusing on access<br />

to healthcare in a suburban area in Japan. Methods: Chart<br />

review was conducted for patients with HCC diagnosed at<br />

a single center. Collected data (1983–2011) was analyzed<br />

including the following data: demographics (age at diagnosis,<br />

sex, residential postcode), Child-Pugh (CP) score, clinical stage<br />

of HCC (TNM classification) at diagnosis, initial treatments<br />

(resection, ablation, TACE, palliative care) and outcomes.<br />

Patients were divided into 2 groups according to their residential<br />

postcodes for mainland and island, in which they needed<br />

to take ships to visit our hospital until the bridges opened in<br />

Apr 1999. Thus the era was divided into 2 groups based<br />

on the changes in the access pattern (former era, before Apr<br />

1999; latter era, after Apr 1999). Local ethic board approved<br />

the study but written consent form was waived due to the retrospective<br />

manner. Results: Data analyses were conducted in<br />

371 HCC patients (male 78% and 21% residing on island).<br />

Mean age was 64.4+/-9.4. Distribution of CP scores was as<br />

follows: A, 67%; B, 20%; and C, 13%. Distribution of clinical<br />

stages of HCC was as follows: I, 14%; II, 35%; III, 31%; IVa,<br />

18%; and IVb, 1.4%. Patients in the latter era (n=108) were<br />

diagnosed at earlier stages (Stages I+II) than that in the former<br />

era (n=263) (P=0.046). During the median observation period<br />

of 814 days, the latter era had a significantly longer survival<br />

rate than that of the former era (P


472A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

alcohol-attributable liver disease in USA is unknown. Therefore,<br />

the aim of this study is to investigate the epidemiological<br />

trends of alcohol liver disease in USA. Methods: The UNOS<br />

liver transplant registry and National Vital Statistics Systems<br />

(NVSS) was assessed for individuals with alcohol liver disease<br />

(ALD). Using the UNOS database, we identified all adult (><br />

18 years) candidates during the 10-year study period (2004<br />

– 2013) who were (i) added to liver transplant waiting list<br />

and (ii) candidates who received liver transplant (LT). Using<br />

the NVSS database, we identified all adults (> 18 years) with<br />

a cause of death attributed to alcohol liver disease using the<br />

ICD code (K70). Data extracted from both (UNOS and NVSS)<br />

databases included patient demographics (i.e. age, gender<br />

and ethnicity), and from the UNOS database (MELD score and<br />

UNOS regions). The age groups in the UNOS database are<br />

categorized as (i) 18 – 34 years (ii) 35 – 49 years (iii) 50 – 64<br />

years and (iv) 65 + years and in the NVSS database (i) 18 –<br />

24 years (ii) 25 – 44 years (iii) 45 – 54 (iii) 55 – 64 years<br />

and (iv) 65 + years. Results: During the study period 15,608<br />

candidates with ALD were added to the waiting list (WL) and<br />

6,502 received liver transplant. In the NVSS mortality dataset,<br />

there were 150,613 with a primary cause of death attributed<br />

to ALD during the study period. In the latter dataset, among all<br />

individuals with a diagnosis of cirrhosis (ALD and non-ALD),<br />

there was a significant increase in mortality for the proportion<br />

with ALD from 46% (2004) to 50% (2013). However, age-specific<br />

analysis showed that the increase in mortality rate was<br />

only significant in the age group < 55 years. Regarding transplantation<br />

trends there was an increase in both the addition of<br />

candidates with ALD to the WL and increase in liver transplant<br />

for candidates with ALD, but only in those in the age group<br />

35 – 49 years. In this age group, there was an increase in WL<br />

from 14.2% (2004) to 28.4% (2013) and increase in LT from<br />

11% (2004) to 25.3% (2013). There was no change in either<br />

WLR or LT in the other age groups. Conclusion: The burden of<br />

ALD in USA in the last decade is increasing. ALD is becoming<br />

a major public health crisis in the younger age group, and it is<br />

now the most common indication for both liver transplant and<br />

wait list addition in the age group 35 -49 years.<br />

Disclosures:<br />

Edson S. Franco - Grant/Research Support: Biospheres Medical<br />

Angel Alsina - Speaking and Teaching: Bayer, Novartis<br />

The following authors have nothing to disclose: Nyingi M. Kemmer, Thure Caire<br />

527<br />

Evaluation of DAA (direct-acting antiviral) access across<br />

the US: the interplay between Hepatitis C (HCV) patient<br />

and insurance type<br />

Jason Katz, Samantha Fernando, Antoinette Wilson, James<br />

Deemer, Marijke Frantsen; Ipsos Healthcare, New York, NY<br />

BACKGROUND In the last 18 months, highly effective DAAs<br />

have become available for HCV treatment, but the cost of these<br />

drugs has made access challenging. In this paper we examine<br />

treatment decision-making across payers by patient type.<br />

METHOD Ipsos Healthcare’s HCV Therapy Monitor, running<br />

since 2005 in the USA, reports on >150 physicians per quarter,<br />

previously annually, across the USA. Physicians provide<br />

patient demographic, disease and treatment data on treated<br />

and untreated HCV patients seen within each study period.<br />

Here, new DAAs include all sofosbuvir- and dasabuvir-containing<br />

regimens. RESULTS Following the launch of sofosbuvir<br />

in December ‘13, HCV treatment rates increased from 12%<br />

to 23% from Q1’14 to Q1‘15. During Q1‘15, rates of treatment<br />

with new DAAs differed across insurance types; 25%<br />

of all privately insured patients were on new DAA regimens,<br />

versus 21% and 16% of Medicare and Medicaid patients,<br />

respectively. All untreated patients who had payer restrictions<br />

as a treatment barrier, regardless of payer type, rose from<br />

5% to 22% from Q1’14 to Q1’15. In Q1’15, physicians’ top<br />

barriers to prescribing new DAAs were restricted to certain<br />

patient type (24%) and insurance coverage (19%). Publically<br />

insured patients are more likely to have advanced fibrosis<br />

(F3-F4) (46%) than privately insured patients (29%). Amongst<br />

more advanced patients across all payer types, F3 patients<br />

rather than F4 patients are most likely to be treated with a<br />

new DAA; this is also observed throughout ‘14. This trend is<br />

more pronounced at the Medicaid group where half of currently<br />

untreated F4 patients are substance abusers and 92%<br />

suffer from another condition. Of the whole Q1’15 untreated<br />

population, 33% are substance abusers and 77% suffer from<br />

another concomitant disorder. The population treated with a<br />

new DAA includes 16% substance abusers and 66% suffering<br />

from comorbidity. CONCLUSION Treatment of HCV patients<br />

with a new DAA is heavily influenced by fibrosis, payer type,<br />

comorbidities, and substance abuse status. Publically insured<br />

patients are more likely to have advanced fibrosis than the<br />

privately insured but the latter are more likely to be on a new<br />

DAA. Untreated patients are more likely to be substance abusers<br />

and suffer from comorbidities. These results reflect the adjudications<br />

made by US payers, many deciding to treat patients<br />

with advanced fibrosis who are free of conditions that could<br />

mitigate treatment outcomes. This also highlights that a large<br />

proportion of the population suffering from advanced fibrosis<br />

currently face access barriers to new DAAs because of their<br />

patient profile and insurance coverage.<br />

Disclosures:<br />

Jason Katz - Consulting: Abbvie, Gilead, Bristol Myers Squibb<br />

The following authors have nothing to disclose: Samantha Fernando, Antoinette<br />

Wilson, James Deemer, Marijke Frantsen<br />

528<br />

Geriatrics and cirrhosis: changing epidemiology of<br />

chronic liver disease among the elderly, 2004-2013<br />

Michael Hagan, Maria A. Kouznetsova, Sumeet K. Asrani; Baylor<br />

University Medical Center, Dallas, TX<br />

Introduction Chronic liver disease (CLD) morbidity has<br />

increased over the last decade, with a marked increase in<br />

admissions among the elderly. Methods: We examined all CLD<br />

related encounters (2004-13) in the elderly (>65 years) in the<br />

Baylor Health Care System serving Dallas-Fort Worth in rural,<br />

urban and community settings (8 hospitals, catchment 7 million,<br />

>130,000 annual admissions). Results: There were 26,816<br />

CLD related admissions (2004-2013); 23% of the admissions<br />

were in the elderly. Elderly patients had a higher prevalence of<br />

NAFLD/cryptogenic cirrhosis (44%) vs. 31% (45-54) and 41%<br />

(55-64) and primary liver malignancy (6.6%) vs. 3.3% (45-54)<br />

and 4.5% (55-64). Comorbidities were high: CAD (25%), CHF<br />

(22%), COPD (13%), DM (44%). Infection was higher in the<br />

elderly (33%) vs. 26% (45-54) and 25% (55-64). Disposition<br />

to home was lowest (51%) and inpatient mortality/hospice was<br />

highest in the elderly (20%) vs. 11% (45-54) and 14% (55-64).<br />

From 2004-05 to 2012-13, the largest increase in admissions<br />

were ages 55-64 (126% increase) and 65+ (83% increase).<br />

In the elderly, admissions increased three fold (5.5 to 14.4%).<br />

NAFLD/cryptogenic liver disease increased (32% to 50%) and<br />

viral hepatitis decreased (17% to 9.5%). Comorbid conditions<br />

increased (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 473A<br />

26% to 19%. Paracentesis and thoracentesis increased, 12%<br />

to 20%. Conclusion: Elderly subjects with CLD present with<br />

more comorbidities, more complicated liver disease, higher<br />

prevalence of NAFLD rather than viral hepatitis, higher procedure<br />

utilization and substantial inpatient mortality. The epidemiology<br />

of CLD in the elderly is changing and warrants chronic<br />

disease management models to mitigate the burden expected<br />

to be borne by Medicare.<br />

Disclosures:<br />

The following authors have nothing to disclose: Michael Hagan, Maria A. Kouznetsova,<br />

Sumeet K. Asrani<br />

529<br />

Communication, sequence and system modeling of GI/<br />

Hepatology outpatient and inpatient work flows for targeting<br />

clinical decision support tools<br />

Anne Miller 3,4 , Jejo D. Koola 3,4 , Michael E. Matheny 3,4 , Julie<br />

Ducom 1 , Jason M. Slagle 4 , Erik J. Groessl 1,5 , Sterling M. Dubin 1,2 ,<br />

Freneka F. Minter 3,4 , Jennifer H. Garvin 6,7 , Matthew B. Weinger 4 ,<br />

Samuel B. Ho 1,5 ; 1 VA San Diego Healthcare System, San Diego,<br />

CA; 2 Medicine, University of California, San Diego, San Diego,<br />

CA; 3 VA Tennessee Valley Healthcare System, Nashville, TN;<br />

4 Vanderbilt University, Nashville, TN; 5 University of California,<br />

San Diego, San Diego, CA; 6 VA Salt Lake City Healthcare System,<br />

Salt Lake City, UT; 7 University of Utah, Salt Lake City, UT<br />

Objectives: Increasing numbers of cirrhotic patients have high<br />

re-admission and mortality rates. Interventions that improve<br />

care coordination with well-targeted clinical decision support<br />

(CDS) can improve outcomes. We analyzed work processes<br />

in outpatient and inpatient settings to identify opportunities for<br />

CDS placement. Methods: Ethnographic observations, guided<br />

by user-centered design principles, occurred at two VA medical<br />

centers. Within 24 hours, pairs of experienced observers met<br />

to interpret their notes. The debriefings elucidated the structure<br />

and meaning of work as practiced resulting in: 1) Affinity diagrams<br />

to aggregate interpreted themes, insights & intervention<br />

design ideas; 2) Communication models to elucidate information<br />

flows across roles; 3) Sequence models to define work<br />

in goal-oriented terms, and 4) Artifact analyses to elucidate<br />

cognitive & problem solving structures. Results: At Site A, 26<br />

physicians, 3 nurses and 1 clerk were observed in: 1) gastrointestinal<br />

(GI) outpatient clinics (31%); 2) inpatient medical<br />

rounds (69%). At Site B, 14 physicians were observed in: 1)<br />

GI outpatient (14%) & Primary Care clinics (29%); 2) Emergency<br />

Department (21%); 3) inpatient medical rounds (29%)<br />

and 4) inpatient GI consult rounds (7%). 2/3 of participants<br />

were trainees & 1/3 were attending physicians. We categorized<br />

168 and 147 notes, respectively, during debriefings<br />

into 3 major themes each with several sub-themes. Team and<br />

Care Coordination addressed work allocation among local<br />

and distributed providers. Alerting and Reviewing addressed<br />

information access and retrieval so as to confer meaning to<br />

patients’ situations. Executing patients’ plans-of-care addressed<br />

the implementation of clinical goals & tasks. We identified<br />

resident physicians as the primary vertical and horizontal communication<br />

and coordination hubs. Using artifact analyses we<br />

found that residents’ paper patient printouts most effectively<br />

supported their clinical work. Using sequence analysis we identified<br />

patient assessment & planning as critical resident tasks<br />

to be targeted for CDS interventions (e.g., automated dashboards,<br />

data visualization, diagnostic and therapeutic checklists).<br />

Conclusions: The success of quality improvement depends<br />

on how well interventions are integrated into routine work.<br />

This study identified 4 intervention design goals to enhance<br />

cirrhosis care: 1) interventions should assume distributed work<br />

environs; 2) target residents; 3) integrate CDS with clinical<br />

assessment & planning processes rather than global point of<br />

order entry; and 4) be provided in both electronic & hardcopy<br />

forms. The design/integration of specific CDS tools is ongoing.<br />

Disclosures:<br />

Samuel B. Ho - Grant/Research Support: Genentech, Gilead; Speaking and<br />

Teaching: Prime Education, Inc<br />

The following authors have nothing to disclose: Anne Miller, Jejo D. Koola,<br />

Michael E. Matheny, Julie Ducom, Jason M. Slagle, Erik J. Groessl, Sterling M.<br />

Dubin, Freneka F. Minter, Jennifer H. Garvin, Matthew B. Weinger<br />

530<br />

Reduction of 30 day Readmission Rate in Patients with<br />

End Stage Liver Disease<br />

Simona Rossi 1 , Manisha Verma 1 , Catherine Reynolds 2 , Dee Morrison<br />

2 , June Smith 2 , Cindy Mcglone 1 , Eyob L. Feyssa 1 , Victor J.<br />

Navarro 1 ; 1 Medicine, Albert Einstein Medical Center Philadelphia,<br />

Philadelphia, PA; 2 Nursing, Albert Einstein Medical Center Philadelphia,<br />

Philadelphia, PA<br />

BACKGROUND: Patients with end stage liver disease (ESLD) are<br />

a unique population prone to high hospital readmission rates,<br />

as evidenced by a 37% thirty day readmission rate (Volk M,<br />

2012) compared to a 20% readmission rate in the Medicare<br />

population of patients excluding those with liver disease. AIM:<br />

We tested an intervention designed specifically for patients with<br />

ESLD, intended to reduce 30 day readmission rates. METHODS:<br />

Using tools described in the evidence based literature for non-<br />

ELSD as a starting point, a liver specific discharge process was<br />

developed comprising 1) patient and family centered nurse<br />

administered medication training for adherence and side effect<br />

monitoring accompanied by performance feedback to assure<br />

understanding, 2) salt and free water restriction counseling,<br />

and 3) 48 hour post-discharge scripted telephone interviews<br />

for symptom and medication side effect monitoring, and follow<br />

up adherence. This process was implemented in June 2014.<br />

Nurses were educated to perform these interventions. Positive<br />

findings (e.g.; new symptoms) reported during the phone interview<br />

were managed with pre-determined responses, including<br />

electronic alerts to a physician for immediate intervention.<br />

Mean paired differences were compared for 30 day readmission<br />

rates during the six month period prior to implementation<br />

of the process, following a standard discharge process, and for<br />

the six months after implementation of the new discharge process.<br />

As this was a quality improvement exercise, all patients<br />

admitted to the Inpatient Liver Unit of the Einstein Medical Center,<br />

Philadelphia, were included. Results: The mean (SD) 30<br />

day hospital readmission rate for the six month period prior<br />

to implementation was 30.18(6.76), compared with 21.83<br />

(8.3) following implementation (p=0.19). Conclusion: In the<br />

ESLD population, strategies to reduce hospital readmissions


474A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

are needed. Our process demonstrated the possibility that a<br />

reduction in readmission rates could be achieved with a multifaceted<br />

process that incorporates nurse directed education with<br />

performance feedback and telephonic symptom assessment,<br />

coupled with rapid provider communications. A larger study<br />

will be required to confirm these findings.<br />

Disclosures:<br />

Simona Rossi - Consulting: BMS; Speaking and Teaching: gilead<br />

Eyob L. Feyssa - Advisory Committees or Review Panels: Gilead; Grant/Research<br />

Support: J&J, BMS, Abbot, Salix, Conatus, Cumberland, Merck; Speaking and<br />

Teaching: Gilead, BMS, Abbvie<br />

The following authors have nothing to disclose: Manisha Verma, Catherine Reynolds,<br />

Dee Morrison, June Smith, Cindy Mcglone, Victor J. Navarro<br />

531<br />

Thirty-day Readmission Following Liver Transplant in<br />

Medicare Beneficiaries – Incidence, Etiology, Outcomes,<br />

and Cost<br />

Lorna M. Dove 1 , Mary Jo Braid-Forbes 2 , Kevin F. Forbes 2 , Daniel<br />

J. Lerner 3 ; 1 Internal Medicine, Columbia University College of<br />

Physicians and Surgeons, New York, NY; 2 Braide-Forbes Health<br />

Research, Silver Spring, MD; 3 Health Sciences West, Scarsdale,<br />

NY<br />

BACKGROUND: Readmission of Medicare beneficiaries (MBs)<br />

within 30 days of hospitalization (30dRA) is considered a<br />

quality indicator, and it increases the total amount of annual<br />

Medicare spending. In 2015 the Centers for Medicare and<br />

Medicaid Services (CMS) can reduce reimbursements by up to<br />

3% for hospitals with high 30dRA rates for specific conditions.<br />

More than one-third of liver transplant (LT) recipients in the US<br />

are MBs, but little is known about 30dRA following LT in this<br />

group. PURPOSE: The purpose of this study is to define the<br />

incidence, etiology, outcomes, and cost of 30dRA following<br />

LT in adult MBs. METHODS: A total of 1,359 adult Medicare<br />

fee-for-service beneficiaries who underwent LT and survived<br />

to discharge in 2011 were identified and used to define the<br />

incidence of 30dRA following LT, and the principal diagnoses,<br />

outcomes, patient characteristics, and inpatient costs associated<br />

with those 30dRA. RESULTS: The incidence of 30dRA<br />

following LT in MBs was 37%. This is substantially higher than<br />

published 30dRA rates for MBs following acute myocardial<br />

infarction (AMI), heart failure (HF), pneumonia, and coronary<br />

artery bypass grafting (CABG). The most common principal<br />

diagnoses for 30dRAs were complications of a device implant<br />

or graft (21.5%) or medical or surgical care (12.9%), fluid and<br />

electrolyte disorders (5.4%), septicemia (4.4%), and acute/<br />

unspecified renal failure (4.0%). The most common procedures<br />

at readmission were transfusion of blood products (11.7%),<br />

non-cardiac vascular catheterization (8.4%), and non-operating<br />

room GI procedures (8.2%). Factors associated with<br />

increased risk of readmission after risk adjustment were being<br />

Medicaid eligible (dual eligible) (OR 1.64, 95% CI [1.26-<br />

2.13]), and having renal failure at the time of transplantation<br />

(OR 1.44 95%CI [1.06-1.94]). The association of dual-eligibility<br />

with 30dRA is important because more than one-third<br />

of the MBs who underwent LT were dual-eligible, significantly<br />

more than in the general Medicare population. The mean cost<br />

of a 30dRA was $21,054. One year after hospitalization<br />

for LT, patients who had a 30dRA had significantly reduced<br />

survival (11.7% vs. 3.7%; P=0.01), and their total inpatient<br />

care costs were 34% higher, compared to patients without a<br />

30dRA ($252,985 vs. $189,352). CONCLUSIONS: In MBs,<br />

the 30dRA rate following LT was 37%, higher than reported<br />

for AMI, HF, pneumonia, or CABG. Complications were the<br />

most common cause for 30dRA. Patients with a 30dRA had<br />

decreased survival and increased inpatient costs at one year.<br />

Future research should focus on strategies to reduce complications<br />

from LT, and reduce 30dRA in patients at increased risk.<br />

Disclosures:<br />

The following authors have nothing to disclose: Lorna M. Dove, Mary Jo Braid-<br />

Forbes, Kevin F. Forbes, Daniel J. Lerner<br />

532<br />

Epidemiology and survival in patients with cirrhosis in a<br />

large commercial insurer database<br />

Steven J. Scaglione 1,3 , Leanne N. Metcalfe 2 , Stephanie Kliethermes<br />

3 , Rebecca Tsang 1 , Vik Goyal 4 , Allyce Caines 4 , Shaham<br />

Mumtaz 4 , Amy Luke 3 , Scott Cotler 1 ; 1 Hepatology, Loyola University<br />

Medical Center, Maywood, IL; 2 Enterprise Clinical Analytics,<br />

Health Care Services Corporation, Chicago, IL; 3 Preventive Medicine,<br />

Loyola University Medical Center, Maywood, IL; 4 Internal<br />

Medicine, Loyola University Medical Center, Maywood, IL<br />

Background: There are limited data regarding the epidemiology<br />

of cirrhosis in the general population (Scaglione, Volk,<br />

2014). We analyzed a large commercial insurance provider<br />

database to investigate the prevalence of cirrhosis, differences<br />

in survival in compensated versus decompensated cirrhosis,<br />

and differences in survival between NASH and HCV-related<br />

cirrhosis with and without diabetes. Methods: Retrospective<br />

matched case-control analysis estimating the prevalence in<br />

clinical-demographic, risk factors and survival associated with<br />

cirrhosis from Sept 2008-Aug 2013 by Health Care Services<br />

Corporation (HCSC), a commercial insurer providing healthcare<br />

coverage in 5 states. Cirrhosis was defined as ICD-9 code<br />

571.2, 571.5, or 571.6. The index date was the date of the<br />

first cirrhosis ICD-9-code occurring in the database after the<br />

first 6 months of continuous enrollment. Decompensated cirrhosis<br />

was defined as ICD-9 571.2, 571.5, or 571.6 AND a<br />

2nd diagnostic code for decompensation (any ICD-9: 070.44,<br />

070.71, 348.3x, 456.0, 456.1, 456.2x, 572.2, 572.3,<br />

572.4, 782.4, 789.59). Compensated cirrhosis is defined<br />

as those with ICD-9 571.2, 571.5, 571.6 and lacking ICD-9<br />

for decompensation as above. HCV Cirrhosis was defined by<br />

diagnostic codes for chronic HCV (ICD-9: 070.54 or 070.70)<br />

on at least two occasions a month or more apart. NAFLD<br />

was defined as by ICD-9 571.8 at any time in the absence<br />

of codes for other causes of liver disease. Results: During the<br />

study period, there were 18,202 persons with cirrhosis among<br />

3,478,093 enrollees, yielding a prevalence of 0.52%. Cirrhotics<br />

were older (55.6 vs 43.7,p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 475A<br />

533<br />

Charges for Patients Hospitalized with Alcoholic Cirrhosis<br />

Exceed All Other Etiologies of Cirrhosis Combined<br />

and are Driven by Readmissions and Volume<br />

Monica Schmidt 1 , Paul H. Hayashi 2 , Ramon Bataller 2 , Alfred S.<br />

Barritt 2 ; 1 UNC Liver Center & Gillings School of Global Public<br />

Health, University of North Carolina Chapel Hill, Chapel Hill, NC;<br />

2 UNC Liver Center, Chapel Hill, NC<br />

Background:Total charges for hospitalized patients with cirrhosis<br />

are greater for alcoholic liver disease than for all other<br />

etiologies of cirrhosis combined. It is not known if costs are<br />

driven by a higher volume of hospitalizations, longer stays,<br />

or readmissions. Methods:We used the HCUP State Inpatient<br />

Database (SID) files for New York and Florida to determine<br />

costs, readmission rates, and predictors of 30-day readmission<br />

for patients with cirrhosis from 2010-2012. The SID<br />

provides longitudinal data for individual patients. A random<br />

effects model was used to evaluate predictors of 30 day readmissions.<br />

Results:215,886 discharges and 96,295 patients<br />

with cirrhosis were identified. Alcoholic cirrhosis accounted<br />

for 53% of total admissions and 51% of cirrhosis related<br />

inpatient charges, totaling $6B between 2010-2012. Costs<br />

per admission were $53,838 for alcohol and $56,326 for<br />

non-alcoholics. Alcoholics had greater aggregate charges at<br />

the index admission and 30, 60 and 90 days(Figure 1). 30<br />

day readmission rates were 14% for alcoholics and 12% for<br />

non-alcoholics(p<br />

31 IU/L in women, ALT> 40 IU/L or AST> 37 IU/L in men)<br />

and excessive alcohol use (>20 g/day in men, >10 g/day in<br />

women). Hepatitis C (HCV) infection was defined as detectable<br />

HCV RNA. Presumed NAFLD was defined as elevated liver<br />

enzymes in the absence of indicators of chronic hepatitis B<br />

and C infection (negative antibody serology) and no excessive<br />

alcohol use. Type 2 diabetes (DM) was defined as the use of<br />

glucose-lowering agents or fasting blood glucose >125 mg/dL.<br />

The NHANES medical history questionnaire was used to determine<br />

the presence of major chronic diseases (cardiovascular,<br />

pulmonary, renal, malignancies) which were further tested as<br />

predictors of mortality in a Cox proportional hazard model.<br />

Results: In 1999-2012, 11,726 BB subjects [age at enrollment<br />

50.2±0.3 years, 72.5% White, 10.9% Hispanics, 11.4% African-American,<br />

36.5% obese (BMI≥30), 10.6% with DM] were<br />

included. The most common cause of CLD in BB was NAFLD<br />

(11.08% of BB population), followed by ALD (2.82%) and HCV<br />

(2.27%). These prevalence rates in the leading-edge BB and<br />

late BB were: NAFLD: 10.4±0.6% vs. 11.7±0.6%, p=0.11;<br />

ALD: 2.0±0.3% vs. 3.5±0.4%, p=0.0044, HCV: 2.8±0.3% vs.<br />

3.6±0.4%, p=0.09. During average 92.4 months of follow-up,<br />

355 (3.84%) BB participants with mortality data died. Multivariate<br />

analysis showed that HCV infection [adjusted hazard ratio<br />

(aHR) = 2.06 (1.03-4.11)] was independently associated with<br />

increased mortality in BB. Additionally, older age [aHR = 1.05<br />

(1.02-1.08) per year], DM [HR = 1.90 (1.21-2.98)], hypertension<br />

[aHR = 2.19 (1.58-3.04)], excessive alcohol use [aHR<br />

= 2.62 (1.73-3.97)], the presence of cardiovascular diseases<br />

[aHR = 1.70 (1.06-2.72)], and chronic kidney disease [aHR =<br />

3.46 (1.87-6.39)] were associated with mortality in BB population.<br />

Conclusions: NAFLD, ALD and HCV are common causes<br />

of CLD in BB. However, only HCV remains an independent<br />

CLD predictor of mortality in BB cohort. As more HCV patients<br />

are being cured, NAFLD will most likely account for most of the<br />

CLD burden in the BB cohort.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.


476A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Maria Stepanova, Sarah Elfeky,<br />

Alexandrea Srishord, Carey Escheik, Leo Mclaughlin, Mariam Afendy, Robert J.<br />

Wong, Sharon L. Hunt<br />

535<br />

A New Approach to Patient Centered Care in Hepatology:<br />

Patient Reported Outcomes Assessment<br />

Manisha Verma 1 , Victor J. Navarro 1 , Shana D. Stites 2 , Eyob L.<br />

Feyssa 1 , Simona Rossi 1 ; 1 Hepatology, Einstein Healthcare Network,<br />

Philadelphia, PA; 2 Center for Urban Health, Einstein Healthcare<br />

Network, Philadelphia, PA<br />

Background: Patient Reported Outcomes (PRO) are measures of<br />

symptoms, health behavior &/or experiences reported directly<br />

by patients. Integrating PRO Assessment (PA) in routine care<br />

improves patient centered care (PCC). PROMIS-29 (Patient<br />

Reported Outcomes Measurement Information System) is a<br />

web-based NIH supported validated set of 29 questions used<br />

for PA. Our aim is to identify if PA influences PCC in patients<br />

with cirrhosis, through enhanced communication, & facilitating<br />

detection of physical/ psychological issues. Methods: This is<br />

a prospective nonrandomized study of the feasibility, acceptability,<br />

& clinical value of PA in cirrhotics. PA included 1)<br />

provider training in PA score interpretation, 2) administering<br />

PROMIS-29, & 3) reporting PA results (T-scores) to patients &<br />

providers for point of care use. T-score measures the symptom<br />

assessed. Higher scores reflect greater symptom assessed. All<br />

outpatients with cirrhosis were eligible. Patient and provider<br />

feedback surveys were administered after each visit to assess<br />

acceptability/feasibility. Descriptive statistics were used to<br />

assess baseline measures; paired t-tests were used to compare<br />

baseline and 3 month data. Results: 50 patients were enrolled;<br />

3 month data were available for 20. There were no statistically<br />

significant differences in baseline measures based on<br />

gender, race, or duration of liver disease (p>0.05). At baseline,<br />

pain, anxiety, & fatigue were worse in our sample vs the<br />

US population (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 477A<br />

537<br />

Physician versus Patient Perceptions of Medical Care<br />

Quality in Primary Biliary Cirrhosis<br />

Andrew Saich, Herbert Swanson, Tracy J. Mayne; Intercept Pharmaceuticals,<br />

Inc., San Diego, CA<br />

Numerous <strong>studies</strong> have demonstrated significant self-enhancement<br />

bias in physician perceptions of medical care delivery, in<br />

which physicians rate care delivery more highly than patients.<br />

Objective: To compare patient and physician perceptions of<br />

care in patients with primary biliary cirrhosis (PBC) and the<br />

hepatologists and gastroenterologists treating PBC patients,<br />

with the goal of improving care delivery. Methods: From<br />

December 11, 2014 to January 12, 2015, we conducted surveys<br />

with board certified gastroenterologists (262) and hepatologists<br />

(60) practicing for at least two years and had treated<br />

at least two PBC patients in the past six months. From January<br />

5-30, 2015, we conducted surveys with 214 patients with PBC<br />

who were under a physician’s care. Results: The figure shows<br />

physician and patient ratings of their confidence in physician<br />

delivery of 7 areas of PBC care. Only 2 areas differed by more<br />

than 5 points: physicians under-rated their skills in helping<br />

patients manage PBC symptoms by a mean 6 points, and overrated<br />

their bedside manner by a mean 10 points, compared to<br />

patient ratings. Beyond care provisions, there were meaningful<br />

differences between patients and physicians perceptions on<br />

how PBC affects patient health-related quality of life. Half of<br />

PBC patients reported that PBC impacts a “great deal” how<br />

they feel physically (50% vs. 25% physicians), their ability<br />

to work (39% vs. 16% physicians), and their ability to do the<br />

things they enjoy (36% vs. 17% physicians). Despite the impact<br />

of PBC on these symptoms, only 42% of patients strongly agree<br />

that additional treatment options would treat PBC more effectively<br />

(versus 65% of physicians). Conclusion: Hepatologists<br />

and gastroenterologists do not demonstrate systematic self-enhancement<br />

bias in treating PBC patients. While they may overrate<br />

their bedside manner, they underrate their management<br />

of PBC symptoms. However, physicians and patients both<br />

rated diagnosis and treatment more highly than monitoring<br />

disease progression and managing symptoms, and physicians<br />

recognize the need to improve practice. If physicians’ estimation<br />

and management of patient quality of life improves, then<br />

patient perception of the treatment effectiveness and optimism<br />

toward treatment innovation may also improve. Interventions to<br />

improve physician communication around disease progression,<br />

and patient communication around symptoms and symptom<br />

management could enhance practice.<br />

Disclosures:<br />

Herbert Swanson - Management Position: Intercept<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: Andrew Saich<br />

538<br />

Redesigning Outpatient Care to Reduce 30-day Hospital<br />

Readmissions Among Recent Liver Transplant Recipients.<br />

Chanda Ho, Nina Topic, Raphael Merriman, Robert W. Osorio,<br />

Garrett M. Hisatake, R Todd Frederick; California Pacific Medical<br />

Center, San Francisco, CA<br />

Background: Hospital readmission after liver transplantation<br />

(LT) is associated with reduced patient and graft survival and<br />

represents a growing health care burden. Prevention of hospital<br />

readmission has been an area of focus by the Centers<br />

for Medicare and Medicaid Services with reduced payments<br />

for readmissions for certain diagnoses. However, hospital<br />

readmissions and strategies to reduce them among recent LT<br />

recipients have not been well characterized. Methods: We<br />

collected baseline hospital readmission data following LT to<br />

our medical center from 2010-2012. Reasons for readmission<br />

were reviewed by a hepatologist and a surgeon. We captured<br />

median length of stay (LOS) both after LT and during readmission.<br />

Based on these data, in October 2012, we implemented<br />

a quality improvement project to reduce our readmission rates.<br />

Readmission data post-intervention were collected for 2013<br />

and 2014. Results: Our 30-day post-LT hospital readmission<br />

rates in 2010, 2011, and 2012 were 40%, 42%, and 43%<br />

respectively. A high proportion of readmissions had a short<br />

LOS (≤ 2days):47% and 48% in 2011 and 2012, respectively.<br />

The reviewers determined a significant proportion of these<br />

admissions could have been managed in the outpatient setting.<br />

Barriers to care management identified included difficulty<br />

scheduling multiple, often complex procedures in the outpatient<br />

setting. The organization of post transplant care involving multiple<br />

expedited procedures (e.g. interventional radiology, endoscopy)<br />

was streamlined to optimize outpatient care coordination<br />

and delivery for the commonly required diagnostic and therapeutic<br />

procedures (e.g. ERCP, liver biopsy, ultrasound). Thirty-day<br />

readmission rates declined in 2013 and 2014 to 25%<br />

and 26%, respectively (p


478A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and 1965. However, many eligible patients are not being<br />

screened. Aim: Determine the rate adherence to the CDC<br />

recommendations, and evaluate differences in demographics<br />

between those screened to those who were not. Methods:<br />

Single center-center, retrospective cohort study of all patients<br />

born between 1945-1965 who were seen at the Primary Care<br />

Group (PCG), Emergency department (ED), Infectious Disease<br />

(ID), Gastroenterology Department (GI) from August 17, 2012<br />

to November 4, 2014 were included in the study. Insurance,<br />

race, results of HCV related laboratory testing were included<br />

in the study. Results: A total of 24965 patients (10 499 males)<br />

born between 1945 and 1965 and no known history or documented<br />

HCV screening history were seen at the University of<br />

Chicago’s ED, PCG, ID and GI. Of the 24 965 eligible individuals<br />

2,104 were screened for chronic HCV (889 males).<br />

Unscreened individuals were younger compared to screened<br />

(56 vs. 58, respectively (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 479A<br />

reflect the position or policy of the Department of Veterans<br />

Affairs or the United States government.<br />

Disclosures:<br />

Christine M. Hunt - Consulting: Otsuka<br />

Ayako Suzuki - Consulting: Novartis<br />

Hans L. Tillmann - Consulting: Novartis; Employment: AbbVie; Grant/Research<br />

Support: Novartis; Stock Shareholder: AbbVie, Abbott, Gilead<br />

Joseph K. Lim - Consulting: Merck, Boehringer-Ingelheim, Gilead, Bristol Myers<br />

Squibb, Janssen; Grant/Research Support: AbbVie, Boehringer-Ingelheim, Bristol<br />

Myers Squibb, Gilead, Hologic, Janssen<br />

The following authors have nothing to disclose: Lauren A. Beste, George N.<br />

Ioannou, Elliott Lowy, Dawn T. Provenzale, Michael J. Kelley, Cynthia A. Moylan,<br />

Omobonike O. Oloruntoba, David Ross<br />

541<br />

Comparable hepatitis C treatment outcomes with oral<br />

direct-acting antivirals between rural outreach clinics<br />

and an academic medical center: A case for expanding<br />

task-shifting<br />

Channa R. Jayasekera, Ryan B. Perumpail, Aijaz Ahmed; Division<br />

of Gastroenterology and Hepatology, Stanford University Medical<br />

Center, Stanford, CA<br />

Purpose: Residence in medically underserved areas is associated<br />

with lower likelihood of receiving hepatitis C treatment.<br />

The safety, tolerability, and effectiveness of oral direct-acting<br />

antivirals (DAA) may obviate the need for specialist expertise<br />

in many patients, and decentralizing treatment from specialists<br />

to more abundant community-based providers may improve<br />

treatment access without compromising safety. At our rural<br />

outreach clinics (OCs), a visiting hepatologist evaluates treatment<br />

eligibility but subsequent routine management is devolved<br />

to licensed vocational nurses (task-shifting). The hepatologist<br />

reviews laboratory results remotely and is available for consultation<br />

and as-needed clinic visits. We assessed the outcomes<br />

of patients treated at these OCs versus those treated directly<br />

by the hepatologist at an academic medical center (AMC).<br />

Methods: We retrospectively compared clinical characteristics,<br />

sustained virological response 12 weeks after treatment completion<br />

(SVR12) rates, treatment discontinuation and adverse<br />

event rates of consecutive patients treated with DAAs by a<br />

single hepatologist between Dec 2013-Sept 2014 at an AMC<br />

and three OCs. Results: 45 AMC and 48 OC patients received<br />

either sofosbuvir+ribavirin (SOF+RBV) or sofosbuvir+simeprevir<br />

(SOF+SMV). AMC and OC patients were comparable in<br />

age, gender, HCV genotype distribution, and prior treatment<br />

experience. More AMC patients were cirrhotic (73.3% vs<br />

50%, p=0.03) and received SOF+SMV (77.8% vs. 56.3%,<br />

p=0.03). Overall SVR12 rates were comparable between the<br />

AMC and OCs (88.8% vs. 83.3%, p=0.87), as were SVR12<br />

rates of non-cirrhotic, cirrhotic, treatment-naïve, treatment-experienced,<br />

genotype 1, 2, and 3 patients (figure 1). One patient<br />

discontinued treatment at each site. There were no treatment<br />

related hospitalizations or deaths. Conclusions: In this analysis<br />

of oral DAA therapy at rural outreach clinics and an academic<br />

medical center, treatment responses were comparable to published<br />

real-world data, and site of treatment or type of provider<br />

did not impact treatment outcome. This task-shifting strategy,<br />

wherein mid-level providers are empowered to manage HCV<br />

treatment in medically underserved areas with remote supervision<br />

of specialists, may hold promise in expanding HCV treatment<br />

access.<br />

Disclosures:<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

The following authors have nothing to disclose: Channa R. Jayasekera, Ryan B.<br />

Perumpail<br />

542<br />

Reducing Wait Times for Radiofrequency Ablation in<br />

Patients with Hepatocellular Carcinoma: A Quality<br />

Improvement Initiative at the Toronto Centre for Liver<br />

Disease.<br />

Mayur Brahmania 1 , Osman Ahmed 1 , Korosh Khalili 2 , Eberhard L.<br />

Renner 1 , Harry L. Janssen 1 , Morris Sherman 1 , David K. Wong 1 ,<br />

Hemant Shah 1 , Jordan J. Feld 1 ; 1 Gastroenterology, University of<br />

Toronto, Toronto, ON, Canada; 2 Medical Imaging, University of<br />

Toronto, Toronto, ON, Canada<br />

Introduction: Radiofrequency ablation (RFA) has become the<br />

recommended curative treatment option for patients with early<br />

stage hepatocellular carcinoma (HCC). Unfortunately, the<br />

expected number of HCC cases requiring RFA will grow at a<br />

greater rate than potential RFA capacity. Thus, novel strategies<br />

will be required to handle the burden. According to guidelines,<br />

waiting times for any malignancy from diagnosis to treatment<br />

should not exceed 30 days. In the current analysis we investigated<br />

if wait times for RFA were associated with outcomes such<br />

as residual disease after RFA, tumor recurrence, liver transplant<br />

or death. Methods: This was a retrospective study analysing<br />

patients with HCC between January 2012 and December<br />

2014 presenting to University Health Network (UHN) hospitals<br />

in Toronto. Multi-disciplinary Tumor Board rounds were held<br />

weekly on patients with newly diagnosed HCC. All patients<br />

receiving RFA alone as treatment for HCC were included. The<br />

time from diagnosis to presentation at Tumor Board rounds<br />

and the time from Tumor Board rounds to the date the patient<br />

received first treatment was documented. Multivariable Cox<br />

regression was used to determine factors associated with tumor<br />

outcome. Results: Of the 150 patients included in the study,<br />

82% percent were male and the median age was 62 years<br />

(Range 25-90). Median tumor size at diagnosis was 19 mm<br />

(Range 0-43 mm), mean MELD was 7 (SD=4.3) and 58% had<br />

Barcelona stage A. The cause of liver disease was hepatitis C<br />

in 41%, hepatitis B in 30%, Hepatitis B/C co-infection in 2%,<br />

and other liver diseases in 27%. The median wait time from<br />

diagnosis to Tumor Board presentation was 21 days (range<br />

12-41), wait time from Tumor Board to Interventional Radiology<br />

consultation was 30 days (21-48) and the mean time from<br />

consultation to RFA was 44 days (IQR 34-58). The median<br />

time from HCC diagnosis to RFA treatment was 98 days (IQR<br />

79-139). 30 patients died or underwent transplant. When<br />

adjusting for Barcelona staging, tumor size and MELD score,<br />

wait times >150 days was not associated with an increased<br />

risk of death or transplant (HR=1.69; p=0.23). Conclusion:<br />

Our study demonstrates wait times for RFA in patients with HCC<br />

are over the recommended waiting time in our province. Wait<br />

times over 150 days showed a non-significant trend towards


480A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

an increased risk of transplant or death. We have identified<br />

system gaps where quality improvement measures can be<br />

implemented to reduce wait times and allocate resources for<br />

future RFA treatment at our centre.<br />

Disclosures:<br />

Eberhard L. Renner - Advisory Committees or Review Panels: Vertex Canada,<br />

Novartis, Astellas Canada, Roche Canada, Gambro, AbbVIe, BMS, Gilead;<br />

Grant/Research Support: Novartis Canada, Gilead; Speaking and Teaching:<br />

Novartis, Astellas Canada, Roche Canada<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Morris Sherman - Consulting: Gilead, Merck, Bayer, Arqule, Celsion, Boehringer<br />

Ingelheim, Janssen, Abbvie; Grant/Research Support: WAKO; Speaking and<br />

Teaching: WAKO<br />

David K. Wong - Grant/Research Support: Gilead, BMS, Vertex, BI<br />

Hemant Shah - Consulting: Merck, Roche, Gilead, Janssen, BMS, Abbvie, Lupin<br />

Jordan J. Feld - Advisory Committees or Review Panels: Merck, Janssen, Gilead,<br />

AbbVie, Theravance, Bristol Meiers Squibb; Grant/Research Support: AbbVie,<br />

Boehringer Ingelheim, Janssen, Gilead, Merck<br />

The following authors have nothing to disclose: Mayur Brahmania, Osman<br />

Ahmed, Korosh Khalili<br />

543<br />

Timing of Hepatitis B and C Diagnosis Relative to Hepatocellular<br />

Carcinoma Diagnosis in the BC Hepatitis Testers<br />

Cohort (BC-HTC)<br />

Naveed Z. Janjua 1,2 , Amanda Yu 1 , Margot E. Kuo 1 , Maryam<br />

Alavi 4 , Ryan Woods 3 , Maria Alvarez 1 , Gregory Dore 4 , Mark Tyndall<br />

1,2 , Mel Krajden 1,2 ; 1 BC Centre for Disease Control, Vancouver,<br />

BC, Canada; 2 University of British Columbia, Vancouver, BC,<br />

Canada; 3 BC Cancer Agency, Vancouver, BC, Canada; 4 University<br />

of New South Wales, Sydney, NSW, Australia<br />

Background: Early diagnosis of hepatitis B (HBV) and hepatitis<br />

C (HCV) followed by treatment could prevent late stage liver<br />

disease. Hepatitis diagnosis after appearance of chronic liver<br />

disease sequelae is an indicator of sub-optimal screening. We<br />

measured the timing of HBV and HCV diagnoses relative to<br />

detection of hepatocellular carcinoma (HCC) as an indicator of<br />

late hepatitis diagnosis. Methods: The British Columbia Hepatitis<br />

Testers Cohort (BC-HTC) consists of 1,417,780 individuals<br />

tested for HCV, HIV and/or reported to public health as a case<br />

of HBV, HCV, HIV or tuberculosis from 1990-2013. Cohort<br />

data were linked with medical visits, hospitalizations, cancers,<br />

prescription drugs and mortality. Late hepatitis diagnosis was<br />

defined as HBV or HCV diagnosis within 2 years prior to or<br />

after HCC diagnosis. HCC diagnosis was based on ICD-O-3<br />

topography code C22.0 and histology codes 81703-81753 in<br />

the BC Cancer Registry. HCV diagnosis date was the date of<br />

the first positive test. HBV diagnosis date was the date HBV was<br />

reported to public health. Hepatitis diagnoses from 1990-2012<br />

and HCC from 1992-2011 were analyzed to allow for a 2<br />

year lead-in and one year follow-up time between hepatitis and<br />

HCC diagnoses. Results: Between 1992 and 2011, 2,250<br />

cases of HCC were diagnosed; 612 (27%) were HBV positive,<br />

868 (39%) HCV positive, and 47 (2%) had both. Overall,<br />

659/34,775 HBV cases (1.9%) and 915/65,842 HCV cases<br />

(1.4%) developed HCC. Among HBV and HCV cases with<br />

HCC, 45% and 31% respectively were diagnosed late. HBV<br />

late diagnosis declined from 100% in 1992 to 53% in 2000<br />

to 24% in 2011. HCV late diagnosis declined from 100% in<br />

1992 to 30% in 2000 to 13% in 2011. In the multivariable<br />

logistic regression model, birth years


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 481A<br />

both tests, the correct answer was chosen 79% of the time after<br />

the rotation. Conclusion: Our novel curriculum for a non-elective<br />

consult rotation has effectively demonstrated improvement in IM<br />

residents’ comfort with and knowledge of CLD. A trend towards<br />

increased career interest in hepatology was also observed,<br />

which suggests more exposure to CLD could positively impact<br />

recruitment to the workforce.<br />

Disclosures:<br />

Donald M. Jensen - Advisory Committees or Review Panels: Merck, Onom Foundation;<br />

Speaking and Teaching: Gilead<br />

K. Gautham Reddy - Advisory Committees or Review Panels: AASLD Transplant<br />

Hepatology Pilot Steering Committee, ACG Training Committee, Program Director’s<br />

Caucus Steering Committee, Gilead, Inercept, Bristol-Myers Squibb; Grant/<br />

Research Support: Intercept, Ocera, Merck, Lumena<br />

Nancy Reau - Advisory Committees or Review Panels: Jannsen, Merck, AbbVie,<br />

Intercept, Salix, BMS, Jannsen; Grant/Research Support: Merck, Gilead, BMS,<br />

AbbVie, Jannsen, BI<br />

Helen S. Te - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Abbvie, Conatus, BMS<br />

The following authors have nothing to disclose: Adam E. Mikolajczyk, Jeanne M.<br />

Farnan, John F. McConville, Andrew I. Aronsohn<br />

545<br />

High Rates of Hepatitis C Virus Infection Among Inpatient<br />

Baby Boomers in an Urban Hospital: a model to<br />

improve linkage to care<br />

Aparna Goel 1 , Lismeiry Paulino 2 , Douglas Dieterich 2 , Ponni V.<br />

Perumalswami 2 ; 1 Gastroenterology, Icahn School of Medicine<br />

Mount Sinai, New York, NY; 2 Liver Diseases, Icahn School of<br />

Medicine Mount Sinai, New York, NY<br />

Background: New York State passed legislation that took effect<br />

in January 2014 requiring healthcare providers to offer Hepatitis<br />

C Virus (HCV) screening to all inpatient baby boomers<br />

and link infected persons to care. Our aims are to determine<br />

the prevalence of HCV-positive baby boomers admitted to an<br />

urban, tertiary care hospital and evaluate the sequential impact<br />

of an electronic medical record (EMR) prompt, provider education,<br />

and data feedback to providers on HCV screening<br />

and linkage to care rates. Methods: The results of HCV antibody<br />

(Ab) tests for individuals born 1945-1965 with inpatient<br />

encounters at Mount Sinai were obtained from November<br />

2013 through April 2015. The type of medical service, level<br />

of provider and patient demographics were collected. HCV<br />

screening and linkage to care rates before and after a series<br />

of interventions were compared: a birth cohort-selective EMR<br />

prompt was implemented in January 2014, comprehensive<br />

education for housestaff and nurses occurred from July-September<br />

2014, monthly data feedback to nurses with screening<br />

rates began in October 2014 and finally patient navigation<br />

was implemented in May 2015. Results: There were 23,835<br />

inpatient baby boomer encounters during the study period.<br />

The median age was 60 years with 56% male and 33% Hispanic.<br />

Overall, 34% had a documented HCV Ab test and<br />

13% (n=1,036) were positive. Screening rates did not significantly<br />

rise following an EMR prompt (p=0.92), education<br />

(p=0.86) or with data feedback to nursing managers (p=0.42).<br />

Provider education and data feedback interventions led to a<br />

rise in newly diagnosed HCV-positive rates from 6% to 15%<br />

(p576,000<br />

patients receiving AMA testing between January 2010 and<br />

December 2013 in a large US commercial laboratory database.<br />

Patients were considered AMA positive if the antibody<br />

titer was > 1:20 or M2 antibody was > 25.0 U. We examined<br />

geographic distribution at a 3-zip code level. Results: 16,492<br />

patients tested AMA positive and had at least 1 ALP measure<br />

over the 4 year period. 5,380 had an ALP >120 U/l before or<br />

within 30 days of their first AMA+ test and were classified as<br />

having probable PBC. There were clear patterns of increased<br />

concentration of incident PBC patients (from west to east):<br />

Southeastern tip of the Alaskan spur; Santa Barbara/Ventura<br />

and inland central California; western Dakotas’ border; central<br />

Oklahoma including Oklahoma City; western Texas and Houston;<br />

Indiana just north of Louisville, Kentucky; an area of the<br />

Appalachians spanning southeastern New York, southern Pennsylvania,<br />

and the northern West Virginia border; and Florida<br />

(excluding the Pan Handle). These areas were not associated<br />

with age, race/ethnicity, urban concentration, or elevation.<br />

Conclusion: These analyses indicate that there are clusters of<br />

incident patients in geographically diverse areas in the United<br />

States. Additional analyses are required to determine if there<br />

may be geographic or demographic factors underlying these<br />

variations.


482A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

Herbert Swanson - Management Position: Intercept<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

The following authors have nothing to disclose: Keith D. Lindor<br />

547<br />

A qualitative assessment of factors impacting adoption<br />

and implementation of USPSTF age-based Hepatitis C<br />

Virus screening guidelines<br />

Amy B. Jessop; HepTREC, University of the Sciences in Philadelphia,<br />

Philadelphia, PA<br />

Background: Baby boomers comprise 70% of U.S. Hepatitis<br />

C (HCV) cases, yet many remain unaware of their infection.<br />

The 2012 CDC and 2013 USPSTF HCV screening guidelines<br />

that expanded to include one-time screening for those born<br />

between 1945 and 1965 were developed to increase identification<br />

of those infected. Insurer policies require or encourage<br />

screening within primary care or family practice settings,<br />

therefore, improvements will be realized only if primary care<br />

providers adopt the guidelines. This qualitative study examines<br />

the awareness and adoption of prevention practice guidelines,<br />

particularly the enhanced USPSTF HCV screening guideline in<br />

a large, suburban family practice setting. Methods: Data collection<br />

employed individual interviews conducted at the practice<br />

site using a semi-structured interview guide developed,<br />

assessed for content validity through literature and expert<br />

review and pilot testing with three practicing primary care<br />

physicians. The family practice site selected for examination<br />

is a community-based family practice and resident training<br />

program. Interviews with all resident (12) and attending physicians<br />

(5) and advanced practice nurse (NP) were taped and<br />

completed between Dec 2014 and Feb 2015. They considered<br />

knowledge and beliefs about practice guidelines (in general<br />

and the USPSTF enhanced HCV guideline in particular) and<br />

self-assessed screening rates. Researchers independently, then<br />

collectively, reviewed transcriptions, assessing responses and<br />

identifying common and emergent themes. Results: All subjects<br />

viewed practice guidelines as useful tools. Attending physicians<br />

reported self-selection of guidelines to adopt and implement<br />

while residents tended to follow advice of the attending<br />

physicians to which assigned. Source of the guideline played<br />

a deciding role in which guidelines were adopted. Awareness<br />

of the expanded USPSTF HCV screening guideline varied by<br />

position, with the NP and attending physicians most aware,<br />

although knowledge of guideline details was limited; some<br />

considered their patients to be of low-risk for HCV infection<br />

and estimates of individual and group HCV screening rates<br />

were low (2-25%). Subjects caring for an HCV-affected patient<br />

reported greater importance to screening. Smart tools, prompting<br />

screening at point of care were deemed likely to improve<br />

practice. Conclusion: Preventive practice guidelines are well<br />

accepted, but knowledge of and attitudes about age-based<br />

HCV screening guidelines is low and lack of reminders at pointof-care<br />

limit screening initiation. Improved education and systems-based<br />

approaches to improve screening practices are<br />

essential to reap benefits of expanded guidelines.<br />

Disclosures:<br />

The following authors have nothing to disclose: Amy B. Jessop<br />

548<br />

Outreach invitations improve HCC surveillance rates:<br />

A Randomized Controlled Trial in a Safety-Net Health<br />

System<br />

Amit G. Singal 1 , Jasmin A. Tiro 2 , Katharine McCallister 2 , Caroline<br />

A. Mejias 2 , Lei Xuan 2 , Ethan Halm 1 ; 1 Internal Medicine, University<br />

of Texas Southwestern, Irving, TX; 2 University of Texas Southwestern,<br />

Dallas, TX<br />

Background: Hepatocellular carcinoma (HCC) surveillance<br />

is associated with improved survival, but its effectiveness is<br />

limited by underuse, particularly in underserved populations.<br />

Aim: Compare the effectiveness of mailed outreach and patient<br />

navigation to increase HCC surveillance rates in a racially<br />

diverse cohort of patients Methods: Patients with documented<br />

or suspected cirrhosis at a safety-net health system were randomized<br />

to receive mailed outreach invitations for surveillance<br />

ultrasonography, mailed outreach plus patient navigation, or<br />

usual care. Documented cirrhosis was defined using ICD-9<br />

codes, and suspected cirrhosis was defined as AST to platelet<br />

ratio index (APRI) ≥1.5 in the presence of liver disease. We<br />

excluded those with Child C cirrhosis who were not transplant<br />

candidates and/or significant comorbid conditions given limited<br />

benefit of surveillance. Patients who did not respond to<br />

outreach invitations within 2 weeks received up to 3 reminder<br />

telephone calls. The primary study outcome was completion of<br />

surveillance ultrasound within 6 months of randomization. We<br />

used intent- to-screen principle for analyses. We are randomizing<br />

1800 patients in groups of 300 (100 patients/arm), with<br />

results of the first 2 groups presented here. Results: Baseline<br />

patient characteristics (n=600) across the arms were similar.<br />

Mean age was 55 years and 60% were men. The cohort was<br />

racially diverse with 36% Black, 35% Hispanic, and 28%<br />

White. Most (79%) had documented cirrhosis, while 21% had<br />

suspected cirrhosis. Cirrhosis was due to HCV in 54%, alcohol<br />

19%, NASH 15%, and HBV 4%. Rates of scheduled surveillance<br />

ultrasound were significantly higher in the outreach/navigation<br />

(34.0%) and outreach alone (26.5%) arms than usual<br />

care (11.5%; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 483A<br />

549<br />

Engagement in care of high risk hepatitis C patients<br />

with interferon-free therapies<br />

John Dever 1,2 , Julie Ducom 1 , Ariel Ma 1 , Michael Yang 1 , Erika<br />

Cherk 1 , Ann Herrin 1 , Erik J. Groessl 1,2 , Samuel B. Ho 1,2 ; 1 VA San<br />

Diego Healthcare System, San Diego, CA; 2 University of California,<br />

San Diego, San Diego, CA<br />

Objectives: New interferon-free direct acting antiviral (DAA)<br />

treatments have the potential to expand the numbers of patients<br />

receiving treatment for hepatitis C virus (HCV); however patient<br />

access and willingness to engage in care is unknown. Our<br />

aim was to determine patient engagement and patient-related<br />

barriers to care for accessing new interferon-free DAA in a realworld<br />

practice setting. Methods: Patients with viremic HCV and<br />

high-risk for fibrosis by FIB-4 score were identified using the<br />

national VA HCV registry within the San Diego VA Healthcare<br />

System in October 2014. Patients were categorized as actively<br />

involved in HCV clinic care or those who were not seen in<br />

HCV clinic within the previous 6 months and without a future<br />

appointment. Patients were contacted by letter and phone call<br />

and notified of the availability of new more effective DAA treatments<br />

and the consequences of not receiving treatment, and<br />

invited to schedule an appointment in HCV clinic. Patient characteristics<br />

and subsequent outcomes were determined using<br />

medical records. Logistic regression was used to examine factors<br />

associated with willingness to engage in care. Results:<br />

481 patients were found to have FIB-4 > 2.4 (307 with FIB-4<br />

> 3.25 with high likelihood of cirrhosis). Of the 481 patients,<br />

177 patients (37%) were actively followed in clinic and eligible<br />

for treatment. Treatment ineligible included 24 (5%) in<br />

hospice or deceased, 66 (14%) relocated, and 12 (2%) with<br />

severe comorbidity. 202 treatment eligible patients (42%) were<br />

identified as never seen by a HCV clinic provider or lost to follow-up.<br />

Of these, 123 (61%) remained non-engaged and gave<br />

no response to outreach (follow up 2-6 mo), 6 patients (3%)<br />

declined evaluation and treatment, and 74 (36%) responded<br />

and engaged with HCV clinic. Non-engaged patients compared<br />

with engaged were more likely to have been homeless<br />

(48 vs 28%), have active alcohol/drug use (44 vs 31%), psychiatric<br />

diagnoses (69 vs 63%), and # co-morbid disorders<br />

(3.17 vs 3.07). Significant variables in logistic regression for<br />

non-engagement included history of homelessness, COPD, and<br />

hypertension. Conclusions: Among treatment-eligible, high priority<br />

HCV patients most at risk for advanced fibrosis that were<br />

not currently attending HCV clinic, 64% were unable or unwilling<br />

to engage in HCV care after a one time outreach (32% of<br />

the total treatment eligible). Homelessness and certain comorbidities<br />

are significant factors related to non-engagement.<br />

These data indicate that more sustained or intensive outreach<br />

efforts are needed in order to effectively expand access to<br />

many of those most in need of treatment.<br />

Disclosures:<br />

Samuel B. Ho - Grant/Research Support: Genentech, Gilead; Speaking and<br />

Teaching: Prime Education, Inc<br />

The following authors have nothing to disclose: John Dever, Julie Ducom, Ariel<br />

Ma, Michael Yang, Erika Cherk, Ann Herrin, Erik J. Groessl<br />

550<br />

Thirty-day Readmission for Infection Following Liver<br />

Transplant is Associated with Increased Mortality<br />

Lorna M. Dove 1 , Daniel J. Lerner 2 ; 1 Internal Medicine, Columbia<br />

University College of Physicians and Surgeons, New York, NY;<br />

2 Health Sciences West, Scarsdale, NY<br />

BACKGROUND: Over thirty percent of patients are readmitted<br />

within 30 days of liver transplantation (30dRA). The role of<br />

infection in 30dRA of these immunosuppressed patients is not<br />

well defined. PURPOSE: The purpose of this study is to define<br />

the incidence, etiology, and cost of 30dRA for infection following<br />

liver transplantation (LT). METHODS: Utilizing the Agency<br />

for Healthcare Research and Quality’s Healthcare Cost and<br />

Utilization Project’s State Independent Databases for 4 geographically<br />

distinct states: Florida, Massachusetts, New York,<br />

and Washington, we identified patients who underwent LT in<br />

2012, based on the ICD-9-CM code for LT, 50.59. Patients<br />

with a 30dRA following LT were divided into 2 groups: those<br />

with and without an infection-related principal diagnosis (IRPD)<br />

for the 30dRA. Patients with a 30dRA for rehabilitation on the<br />

day of discharge were excluded. Approximately 8% of 30dRA<br />

were not captured because they occurred in the following<br />

year. RESULTS: The incidence of 30dRA following LT was 30%<br />

(332/1099). Nineteen percent (64/332) of those 30dRA had<br />

an IRPD. There was no significant difference in the mean age,<br />

gender, race, number of chronic conditions, primary payor, or<br />

median household income between the two groups. There were<br />

21 different IRPD, and the most common were postoperative<br />

infection (31%; 20/64), intestinal infection due to C. difficile<br />

(16%; 10/64), unspecified septicemia (13%; 8/64), and urinary<br />

tract infection (8%; 5/64). Patients with a 30dRA for an<br />

IRPD were more likely to die during that 30dRA, compared<br />

to those with a non-IRPD (5% vs. 0%; P


484A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

551<br />

Evacuation after the Fukushima Daiichi Nuclear Power<br />

Plant accident as a cause of liver dysfunction: The<br />

Fukushima Health Management Survey<br />

Atsushi Takahashi 1,2 , Tetsuya Ohira 2,3 , Mitsuaki Hosoya 2,4 , Seiji<br />

Yasumura 2,5 , Masato Nagai 2,3 , Hiromasa Ohira 1 , Shigeatsu<br />

Hashimoto 2,6 , Hiroaki Satoh 2,6 , Akira Sakai 2,7 , Akira Ohtsuru 2,8 ,<br />

Yukihiko Kawasaki 2,4 , Hitoshi Suzuki 2,9 , Gen Kobashi 10 , Kotaro<br />

Ozasa 11 , Shunichi Yamashita 2,12 , Kenji Kamiya 2,13 , Masafumi<br />

Abe 2 ; 1 Gastroenterology and Rheumatology, Fukushima Medical<br />

University School of Medicine, Fukushima, Japan; 2 Radiation Medical<br />

Science Center for the Fukushima Health Management Survey,<br />

Fukushima Medical University School of Medicine, Fukushima,<br />

Japan; 3 Epidemiology, Fukushima Medical University School of<br />

Medicine, Fukushima, Japan; 4 Pediatrics, Fukushima Medical<br />

University School of Medicine, Fukushima, Japan; 5 Public Health,<br />

Fukushima Medical University School of Medicine, Fukushima,<br />

Japan; 6 Nephrology, Hypertension, Diabetology, and Endocrinology,<br />

Fukushima Medical University School of Medicine,<br />

Fukushima, Fukushima, Japan; 7 Radiation Life Sciences, Fukushima<br />

Medical University School of Medicine, Fukushima, Japan; 8 Radiation<br />

Health Management, Fukushima Medical University School<br />

of Medicine, Fukushima, Japan; 9 Cardiology and Hematology,<br />

Fukushima Medical University School of Medicine, Fukushima,<br />

Japan; 10 Public Health, Dokkyo Medical University, Tochigi,<br />

Japan; 11 Epidemiology, Radiation Effects Research Foundation,<br />

Hiroshima, Japan; 12 Japan and Atomic Bomb Disease Institute,<br />

Nagasaki University, Nagasaki, Japan; 13 Research Institute for<br />

Radiation Biology and Medicine, Hiroshima University, Hiroshima,<br />

Japan<br />

Background and Aim: The Great East Japan Earthquake and<br />

Fukushima Daiichi Nuclear Power Plant accident caused residents<br />

to forego their normal lives and evacuate. The aim of<br />

this study was to evaluate liver function before and after this<br />

disaster. Methods: This was a longitudinal survey of 26,006<br />

Japanese men and women living near the Fukushima Daiichi<br />

Nuclear Power Plant in Fukushima prefecture. This study was<br />

undertaken using data from annual health checkups of persons<br />

40-90 years old from 2008 to 2010. Follow-up examinations<br />

were conducted from June 2011 to March 2013, with a<br />

mean follow-up of 1.6 years. Changes in liver function before<br />

and after the disaster were compared among evacuees and<br />

non-evacuees. Liver dysfunction was defined as aspartate aminotransferase<br />

≥40 IU/L, alanine aminotransferase ≥40 IU/L, or<br />

γ-glutamyl transpeptidase ≥50 IU/L. Subjects were divided into<br />

three alcohol intake groups: 1) non-drinkers, defined as drinking<br />

no alcohol; 2) light drinkers, defined as


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 485A<br />

preventive liver care measures for IBD patients<br />

Disclosures:<br />

K. Gautham Reddy - Advisory Committees or Review Panels: AASLD Transplant<br />

Hepatology Pilot Steering Committee, ACG Training Committee, Program Director’s<br />

Caucus Steering Committee, Gilead, Inercept, Bristol-Myers Squibb; Grant/<br />

Research Support: Intercept, Ocera, Merck, Lumena<br />

Helen S. Te - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Abbvie, Conatus, BMS<br />

Nancy Reau - Advisory Committees or Review Panels: Jannsen, Merck, AbbVie,<br />

Intercept, Salix, BMS, Jannsen; Grant/Research Support: Merck, Gilead, BMS,<br />

AbbVie, Jannsen, BI<br />

The following authors have nothing to disclose: John N. Gaetano, Andrew I.<br />

Aronsohn<br />

Disclosures:<br />

Robert E. Smith - Speaking and Teaching: Gilead, Salix, Abbvie, Janssen<br />

The following authors have nothing to disclose: Chuan L. Miao, Sara West, Korta<br />

Yuasa, Michael Komar, Stacy Prall<br />

553<br />

Disparities in the Quality of Death and Resource Utilization<br />

in End-Stage Liver Disease<br />

John N. Gaetano, K. Gautham Reddy, Andrew I. Aronsohn, Helen<br />

S. Te, Nancy Reau; Department of Medicine, University of Chicago<br />

Hospitals, Chicago, IL<br />

Background: Various <strong>studies</strong> have shown increased utilization<br />

of life-sustaining measures as well as increased end-of-life costs<br />

in minorities. It is unknown if this disparity exists in the end-of life<br />

care among patients with end-stage liver disease. Methods: The<br />

Nationwide Inpatient Sample from 2012 was used to capture<br />

patients with cirrhosis and a manifestation of hepatic decompensation<br />

(variceal bleeding, hepatorenal syndrome, ascites,<br />

spontaneous bacterial peritonitis, or portosystemic encephalopathy),<br />

above the age of 39, who died in the hospital. The<br />

primary predictor variable was race, while the primary outcome<br />

measure was the use of life sustaining measures (cardiopulmonary<br />

resuscitation, mechanical ventilation, reintubation,<br />

hemodialysis, total parenteral nutrition, tracheostomy, and use<br />

of vasopressors). Patient and hospital demographic information<br />

as well as patient comorbidities were controlled for in multivariate<br />

analysis. Results: There were 6,600 patients with cirrhosis<br />

of the liver, age 40 or greater, who died in the hospital. Of<br />

these, 4,178 had at least one form of hepatic decompensation<br />

and were then included in to our analysis. Comorbidity<br />

indexes across race were not significantly different. On univariate<br />

analysis, black and Hispanic race had significantly<br />

more frequent use of life sustaining measures (69% and 65%,<br />

respectively) compared to white race (58%, p


486A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Vincent Di Martino - Board Membership: Merk schering Plough France, Merk<br />

schering Plough France, Merk schering Plough France, Merk schering Plough<br />

France; Speaking and Teaching: Roche France, Gilead France, Roche France,<br />

Gilead France, Roche France, Gilead France, Roche France, Gilead France<br />

The following authors have nothing to disclose: Armand Garioud, Jean françois<br />

D. Cadranel, Jean-Baptiste Nousbaum, Allaoua Smaïl, Hortensia Lison, Thierry<br />

Thevenot, Mourad Medmoun, Djamel Belloula, Firouzeh Kazerouni, Khaled<br />

Hadj-Nacer, Alain Cazier<br />

Table 1. Predictive parameters, risk score and estimated probability<br />

of survival.<br />

555<br />

Predicting survival in liver abscess: a risk score based<br />

on 8,423 patients hospitalized with liver abscess, Thailand.<br />

Kittiyod Poovorawan 1 , Wirichada Pan-ngum 2 , Ngamphol<br />

Soonthornworasiri 2 , Chatporn Kittitrakul 1 , Polrat Wilairatana 1 ,<br />

Sombat Treeprasertsuk 3 , Bubpha Kitsahawong 4 , Kamthorn Phaosawasdi<br />

4 ; 1 Department of Clinical Tropical Medicine, Faculty<br />

of Tropical Medicine, Mahidol University, Bangkok, Thailand;<br />

2 Department of Tropical Hygiene, Faculty of Tropical Medicine,<br />

Mahidol University, Bangkok, Thailand; 3 Department of Medicine,<br />

Faculty of Medicine, Chulalongkorn University, Bangkok, Thailand;<br />

4 Vichaiyut hospital and Medical Center, Bangkok, Thailand<br />

We aimed to establish a risk score for predicting 60-day survival<br />

in patients hospitalized with liver abscess, using Nationwide<br />

Hospital Admission Data from 2009 to 2013 provided<br />

by the National Health Security Office (NHSO), Thailand. All<br />

patients with a primary diagnosis of pyogenic liver abscess and<br />

amoebic liver abscess (ICD10-K750, A064) were included.<br />

Baseline characteristics, comorbidities, hospital course, and<br />

survival were analyzed, and all-causes mortality data were<br />

retrieved from national death registration. A total of 11,296<br />

admissions comprised of 8,423 patients from 844 hospitals<br />

across Thailand were eligible for analysis. The mean age was<br />

52±17 years and 66.1% of patients were male. Median length<br />

of hospitalization was 8 days (IQR 4 to 13 days), and the<br />

mean cost of hospitalization was 846±1,574 USD. Overall<br />

in-hospital mortality rate was 2.8%, and overall long-term survival<br />

was significantly higher among patients with amoebic<br />

liver abscess compared to pyogenic liver abscess (HR 0.54;<br />

95% CI, 0.42-0.68; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 487A<br />

Disclosures:<br />

The following authors have nothing to disclose: Michelle T. Jesse, Reema Habra,<br />

Steven Mekaru, Eric Goldstein, Anne Eshelman, Marwan S. Abouljoud<br />

557<br />

Palliative Care Services are Underutilized in Liver Transplant<br />

Candidates<br />

Priya Kathpalia 1 , Alexander Smith 2 , Jennifer C. Lai 1 ; 1 Division of<br />

Gastroenterology and Hepatology, University of California, San<br />

Francisco, San Francisco, CA; 2 Division of Geriatrics, University of<br />

California, San Francisco, San Francisco, CA<br />

Background: Patients with end-stage liver disease (ESLD)<br />

experience high symptom burden and 50% risk of death at<br />

5 years from progressive liver disease without liver transplant<br />

(LT). These patients represent a population with high needs<br />

for palliative care. Utilization practices of palliative care of LT<br />

candidates are not well characterized. Methods: All adult LT<br />

candidates who died or were delisted from the wait-list at a<br />

single high volume LT center from 2013-14 were evaluated.<br />

Excluded were those who were delisted for noncompliance<br />

or were transplanted at another center. Reasons for palliative<br />

care consultation were obtained from review of electronic<br />

health records. Logistic regression assessed factors associated<br />

with palliative care consultation in these patients. Results: Of<br />

683 patients listed for LT in 2013-14, 107 (16%) ultimately<br />

died (n=62) or were delisted for being too sick for LT (n=45):<br />

median age was 58 years, 34% were female, 53% were white,<br />

61% had Child Pugh Class C cirrhosis, and 52% had hepatocellular<br />

carcinoma. Among these 107 patients who died/were<br />

delisted, 17% (n=18) received a palliative care consult, 89%<br />

(16/18) of which occurred in the inpatient setting. The median<br />

(interquartile range [IQR]) number of days from palliative care<br />

consultation to death was 4 (1-11 days), with 50% of the consultations<br />

occurring within 3 days of death. Reasons for palliative<br />

care consultation were: assistance with transition to comfort<br />

care (78%), goals of care without transition to comfort care<br />

(11%) and symptom management (11%). Even among the 26<br />

patients delisted for advanced HCC, only 12% were referred<br />

to palliative care. In univariable analysis, age (OR 0.92; 95%<br />

CI 0.87-0.98; p


488A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

analyzed a large commercial insurance provider database to<br />

investigate the epidemiology, the pattern of treatment, and<br />

survival in patients with HCC. Methods: This is a retrospective<br />

matched case-control analysis estimating the epidemiology of<br />

HCC as well as clinical characteristics, treatment patterns and<br />

survival in enrollees with HCC in Health Care Services Corporation<br />

(HCSC), a commercial insurer providing health-care coverage<br />

in 5 states, from September 2008-August 2013. HCC<br />

was defined as ICD-9 code 155.0 after the first six months<br />

of enrollment. Cirrhosis was defined by ICD-9 codes 571.2,<br />

571.5, and 571.6. Treatment was grouped as following: surgical<br />

resection (CPT: 47120, 47122, 47125, 47130, 47712),<br />

ablation (CPT: s 47370, 47371, 47379, 47380, 47381,<br />

47382), TACE (CPT: 36247, 36248, 36246, 75726, 75774,<br />

37243, with code 96420) TARE (CPT: 36247, 36248, 36246,<br />

75726, 75774, 37243, with code 79445). Sorafenib use<br />

was defined as:ICD9 155.0 and a claim for sorafenib. Liver<br />

transplant was defined as ICD-9: v42.7 AND CPT 47135 or<br />

47136. Results: During the study period, there were 3465<br />

with an ICD-9 code 155.0 among 3,478,093 enrollees yielding<br />

a prevalence of HCC of 0.09%. Among patients with cirrhosis,<br />

7.95% had a diagnosis of HCC. When compared to<br />

those with cirrhosis without HCC, those with HCC were older<br />

(58.1 vs 55.5, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 489A<br />

≥1.45 were counseled by the HIV clinic and referred to the<br />

Division of Gastroenterology and Hepatology for liver liver<br />

fibrosis assessment and evaluation for HCV therapy, as reimbursement<br />

of second-generation direct-acting antivirals is<br />

restricted to patients with ADVFIB (liver stiffness >9.5kPa/<br />

METAVIR F3/F4) in Austria. Results: Among 1348 HIV-positive<br />

patients, 31%(414/1348) of patients were HCV-antibody<br />

positive and 17%(229/1348) patients had detectable serum<br />

HCV-RNA. The majority of HIV/HCV (62%) had HCV-genotype<br />

1 (subtype 1a:75%/1b:25%), while HCV-genotypes 2,<br />

3, and 4 were observed in 1%, 28%, and 9% of patients.<br />

One hundred thirteen HIV/HCV with a FIB-4 index ≥1.45<br />

were counseled by the HIV clinic. Among these, 58%(66/113)<br />

underwent liver fibrosis assessment either by transient elastography<br />

(TE) or biopsy. ADVFIB was confirmed in 50% (33/66)<br />

and IFN-free treatments were initiated in 17 of these patients<br />

(SOF/RBV:12%[2/17]; SOF/DCV:76%[13/17]; SOF/<br />

LDV:12%[2/17]). Importantly, among patients with a FIB-4<br />

index ≥1.45 in whom liver fibrosis was not assessed by either<br />

TE or biopsy (42%[47/113]), 13 patients had a FIB-4 index<br />

>3.45, which is diagnostic for ADVFIB. None of these patients<br />

received HCV therapy. Another 9 patients with a FIB-4 index<br />

576,000<br />

patients receiving AMA testing between 1/10 - 12/13 in<br />

a large US commercial laboratory database. Patients were<br />

considered AMA+ if antibody titre > 1:20 or M2 antibody ><br />

25.0 U. We examined patient demographics, site of testing<br />

& testing history. Results: 16,492 patients tested AMA+ and<br />

had at least 1 alkaline phosphatase (ALP) test over the 4 year<br />

period; 6,107 had an ALP >120 U/l that period and could be<br />

classified as probable PBC. 53% of initial AMA+ tests were<br />

conducted in a hospital setting. The specialty of the ordering<br />

outpatient physician was gastroenterologists (18%), primary<br />

care (14% including NP/PAs) and Rheumatologists (2%). A significant<br />

minority of 1st AMA+ tests were male (19%). The age<br />

at 1st AMA+ tests: <br />

65 (40%). 55% of patients had >3 ALP tests before 1st positive<br />

AMA; 67% had at least one ALP > 115 U/l before 1st AMA+<br />

test, and 40% had >2 tests with ALP > 115. Conclusion: The<br />

63% of AMA+ patients without elevated ALP (ie, probably<br />

PBC) is similar to that reported in AASLD guidelines. Most PBC<br />

patients’ 1st AMA+ test was in a hospital setting, suggesting<br />

that they are being diagnosed as the result of standard screening<br />

through other emergent conditions, not as part of regular<br />

care. Given that early stage PBC is asymptomatic, this may<br />

not be unexpected, except that 67% of patients had > 1 and<br />

40% > 2 ALP test results > 115 U/l before 1st AMA+ test.<br />

This suggests that outpatient physicians may not be following<br />

up on elevated ALP with AMA testing. The distribution of age<br />

is similar to prevalence <strong>studies</strong>, but the proportion of men testing<br />

AMA+ is more than double what was expected. Other<br />

<strong>studies</strong> suggest that men are diagnosed at a later PBC stage<br />

than women, and have a higher mortality rate, thus a higher<br />

proportion of incident men would ultimately produce lower<br />

prevalence. Research on healthcare resource use shows that<br />

in the US, men are less likely to use primary care, and instead<br />

receive care through emergency rooms, which may explain the<br />

high rate of hospital testing.<br />

Disclosures:<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

Herbert Swanson - Management Position: Intercept<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

The following authors have nothing to disclose: Keith D. Lindor<br />

563<br />

Worse Outcomes Among Uninsured Patients with Cirrhosis<br />

– Towards Understanding the Obstacles to Hepatology<br />

Care<br />

Daniela Ladner 1 , Kathryn Skibba 1 , Kathryn Jackson 1 , Niloufar<br />

Safaeinili 1 , David S. Goldberg 2 , Lisa B. VanWagner 1 , Satyender<br />

Goel 1 , Abel Kho 1 , Amna Daud 1 , Nazanin Salehitezangi 1 , Anton<br />

I. Skaro 1 ; 1 Northwestern University, Chicago, IL; 2 University of<br />

Pennsylvania, Philadelphia, PA<br />

Background: Cirrhosis of the liver is a leading cause of morbidity<br />

and mortality in the US and worldwide. Advanced hepatology<br />

care and transplantation can improve survival and quality<br />

of life. However, discrete obstacles to access to appropriate<br />

liver-specific care exist. Due to an absence of epidemiologic<br />

data collected systematically on the cirrhotic patient population<br />

the nature and magnitude of the problem remains poorly characterized.<br />

Longitudinal measurement of MELD is an adequate<br />

surrogate of access to care. Therefore, we examined outcomes


490A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of patients with cirrhosis identified within the greater Metropolitan<br />

Chicago area by query of the HealthLNK database. Methods:<br />

HealthLNK captures over 2 million patients treated within<br />

6 diverse healthcare integrated delivery networks in the greater<br />

Chicago area (Northwestern Medicine, Univ of Chicago, Rush<br />

Univ MC, University of Illinois of Chicago MC, and Loyola Univ<br />

MC, and Cook County Health Systems) between 2006-2012.<br />

Cirrhosis was defined by ICD-9 codes 571.2 OR 571.5 OR<br />

571.6 OR Fib-4 score > 3.25. Demographics, and co-morbidities<br />

assessed by ICD-9 codes, CPT codes, medications and laboratory<br />

values were collected. Cirrhotic patients were stratified<br />

into groups according to whether or not the MELD score was<br />

measured. Comparisons by group were tested using t-tests or<br />

chi-squared tests for continuous and non-continuous variables<br />

respectively. Results: 24,185 cirrhotic patients were identified<br />

and 11,481 (47%) had INR, bilirubin, and creatinine biochemistries<br />

drawn at one or more points in time. There were no<br />

differences in age or sex among those patients with no labs<br />

available to calculate a MELD score versus those with labs to<br />

calculate a MELD score (p2) were assimilated into<br />

a screening tool. The tool was retrospectively applied to all<br />

patients admitted to the Bristol Royal Infirmary with a diagnosis<br />

of cirrhosis over 90 consecutive days from 1st July 2013<br />

(n=54). Mortality 1 year post initial admission was calculated.<br />

Results Of 54 index cases 7 were not analysed due to incompleteness<br />

of data or loss to follow up. On the basis of optimum<br />

sensitivity and specificity for death at 1 year a score of 3 or<br />

more was considered a threshold for identifying likely poor<br />

prognosis (table). Conclusions The tool has been trialed and<br />

audited over the past year. Patients who screen positive (≥<br />

3 criteria) are discussed at a weekly hepatology multidisciplinary<br />

meeting (MDM). Assuming MDM agreement consultant<br />

led patient discussion, a poor prognosis letter to the GP, and a<br />

palliative medicine referral are triggered. Communication skills<br />

training has been delivered to consultant and junior staff by<br />

the palliative medicine team. As deaths from liver disease continue<br />

to increase, identifying patients who stand to benefit from<br />

advanced care planning and timely palliative care intervention<br />

will become increasingly important.<br />

Accuracy of tool at varying ‘thresholds’ for predicting mortality at<br />

1 year<br />

Disclosures:<br />

The following authors have nothing to disclose: Benjamin E. Hudson, Kelly Ameneshoa,<br />

Peter Collins, Andrew J. Portal, Fiona H. Gordon, Julia Verne, Anne<br />

McCune<br />

565<br />

Distance to Transplant Center and Waiting List Outcomes<br />

Jordan Voss, Adam McCann, Sean C. Kumer, Timothy Schmitt,<br />

Richard Gilroy; Kansas University Medical Center, Kansas City, KS<br />

Objective: To analyze the impact of a patient’s distance from<br />

a transplant center on liver transplant waiting list outcomes.<br />

Methods: All patients waitlisted at a single transplant center<br />

from January, 2010 to May, 2014 were evaluated for waitlist<br />

outcome, distance to transplant center and model for end-stage<br />

liver disease (MELD) score at time of listing, transplant, and<br />

death. Distance to transplant center was evaluated both as a<br />

continuous variable and categorically for distances of 25, 50,<br />

100, and 200 miles. Logistic regression was used to evaluate<br />

the effect of distance on likelihood of death on the waiting list.<br />

Results: 617 patients were waitlisted during the study period. In<br />

univariate analysis distance was very weakly associated with a<br />

lower MELD at both listing and death: R=-0.086, p=0.033 and<br />

R=-0.181, p=0.036, respectively. Logistic regression revealed<br />

that distance does not predict death on the waiting list (OR<br />

1.000, p=0.851). Median test and Mann-Whitney U Test indicated<br />

that MELD scores at listing, transplant, and death are<br />

not significantly different among patients living within 25, 50,<br />

100, and 200 miles of the transplant center compared to those<br />

living outside the specified distances. Conclusions: A patient’s<br />

distance from the liver transplant center does not appear to<br />

impact MELD score at the time of listing and once listed does<br />

not impact MELD at transplant/death or likelihood of death on<br />

the waiting list. How socioeconomic and racial/ethnic variables<br />

impact this warrants further consideration given the differ-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 491A<br />

ence in these variables between rural and urban populations,<br />

and is being evaluated in subsequent analyses.<br />

Disclosures:<br />

Richard Gilroy - Speaking and Teaching: Salix, NPS, Gilead, AbbVie, Novartis<br />

The following authors have nothing to disclose: Jordan Voss, Adam McCann,<br />

Sean C. Kumer, Timothy Schmitt<br />

566<br />

Deleterious effects of cirrhosis on outcomes after firearm<br />

injury<br />

Sentia Iriana 1 , Elliot B. Tapper 2 , Vinay Sundaram 1 ; 1 Medicine,<br />

Cedars-Sinai Medical Center, Los Angeles, CA; 2 Medicine, Beth<br />

Israel Deaconess Medical Center, Boston, MA<br />

Aims: Cirrhotic patients have poor outcomes from traumatic<br />

injury. The impact of firearm injury in this population is unexplored.<br />

The aim of this study was to assess the prevalence and<br />

outcomes related to firearm injury related hospitalization in cirrhosis<br />

patients. Methods: We analyzed the Nationwide Inpatient<br />

Sample, 2009-2011. Patients were identified by having a<br />

primary or secondary discharge diagnosis associated with firearm<br />

injury. We utilized multivariable logistic regression models<br />

for the outcome of inpatient mortality. Deyo modification of the<br />

Charlson index was used to account for non-hepatic comorbidity.<br />

Results: A total of 348,823 discharge records were studied,<br />

of which 317,592 had cirrhosis alone, 26,044 had firearm<br />

injury without cirrhosis, and 2,730 had cirrhosis with firearm<br />

injury. Among cirrhotics with firearm injury, 910 (33.3%) were<br />

accidental and 1820 (66.7%) were assault related. Cirrhotics<br />

with firearm injury were younger compared to cirrhotics without<br />

firearm injury (54.9 vs 58.7 years, p


492A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

lyzed using Chi2 test for trend and linear regression analysis.<br />

Results Among 34,308 HIV-infected patients enrolled between<br />

2000 and 2012, 5,562 were HCV coinfected. HCV prevalence<br />

declined from 28% to 15% during the study period<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 493A<br />

outcomes in a multi-center cirrhotic cohort. Methods: NACSELD<br />

(North American Consortium for End-Stage Liver Disease)<br />

enrolls hospitalized cirrhotics from 16 centers. Data collected<br />

are admission/cirrhosis details, inpatient details & 30-day mortality.<br />

Details & outcomes of pts given (Alb) /not given albumin<br />

(no Alb) and regional variations were compared. Within<br />

Alb, evidence-based indications (AKI, SBP, paracentesis) were<br />

compared to others. Results: 988 cirrhotics (64% men, MELD<br />

19, 43%alcohol) were included. 597(60%) pts received a<br />

280±267 gm of albumin. Alb pts had worse cirrhosis severity<br />

but similar demographics as non-alb(Table).Albumin use: The<br />

majority(76%) was evidence-based (47%AKI, 34%paracentesis<br />

& 14%SBP). Remaining 24% was for hyponatremia (12%),<br />

non-SBP infection (6%) & anasarca (6%). This resulted in AKI<br />

resolution in 28%, reduced anasarca in 22%, improved Na in<br />

12% & allowed paracentesis in 30% of pts. 4% had adverse<br />

events (3% fluid overload & 1% pulmonary edema) Regional<br />

variation: 4 regions: 3 US (1: North, 2: South, 3: Southwest)<br />

& Canada (region 4) were studied. Region 3 had the highest<br />

albumin use (69%) with lowest evidence-based use (57%) compared<br />

to Region 2 that used the least(45%,p


494A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Luca Fabris, Stefano Okolicsanyi,<br />

Matteo Rota, Paolo Cortesi, Luca S Belli, Stefano Fagiuoli, Luciana Scalone,<br />

Giancarlo Cesana, Mario Strazzabosco<br />

572<br />

Knowledge about hepatitis B transmission risks among<br />

health professionals in Tanzania<br />

Jose Debes 1 , Johnstone Kayandabila 2 , Hope Pogemiller 1 ; 1 University<br />

of MInnesota, Minneapolis, MN; 2 Arusha Lutheran Medical<br />

Centre, Arusha, United Republic of Tanzania<br />

Introduction: It is well known that hospital workers are at<br />

increased risk for contracting hepatitis B virus (HBV). This risk<br />

it is increased in settings of high HBV seroprevalence, such as<br />

sub-Saharan Africa. Despite this, it is unclear whether healthcare<br />

providers in developing countries are aware of the transmission<br />

risks and adhere to universal precaution strategies.<br />

In this study we aimed to evaluate the knowledge and understanding<br />

of HBV among health-care workers in hospitals in<br />

Tanzania. Methods: We provided a HBV survey to staff in 2<br />

hospitals in northern Tanzania. The survey consisted of nine<br />

multiple-choice questions that inquired about understanding<br />

of HBV serostatus, vaccination and risks of transmission of<br />

HBV. The survey was written in English and Swahili (the local<br />

language) and distributed among medical and non-medical<br />

staff (laboratory technicians, students, etc). Multivariate analyses<br />

were performed using Fishers exact test and Chi-square.<br />

Results: We received voluntary participation from 114 subjects<br />

in our survey. The mean age of the participants was 33<br />

y/o, 67% were females, and 61% responded to the survey in<br />

English. Ninety one percent of subjects had no knowledge of<br />

their HBV surface antigen status, and 89% indicated they never<br />

received a vaccine for HBV, with lack of knowledge about<br />

the vaccine being the most common reason (34%). Seventy<br />

percent of participants new about HBV complications and 60%<br />

responded correctly inquires about transmission routes. There<br />

was a significant difference in knowledge of HBV serostatus<br />

and vaccination between participants with a medical background<br />

(consultants, interns, etc) and others, p=0.01 and<br />

p-0.001 respectively. However, only 33% of consultants knew<br />

about their HBV serostatus or were vaccinated for it. There<br />

was no significant difference between knowing HBV transmission<br />

route or in-hospital risk and whether the survey was<br />

answered in English or local language, or hospital position of<br />

the responder. Conclusions: Our study shows a surprisingly low<br />

knowledge of HBV serostauts and vaccination status among<br />

hospital workers in Tanzania. Studies are needed in other parts<br />

of sub-Saharan Africa to further understand this gap in knowledge,<br />

and HBV awareness programs should be promoted.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jose Debes, Johnstone Kayandabila,<br />

Hope Pogemiller<br />

573<br />

Hepatitis C Virus Sustained Virologic Response Rates<br />

among Patients Treated in Primary Care Settings in New<br />

York State<br />

Colleen Flanigan, Jason Pendergast; New York State Department<br />

of Health, Albany, NY<br />

Background Hepatitis C virus (HCV) treatment has been rapidly<br />

changing over the last few years. With the arrival of direct<br />

acting anti-viral therapies, HCV treatment is now highly effective.<br />

As a result of these more effective treatments, the demand<br />

for HCV treatment is increasing, adding strain to an already<br />

limited number of specialists available to treat HCV. The newer<br />

treatments are less complex, more tolerable and treatment<br />

durations can be as short as eight weeks, thus, allowing for<br />

primary care providers (PCPs) to treat HCV. To expand access<br />

to HCV care and treatment in New York State (NYS), the NYS<br />

Department of Health provides funding to 13 primary care sites<br />

to integrate HCV care and treatment. These primary care sites<br />

include: community health centers, hospital-based clinics, a<br />

methadone maintenance treatment program and a residential<br />

drug treatment program. At each of these settings, a PCP and<br />

a multidisciplinary team manages and treats persons infected<br />

with HCV. Each PCP is an experienced HCV provider with the<br />

knowledge and skills to safely and effectively manage and<br />

treat HCV. Each PCP has an agreement with a liver specialist<br />

for consultation. The purpose of this study was to evaluate<br />

HCV treatment sustained virologic response (SVR) rates<br />

among patients receiving HCV treatment in a primary care<br />

setting. Methods This analysis looks at HCV SVR rates among<br />

HCV monoinfected and HIV/HCV coinfected patients enrolled<br />

at 13 primary care sites across NYS. All primary care sites<br />

report client level data using the AIDS Institute Reporting System<br />

(AIRS). AIRS is a comprehensive client and service/encounter<br />

reporting application. AIRS also collects HCV treatment type<br />

and treatment outcome (SVR). Results From October 1, 2010<br />

through April 30, 2015, 2,447 HCV infected patients were<br />

served by the 13 primary care sites. Of these patients, 816<br />

(33.3%) initiated HCV treatment and of these patients 743<br />

(91.1%) had final treatment outcomes available for analysis.<br />

The overall SVR rate among these patients was 61.5%; 70.1%<br />

among HCV monoinfected and 50.8% among HIV/HCV coinfected<br />

patients. SVR rates ranged from 49.3% among those<br />

treated with the combination pegylated interferon and ribavirin<br />

to as high as 96.6% among those treated with ledipasvir-sofosbuvir.<br />

Conclusions These findings illustrate that HCV treatment<br />

SVR rates are just as good in primary care settings as they are<br />

in specialty settings. Therefore, integrating HCV treatment into<br />

primary care can help increase capacity for HCV treatment<br />

and allow more people to be cured of HCV.<br />

Disclosures:<br />

The following authors have nothing to disclose: Colleen Flanigan, Jason Pendergast<br />

574<br />

External Validity of Trials in Non-Alcoholic Fatty Liver<br />

Disease: Systematic Review of Randomised Controlled<br />

Trials and Non-Randomised <strong>studies</strong><br />

Richard Parker 1 , James Hodson 2 , Ian A. Rowe 1 ; 1 Centre for Liver<br />

Research, University of Birmingham, Birmingham, United Kingdom;<br />

2 University Hospitals Birmingham NHS Foundation Trust, Birmingham,<br />

United Kingdom<br />

Introduction Trials with good external validity allow clinicians<br />

and patients to have confidence in new treatments. Individuals<br />

in clinical trials are often highly selected, in part to remove<br />

confounding factors, but this can introduce bias and limit the<br />

external validity of a trial. In this study, we surveyed observational<br />

<strong>studies</strong> and randomised controlled trials (RCT) to gauge<br />

the external validity of trials in non-alcoholic fatty liver disease<br />

(NAFLD). Methods Structured literature searches were undertaken<br />

for observational <strong>studies</strong> (ObS) and RCT in NAFLD. Data<br />

were extracted from included <strong>studies</strong>. Participant characteristics<br />

were compared between observational <strong>studies</strong> and RCT, and<br />

between subsets of observational <strong>studies</strong> describing with biopsy-proven<br />

non-alcoholic steatohepatitis (NASH) or advanced<br />

fibrosis (>F2). For age and BMI, weighted means were calculated<br />

and compared with t-tests, for prevalence of Diabetes mellitus<br />

(DM) and gender, Chi-squared was used for comparisons.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 495A<br />

Results 112 <strong>studies</strong> were included: 47 RCT and 65 ObS. Sixteen<br />

observational <strong>studies</strong> described characteristics of patients<br />

with NASH and 31 described characteristics of patients with<br />

advanced fibrosis. RCT participants differed from individuals in<br />

ObS with regard to age, BMI, prevalence of DM, and gender<br />

(p


496A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

disseminate knowledge on new HCV treatments are needed to<br />

decrease the burden of HCV infection.<br />

577<br />

Interventions to optimize retention in the chronic viral<br />

hepatitis screening, care and treatment cascade: a systematic<br />

review<br />

Kali Zhou 4 , Nick Walsh 1 , Thomas Fitzpatrick 2 , Ji Young Kim 1 ,<br />

Ying-Ru Lo 1 , Joseph D. Tucker 3 ; 1 World Health Organization Western<br />

Pacific Region, Manila, Philippines; 2 University of Washington,<br />

Seattle, WA; 3 University of North Carolina-Chapel Hill, Chapel<br />

Hill, NC; 4 Gastroenterology, UCSF, San Francisco, CA<br />

BACKGROUND: Global mortality burden of viral hepatitis is<br />

estimated at 1.45 million and now surpasses that of HIV. Recent<br />

advances in hepatitis B (HBV) and hepatitis C (HCV) therapeutics<br />

provide the opportunity to effectively treat HBV and cure<br />

chronic HCV, thereby reducing hepatitis related mortality. To<br />

support countries in the delivery of hepatitis programs, we carried<br />

out a systematic review of interventions that may increase<br />

uptake and retention in the hepatitis screening, care and treatment<br />

service cascade. METHODS: We searched Pubmed,<br />

EMBASE, the WHO library, Clinicaltrials.gov, International<br />

Clinical Trials Registry Platform, Psychinfo and Cinahl. Interventional<br />

<strong>studies</strong> with a comparator arm targeting an aspect of the<br />

chronic hepatitis treatment cascade were included. The study<br />

was registered in PROSPERO (42014015094) and carried out<br />

according to PRISMA guidelines. Results from <strong>studies</strong> with similar<br />

outcomes were pooled using RevMan. RESULTS: We identified<br />

7,569 citations and included 56 in the review. All <strong>studies</strong><br />

were from high-income settings. Self-reported HBV testing rates<br />

were higher among recipients of community based HBV-specific<br />

educational interventions compared to non-specific health<br />

interventions (RR=3.97, CI95 3.02-5.22). For HCV screening,<br />

facility-based interventions targeting providers (n=4) (HCV-specific<br />

education, electronic health record prompts) increased<br />

HCV testing rates (RR=3.44, CI95 3.24-3.65) compared to<br />

standard of care; interventions targeting patients (HCV-specific<br />

patient education) (n=4) also increased testing rates (RR=1.43,<br />

CI95 1.29-1.59) compared to standard of care. Interventional<br />

<strong>studies</strong> (n=11) to boost adherence to HCV treatment (patient<br />

support program, adherence literacy, integrated care program)<br />

increased treatment completion rates (RR=1.10, CI95 1.04-<br />

1.16)) compared to standard of care. Finally, when adherence<br />

was measured by sustained virologic response (SVR),<br />

adherence focused interventions (n=17) increased the likelihood<br />

of achieving SVR (RR=1.20, CI95 1.13-1.28) compared<br />

with standard of care. CONCLUSION: There was a lack of<br />

data on community-based interventions to increase HCV testing<br />

uptake, and no data on interventions to improve hepatitis<br />

B treatment uptake and adherence. These areas require further<br />

research. Efforts are needed to translate these findings to<br />

increase chronic viral hepatitis screening uptake and optimize<br />

country programmatic efforts in treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kali Zhou, Nick Walsh, Thomas<br />

Fitzpatrick, Ji Young Kim, Ying-Ru Lo, Joseph D. Tucker<br />

Disclosures:<br />

Anna S. Lok - Advisory Committees or Review Panels: Gilead, MYR, Tekmira;<br />

Consulting: GSK, Merck; Grant/Research Support: AbbVie, BMS, Gilead, Idenix<br />

The following authors have nothing to disclose: Mary Thomson, Monica Konerman<br />

578<br />

Upper Limit Normal of Alanine Aminotransferase for the<br />

Indian Population<br />

Jack Masur 1 , Shyam Kottilil 1 , Eleanor Wilson 2 , Samir R. Shah 3 ;<br />

1 Institute of Human Virology, University of Maryland School of<br />

Medicine, Baltimore, MD; 2 Critical Care Medicine Department,<br />

National Institutes of Health, Bethesda, MD; 3 Global Hospitals,<br />

Mumbai, India<br />

Introduction AASLD recently lowered the Upper Limit Normal<br />

(ULN) of alanine aminotransferase (ALT) level to 30 U/L in<br />

males and 19 U/L in females. Studies have found that relying<br />

on pre-existing normal ranges for ALT measurement in patients<br />

with chronic hepatitis C virus (HCV) infection was insensitive<br />

at identifying patients with minimal to mild necroinflammatory<br />

activity. ALT values have been shown to differ by ethnicity;<br />

as well as by metabolic factors such as BMI, glucose, cholesterol,<br />

and hyperlipidemia; and, by infection from chronic<br />

hepatitis viruses. The existing ULN level in India is 45 U/L,<br />

irrespective of sex. India currently has an HBV prevalence of<br />

3.7% and HCV prevalence of 1%, and many infected persons<br />

are unaware of their chronic carrier status. Additionally,<br />

non-alcoholic fatty liver disease (NAFLD) is an epidemic in<br />

India. Increased sensitivity of testing might allow for prevention<br />

and earlier intervention. The purpose of this study was to test<br />

whether the currently defined ULN of ALT thresholds in India<br />

underestimate the number of patients at risk for chronic liver<br />

disease in the country. Methods We conducted a retrospective<br />

study of patient laboratory data and vital sign measurements<br />

collected from the Metropolis Laboratory in Mumbai, India over<br />

a two-year period. Patients were analyzed based on preset<br />

health parameters that were used to determine inclusion or<br />

exclusion from the “control” cohort. The parameters required<br />

for inclusion in the “control” cohort included: BMI ≤23 kg/<br />

m 2 ; normal blood pressure of 120/80 mm Hg; total cholesterol<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 497A<br />

Disclosures:<br />

The following authors have nothing to disclose: Jack Masur, Shyam Kottilil, Eleanor<br />

Wilson, Samir R. Shah<br />

579<br />

Successful Implementation of a Community-based<br />

Patient Navigation Program to Increase Screening and<br />

Linkage-to-care in High-Risk Patients with Chronic Hepatitis<br />

B Infection<br />

Jennifer M. Newton 1 , Matt Johnson 2 , Sharon Song 2 , Helen S. Te 1 ,<br />

Edwin Chandrasekar 2 , Karen E. Kim 1 ; 1 Medicine, University of<br />

Chicago Medicine, Chicago, IL; 2 Asian Health Coalition, Chicago,<br />

IL<br />

Background: Chronic hepatitis B virus infection (CHB) is estimated<br />

to affect up to 2.2 million people in the United States,<br />

and disproportionately affects foreign-born Asian and African<br />

immigrants. CHB carries a high morbidity and mortality<br />

due to undertreatment, which largely reflects lack of awareness<br />

of infection and poor linkage to care. Purpose: In this<br />

intervention study, we have implemented a community-based<br />

patient navigation program. Utilizing community-health workers<br />

(CHWs) and patient navigators (PNs), we aim to increase<br />

patient awareness of infection and improve linkage to care<br />

for foreign-born Asian, Pacific Islander and African patients<br />

with CHB. Methods: While CHWs provide culturally competent<br />

education, PNs are essential for care coordination. To improve<br />

community linkage to care, we developed a co-curriculum to<br />

enhance care coordination and community linkage for hepatitis<br />

B surface antigen (HBsAg) positive patients. To strengthen<br />

their partnerships, CHWs and HPNs jointly attend site visits<br />

and training sessions for cultural competency, CHB education,<br />

navigating the healthcare system and community awareness.<br />

Results: In the first 6 months of the intervention, 46 (4.7%) of the<br />

974 individuals screened were HBsAg positive. Asian (674,<br />

69.2%) and Black/African (224, 23%) patients accounted for<br />

the largest ethnic groups screened, while Iraq (228, 23.4%)<br />

and Myanmar (147, 15.1%) represent the largest screenings<br />

by country of origin. 247 patients received all 3 HBV serologic<br />

tests (surface antigen, surface antibody, and core antibody).<br />

Of these patients, 111 (44.9%) were previously vaccinated,<br />

60 (24.3%) were susceptible to infection, and 40 (16.2%) had<br />

resolved infection. 891 (91.5%) of the 974 patients screened<br />

received test results and counseling, including all 46 (100%)<br />

patients with active infection. Among patients with active<br />

HBV infection, 37 (80.4%) have been linked to care, with 31<br />

(67.4%) seeking care through specialty clinics and 6 (13%)<br />

through primary care. 3 (6.5%) patients moved out of state,<br />

and 6 (13%) have been lost to follow-up. Conclusions: To our<br />

knowledge, this is the first community-based hepatitis B patient<br />

navigation program to be instituted in the country. This intervention<br />

has been successful in linking high-risk Asian and African<br />

patients to healthcare for CHB, and is a promising model that<br />

may be implemented nationally to decrease the unnecessary<br />

morbidity and mortality associated with CHB.<br />

Disclosures:<br />

Helen S. Te - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Abbvie, Conatus, BMS<br />

The following authors have nothing to disclose: Jennifer M. Newton, Matt Johnson,<br />

Sharon Song, Edwin Chandrasekar, Karen E. Kim<br />

580<br />

Novel Algorithm to Predict Hospital Readmission in<br />

Patients with Cirrhosis Utilizing Machine Learning Classifiers<br />

Spencer L. James 1 , Shawn Shah 2 , Zilla H. Hussain 2 , Neil Volk 2 ,<br />

Joseph Shatzel 2 , Rolland Dickson 2 ; 1 Medical School, Geisel<br />

School of Medicine at Dartmouth, Lebanon, NH; 2 Department of<br />

Medicine, Dartmouth-Hitchcock Medical Center, Lebanon, NH<br />

BACKGROUND: Patients with cirrhosis have high rates of hospital<br />

readmission. Current models used to predict readmission<br />

in cirrhotic patients are based on limited data. The aims of this<br />

study are to use machine learning classifiers (MLC), a novel<br />

discipline of computer science focused on predicting outcomes<br />

from complex datasets, to develop a model which can accurately<br />

predict which cirrhotic patients will be readmitted after<br />

discharge. METHODS: We retrospectively reviewed patients<br />

with a primary diagnosis of cirrhosis admitted to a tertiary<br />

care center over a 21-month period (April 2011 to December<br />

2012). Using admission data, we applied various MLC<br />

and a deep learning technique to develop a predictive model<br />

for readmission in cirrhotics. The MLC utilized features of the<br />

patient’s EHR including age, sex, serum sodium, platelet count,<br />

hemoglobin, hepatic biomarkers, albumin, international normalized<br />

ratio, MELD score, use of rifaximin or lactulose, use of<br />

diuretics, use of a beta-blocker, prior TIPS, presence of hepatocellular<br />

carcinoma, etiology of cirrhosis and socioeconomic<br />

factors (such as income and insurance). We then tested the<br />

model’s ability to predict the likelihood of readmission among<br />

cirrhotic patients in the database. RESULTS: During the study<br />

period, 317 unique patients accounting for a total of 498<br />

admissions were included. The average patient age was 57.9<br />

years (range of 14 to 88), and 58% were male. The mean<br />

number of admissions per patient was 2.4 (range of 1 to 9),<br />

with an average time to readmission of 88 days (range of 1<br />

to 411 days). From this sample, we trained machine learning<br />

classifiers on a randomly selected 80% of the data, and<br />

then tested the predictive accuracy on the remaining 20%.<br />

This train/test process was repeated 100 times. The mean<br />

accuracy to correctly identify hospital readmission varied by<br />

classifier, with random forest classification correctly predicting<br />

patient readmission 70% of the time on average (range of<br />

53% to 80%). Random forest demonstrated a mean sensitivity<br />

and specificity of 50% and 85%, respectively, with a PPV and<br />

NPV of 70% and 69%, respectively, in the native distribution<br />

of the dataset. CONCLUSION: Machine learning classifiers<br />

were capable of predicting hospital readmission for patients<br />

with cirrhosis nearly 70% of the time. As a decision support<br />

tool, machine learning is promising to help risk stratify patients<br />

admitted with cirrhosis. With further validation, such classifiers<br />

could be incorporated into EHR systems to help identify highrisk<br />

patients with cirrhosis who are likely to be readmitted, and<br />

prompt appropriate multidisciplinary intervention.<br />

Disclosures:<br />

The following authors have nothing to disclose: Spencer L. James, Shawn Shah,<br />

Zilla H. Hussain, Neil Volk, Joseph Shatzel, Rolland Dickson


498A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

581<br />

Differentially altered gut microbial composition correlates<br />

with plasma metabolite alterations in patients of<br />

severe alcoholic hepatitis (SAH) and steroid response<br />

Shvetank Sharma, Jaswinder S. Maras, Sukanta Das, Shabir Hussain,<br />

Saggere M. Shasthry, Shiv K. Sarin; Institute of Liver & Biliary<br />

Sciences, New Delhi, India<br />

Purpose: Gut microbiota plays an important role in systemic<br />

metabolism and liver diseases. We hypothesized that gut<br />

microbiota and their metabolic pathways vary in SAH at baseline<br />

& on steroid therapy.Patients and Methods: From108 consecutive<br />

alcoholic liver disease patients, 60 with SAH [chronic<br />

alcohol use >5years, DF>32, liver histology suggestive of<br />

SAH, absence of known liver disease or h/o immunosuppressive<br />

drugs] were treated with prednisolone(40mg/day)<br />

for 28days. Response to steroids was assessed at day7 by<br />

Lille score [0.45=non-responder(NR)].<br />

Stool samples were collected at baseline & day7 of therapy.<br />

250mg frozen samples were processed for DNA isolation &<br />

16S rRNA (V3-V4 region )analysed by NGS. Plasma was analysed<br />

for metabolites by MS/MS.Results: Total of 305 species<br />

from 135 genera were identified among six R & NR each.<br />

Klebsiella(32%), Enterococcus(24%) were identified in NR,<br />

while Bacteroides(26%),Lactobacilli(21%) were identified in R.<br />

Post-therapy, E.coli(45%) & Bifidobacterium(21%) were dominant<br />

in R &E.coli(40%) &Lactobacillus(26%) in NR.At baseline<br />

R & NR were categorized as type1 & type2 enterotypes.<br />

Post-therapy both groups demonstrated type3 composition.34<br />

metabolic pathways were common between all the treatment<br />

groups & time-points.Ferredoxin oxidoreducatse,& LPS biosynthesis,<br />

sulphate transport, Sec-system pathways were uniquely<br />

present in R0 &NR0 respectively, suggesting increase in the<br />

secretory phenotype in NR at baseline. Significant upregulation<br />

(1.5 fold, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 499A<br />

583<br />

Phenotypic and genotypic characterization of drug-induced<br />

liver injury (DILI) with autoimmune features<br />

C. Stephens 1 , A. Ortega 1 , I. Medina-Cáliz 1 , Mercedes Robles<br />

Diaz 1 , Agustin Castiella 2 , Pedro Otazua 3 , E. Zapata 2 , E.M.<br />

Gomez-Moreno 4 , MAngel López-Nevot 4 , Francisco Ruiz-Cabello 4 ,<br />

German Soriano 5 , E. Roman 5,6 , Hacibe Hallal 7 , Jose María<br />

Moreno-Planas 8 , Martin Prieto 9 , M. I. Lucena 1 , Raul J. Andrade 1 ;<br />

1 Instituto de Investigación Biomédica de Málaga (IBIMA), Hospital<br />

Universitario Virgen de la Victoria, Universidad de Málaga,<br />

CIBERehd, Málaga, Spain; 2 Hospital Mendaro, Guipúzcoa, Spain;<br />

3 Hospital Mondragón, Guipúzcoa, Spain; 4 Instituto de Investigación<br />

Biosanitario de Granada, Hospital Universitario Virgen de<br />

las Nieves, Universidad de Granada, Granada, Spain; 5 Hospital<br />

de la Santa Creu i Sant Pau, Universitat Autònoma de Barcelona,<br />

CIBERehd, Barcelona, Spain; 6 Escola Universitària d’Infermeria<br />

EUI-Sant Pau, Universitat Autònoma de Barcelona, Barcelona,<br />

Spain; 7 Hospital Morales Meseguer, Murcia, Spain; 8 Complejo<br />

Hospitalario Universitario de Albacete, Albacete, Spain; 9 Hospital<br />

La Fe, CIBERehd, Valencia, Spain<br />

Background: Positive autoantibody (AAB) titres is a common<br />

presentation of autoimmune hepatitis (AIH), but can also occur<br />

in DILI. The underlying mechanism of selective AAB occurrence<br />

in DILI is unknown. Hence, we aimed to compare demographics,<br />

clinical parameters and HLA allele compositions in DILI<br />

with positive (+) and negative (-) AAB titres and AIH patients.<br />

Material and Methods: High resolution genotyping of HLA<br />

class I (A, B, C) and II (DRB1, DQB1) loci were performed<br />

on 207 DILI (drug-induced autoimmune hepatitis excluded)<br />

and 51 AIH patients and compared to 885 Spanish healthy<br />

controls. Results: Fifty-eight of the 207 DILI patients presented<br />

positive titres for at least one AAB (ANA 76%, ASMA 28%,<br />

AMA 9% or LKM-1 2%) during the DILI episode, while 149<br />

were negative for all four AABs. Comparing demographics<br />

and clinical parameters, AAB+ patients were found to be significantly<br />

older than AAB- patients (58 vs 50 years, p=0.019).<br />

Females predominated among the AIH (71%) compared to<br />

AAB+ (59%) and AAB- (47%), p=0.01. Significant differences<br />

in type of liver injury were present (p=0.0025), with cholestatic/mixed<br />

damage being more prevalent in AAB-, which was<br />

reflected in a higher onset ALP elevation in the same group<br />

(p=0.029). Furthermore, hypertension was significantly more<br />

frequent in AAB+ patients. Compared to controls, HLA alleles<br />

B*08:01 (45% vs 10%, pc=1.0E-12), C*07:01 (47% vs 24%,<br />

pc=0.006), DRB1*03:01(59% vs 26%, pc=2.0E-05) and<br />

DQB1*02:01 (57% vs 22%, pc=3.0E-07) were significantly<br />

more frequent in AIH patients. The frequency of HLA-A*01:01<br />

was increased in the same population, but did not reach significance<br />

after Bonferroni’s correction (33% vs 19%, p=0.02/<br />

pc=0.37). There was a tendency for higher representation of<br />

DRB1*14:01 and DQB1*05:03 in DILI AAB+ compared to<br />

DILI AAB- (14% vs 4.7%, p=0.03; 14% vs 5.4%, p=0.06) and<br />

controls (14% vs 5.0%, p=0.007; 14% vs 5.4%, p=0.01).<br />

Conclusions: The AAB+ patients were found to be older and<br />

less prevalent to develop cholestatic/mixed injury compared<br />

to AAB- patients. The AIH patients were predominately female<br />

with hepatocellular injury and less likely to have underlying<br />

hypertensive conditions. The presence of HLA alleles B*08:01,<br />

C*07:01, DRB1*03:01 and DQB1*02:01 and possibly<br />

A*01:01 appear to enhance the risk of AIH in Caucasians with<br />

Spanish inheritance. These alleles form part of the conserved<br />

extended haplotype 8.1. However, haplotype formations in the<br />

study cohort are currently unknown. HLA alleles DRB1*14:01<br />

and DQB1*05:03 could potentially increase the risk of positive<br />

AAB (particularly ANA) in Spanish DILI patients. Funding:<br />

P10-CTS-6470, PI12/00378, AC-0073-2013, CIBERehd-ISCIII<br />

Disclosures:<br />

Martin Prieto - Advisory Committees or Review Panels: Bristol, Gilead<br />

The following authors have nothing to disclose: C. Stephens, A. Ortega, I. Medina-Cáliz,<br />

Mercedes Robles Diaz, Agustin Castiella, Pedro Otazua, E. Zapata,<br />

E.M. Gomez-Moreno, MAngel López-Nevot, Francisco Ruiz-Cabello, German<br />

Soriano, E. Roman, Hacibe Hallal, Jose María Moreno-Planas, M. I. Lucena,<br />

Raul J. Andrade<br />

584<br />

Effect of Branched Chain Amino Acids on Iron Transport<br />

in the Liver and Small Intestine of a Cirrhotic Rat Model<br />

Yoshinao Kobayashi 1,2 , Ryosuke Sugimoto 2 , Motoh Iwasa 2 , Yoshiyuki<br />

Takei 2 ; 1 Center for Physical and Mental Health, Mie University<br />

Graduate School of Medicine, Tsu, Japan; 2 Department of Gastroenterology<br />

and Hepatology, Mie University Graduate School of<br />

Medicine, Tsu, Japan<br />

BACKGROUND: Branched chain amino acids (BCAA) have<br />

been used as a supplemental therapy to improve malnutrition<br />

and event-free survival in patients with liver cirrhosis (LC). Iron<br />

is known to involve in inflammation and carcinogenesis of the<br />

liver through production of reactive oxygen spices (ROS). We<br />

have reported that BCAA reduce oxidative stress by diminished<br />

hepatic iron accumulation in a cirrhotic rat model. In order to<br />

clarify the effect of BCAA supplementation on iron metabolism<br />

of the liver and small intestine, we investigated effects of BCAA<br />

on hepatic iron accumulation and intestinal iron transporters<br />

in a rat model. Additionally, we investigated effects of BCAA<br />

on expression of intestinal iron transporters using CaCo2 cells.<br />

METHODS: Carbon tetrachloride (CCl 4<br />

) was administered to<br />

male Wistar rats. On the 5 th week, rats were randomly divided<br />

into 3 groups and further maintained for 16 weeks. Group 1<br />

(G1) was treated with basal diet as a control (n=8). Group 2<br />

was treated with non-heme iron diet (60mg/kg a day) (n=7).<br />

Group 3 was treated both with non-heme iron and BCAA<br />

mixture (10mg/kg a day) (n=7). On the 21 st week, liver and<br />

intestine tissues were collected. Expression of iron transporters,<br />

divalent metal transporter 1 (Dmt1) and ferroportin 1 (Fp1),<br />

was analyzed by a semi-quantitative RT-PCR and Western blotting.<br />

Furthermore, Caco2 cells were cultured for 48 hours with<br />

FeCl 3<br />

, deferoxamine (DFO) and/or BCAA. RESULTS: We have<br />

shown that BCAA-supplemented rats showed lower hepatic iron<br />

contents (P< 0.05), lower hepatic hepcidin mRNA expression<br />

(P< 0.001), lower serum hepcidin level (P= 0.05) and lower<br />

hepatic 8-hydroxyl-2’-deoxyguanosine (8-OHdG) production<br />

(P< 0.05), as compared with control rats. Protein expression<br />

of Dmt1 in the small intestine was reduced in G2 (vs. G1, P <<br />

0.05), which tended to be retrieved in G3 (vs. G2, P= 0.08).<br />

Protein expression of intestinal Fp1 tended to be reduced in G2<br />

(vs. G1, P=0.06). The reduced FP1 tended to be retrieved in<br />

G3 (vs. G2, P= 0.09). BCAA did not affect gene and protein<br />

expression of Dmt1 in Caco2 cells. Gene expression of Fp1 in<br />

these cells was not significantly changed by BCAA, whereas<br />

protein expression of Fp1 in the cell membrane fraction was<br />

reduced when cultured with BCAA. Among BCAA, leucine<br />

most strongly repressed FP1 protein expression in Caco2 cells.<br />

CONCLUSION: BCAA, especially leucine, can reduce intestinal<br />

Fp1 protein expression through post-transcriptional pathway,<br />

which was a possible mechanism for reducing intestinal iron<br />

absorption. BCAA and serum hepcidin level reduced by BCAA<br />

supplementation can modulate iron absorption from the intestine<br />

and hepatic iron accumulation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yoshinao Kobayashi, Ryosuke<br />

Sugimoto, Motoh Iwasa, Yoshiyuki Takei


500A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

585<br />

Persistent Generation of Inflammatory Mediators after<br />

Acetaminophen Overdose in Surviving and Non-Surviving<br />

Patients<br />

Benjamin Woolbright, Mitchell R. McGill, Matthew R. Sharpe,<br />

Hartmut Jaeschke; Pharmacology, Toxicology & Therapeutics, University<br />

of Kansas Medical Center, Kansas City, KS<br />

Acetaminophen (APAP) overdose is a leading cause of drug-induced<br />

liver failure. In both human overdose patients and the<br />

murine model, metabolic activation of acetaminophen leads to<br />

formation of protein adducts, mitochondrial dysfunction, necrosis<br />

and an acute inflammatory response. Considerable debate<br />

occurs over the role of the inflammatory response after APAP<br />

overdose, as to whether sterile inflammation promotes liver<br />

injury through inflammasome activation, or if sterile inflammation<br />

resolves injury through clearance of necrotic debris and<br />

stimulation of regeneration. Currently, there is limited information<br />

on the time course of cytokine formation or complement<br />

activation in human patients, especially in surviving (S) and<br />

non-surviving (NS) populations. This study addressed this deficit,<br />

with the hypothesis that NS patients would accumulate<br />

a greater degree of plasma cytokine levels, and have more<br />

severe complement depletion after APAP overdose. Circulating<br />

plasma levels of IL-1ß, TNF-α, IL-6, IL-8, IL-10, MCP-1, G-CSF,<br />

and GRO were measured by multi-plex ELISA and complement<br />

C3 was measured by ELISA in healthy volunteers (n=8),<br />

patients with APAP overdose but no increase in liver enzymes<br />

(n=10), and S (n=13) and NS (n=8) APAP overdose patients<br />

with liver injury, for the first seven days after study admission.<br />

ALT levels were similar between S and NS groups (5,531 ±<br />

639 U/L v. 5,302 ± 1,139 U/L), although bilirubin was significantly<br />

greater in the NS group (4.1 ± 1.0 mg/dL v. 8.6 ±<br />

3.1 mg/dL). Plasma levels of pro-inflammatory cytokines such<br />

as IL-1ß, GRO, and TNF-α did not rise above control levels at<br />

any point. In contrast, both S and NS patients had persistent,<br />

significant elevations in G-CSF, IL-6, IL-8, and anti-inflammatory<br />

cytokines such as IL-10 and MCP-1. Moreover, levels of IL-6,<br />

IL-8, IL-10 and MCP-1 were significantly higher in NS patients<br />

compared to S patients at multiple time points. MCP-1 levels<br />

rose to the greatest degree in both S patients (6,134 ± 1,134<br />

pg/mL) and NS patients (7,074 ± 1,105 pg/mL); whereas<br />

IL-1ß accumulated the least in both S patients (81.4 ± 64.2<br />

pg/mL) and NS patients (39.2 ± 32.6 pg/mL). Plasma C3<br />

levels were depleted at the same rate in S and NS patients, but<br />

recovered only in S patients. In conclusion, these data support<br />

the hypothesis that NS patients become refractory to anti-inflammatory<br />

cytokine signals, which likely fail to resolve injury<br />

or stimulate regeneration. Furthermore, there is limited inflammasome<br />

activation in human patients, as IL-1ß levels never rise<br />

above baseline. Importantly, however, other cytokines may<br />

have predictive value.<br />

Disclosures:<br />

The following authors have nothing to disclose: Benjamin Woolbright, Mitchell R.<br />

McGill, Matthew R. Sharpe, Hartmut Jaeschke<br />

586<br />

Protective effects of connexin43 signaling in acetaminophen-induced<br />

liver injury<br />

Michaël Maes 1 , Mitchell R. McGill 2 , Tereza Cristina da Silva 3 ,<br />

Margitta Lebofsky 2 , Isabel V. Pereira 3 , Joost Willebrords 1 , Sara<br />

Crespo Yanguas 1 , Anwar Farhood 4 , Maria Lucia Zaidan Dagli 3 ,<br />

Hartmut Jaeschke 2 , Bruno Cogliati 3 , Mathieu Vinken 1 ; 1 In Vitro<br />

Toxicology and Dermato-Cosmetology, Vrije Universiteit Brussel,<br />

Brussels, Belgium; 2 Department of Pharmacology, Toxicology and<br />

Therapeutics, University of Kansas Medical Center, Kansas City,<br />

KS; 3 Department of Pathology, School of Veterinary Medicine<br />

and Animal Science, University of São Paulo, São Paulo, Brazil;<br />

4 Department of Pathology, St. David’s North Austin Medical Center,<br />

Austin, MO<br />

Background and aims: Being goalkeepers of liver homeostasis,<br />

gap junctions are also involved in hepatotoxicity. However,<br />

their role in this process is ambiguous, as gap junctions can act<br />

as both targets and effectors of liver toxicity. This particularly<br />

holds true for drug-induced liver insults. In the present study, the<br />

role and functional relevance of connexin32 and connexin43,<br />

the building stones of liver gap junctions, were investigated<br />

in acetaminophen-induced hepatotoxicity. Methods: C57BL/6<br />

mice were overdosed with acetaminophen followed by analysis<br />

of the expression and localization of connexins as well as<br />

monitoring of hepatic gap junction functionality. Furthermore,<br />

acetaminophen-induced liver injury was compared between<br />

mice genetically deficient in either connexin32 or connexin43<br />

and wild type animals. Evaluation of the toxicological response<br />

was based on a set of clinically relevant parameters, including<br />

protein adduct formation, histopathological examination, measurement<br />

of alanine aminotransferase, cytokines, reduced and<br />

oxidized glutathione, and analysis of proliferating cell nuclear<br />

antigen expression. Results: It was found that gap junction communication<br />

deteriorates upon acetaminophen intoxication in<br />

wild type mice, which was associated with a switch in mRNA<br />

and protein production from connexin32 to connexin43. Furthermore,<br />

connexin43-deficient animals showed increased<br />

liver cell death, inflammation and oxidative stress in comparison<br />

with wild type counterparts, whereas the opposite seems<br />

to hold true for mice lacking connexin32. Conclusion: These<br />

results suggest that hepatic connexin43-based signaling protects<br />

against acetaminophen-induced liver toxicity.<br />

Disclosures:<br />

The following authors have nothing to disclose: Michaël Maes, Mitchell R.<br />

McGill, Tereza Cristina da Silva, Margitta Lebofsky, Isabel V. Pereira, Joost<br />

Willebrords, Sara Crespo Yanguas, Anwar Farhood, Maria Lucia Zaidan Dagli,<br />

Hartmut Jaeschke, Bruno Cogliati, Mathieu Vinken<br />

587<br />

The Hepatic “Matrisome” Responds Dynamically to<br />

Inflammatory Injury: Proteomic Characterization of the<br />

Transitional ECM Changes in the Liver<br />

Christine E. Dolin 1 , Veronica L. Massey 1 , Lauren G. Poole 1 ,<br />

Deanna L. Siow 1 , Michael L. Merchant 2 , Daniel W. Wilkey 2 ,<br />

Gavin E. Arteel 1 ; 1 Pharmacology and Toxicology, University of<br />

Louisville, Louisville, KY; 2 Medicine, University of Louisville, Louisville,<br />

KY<br />

Background. The impact of the changes to the hepatic extracellular<br />

matrix (ECM) in the context of collagen accumulation<br />

during fibrosis are well known. However, the hepatic ECM<br />

consists of a myriad of biologically relevant proteins beyond<br />

collagen; furthermore, changes to the hepatic ECM are not<br />

restricted to fibrosis. Indeed, transitional changes to the hepatic<br />

ECM during inflammatory liver injury may be critical in the<br />

balance between restitution versus accumulation of damage.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 501A<br />

The goal of this work was to develop a new proteomic method<br />

to characterize the hepatic ECM (“matrisome”) and document<br />

the impact of ethanol (EtOH) and lipopolysaccharide (LPS) on<br />

this compartment. Methods. Liver sections were solubilized<br />

in a series of increasingly rigorous extraction buffers (NaCl,<br />

SDS, GnHCl), to separate proteins by ‘age’ (e.g., crosslinking).<br />

Protein fragments in the extracts were identified by LC-MS/<br />

MS. ECM proteins were identified, categorized by primary<br />

role and compared qualitatively and quantitatively between<br />

groups. This method was validated using liver sections from<br />

mice administered CCl 4<br />

for 4 weeks to establish fibrosis. The<br />

effects of 6 weeks EtOH diet and/or 24 hours LPS on the patterns<br />

of the matrisome were subsequently determined. Results.<br />

The proteomic approach validated an increase in collagen I<br />

protein in CCl 4<br />

–exposed livers; moreover, several other ECM<br />

proteins and collagens (e.g., III, IV and V) were identified to be<br />

changed by CCl 4<br />

exposure. Additionally, this approach indicated<br />

that the matrisome changed dramatically after steatosis<br />

(EtOH) and/or inflammation (LPS). EtOH caused a dramatic<br />

~30% increase in the number of matrisome proteins with the<br />

least change in the pellet fraction and collagen category; LPS<br />

produced a similar response. The enhancement of LPS-induced<br />

liver damage by EtOH preexposure demonstrated unique protein<br />

changes. The distribution of proteins across the extracts<br />

also changed in some proteins changed in response to insult,<br />

suggesting changes in their ‘age.’ Many proteins that did not<br />

qualitatively change with ethanol, LPS, and/or the combination<br />

quantitatively changed. Conclusions. These results suggest<br />

that this approach can document qualitative and quantitative<br />

changes to the hepatic ECM proteome. This proteomic technique<br />

reveals the intricacies of transitional ECM remodeling,<br />

including new players that may contribute to the mechanisms<br />

by which ethanol enhances hepatic inflammatory injury. These<br />

protein players likely include novel candidates for biomarkers<br />

and drug targets and will be examined more closely in future<br />

<strong>studies</strong>. Supported by NIH/NCI (5R25CA134283-03) and<br />

NIH/NIAAA (1R01AA021978-01).<br />

Disclosures:<br />

The following authors have nothing to disclose: Christine E. Dolin, Veronica L.<br />

Massey, Lauren G. Poole, Deanna L. Siow, Michael L. Merchant, Daniel W.<br />

Wilkey, Gavin E. Arteel<br />

588<br />

Golgi Stress Response Is Associated with Anti-HIV<br />

drug-induced ER Stress and Liver Cell Injury<br />

Hui Han, Cheng Ji; Medicine, Univ Southern California, Los Angeles,<br />

CA<br />

Background & Aim Certain antivirals commonly used in the<br />

Highly Active Antiretroviral Therapies for HIV-infected patients<br />

cause liver injury. Previously we found that endoplasmic reticulum<br />

(ER) stress contributes to the drug-induced liver injury. Here<br />

we further investigated whether Golgi stress response underlies<br />

the drug-induced hepatotoxicity. Methods HepG2 cells were<br />

treated with two HIV protease inhibitors, ritonavir (RIT) and<br />

lopinavir (LOP), ER stress-inducing agents, tunicamycin (Tm)<br />

or thapsigargin (Tg), or Golgi stress-inducing agents, sodium<br />

monesin (SM) or Benzyl 2-acetamido-2-deoxy-α-D-galactopyranoside<br />

(BADG). Cell death rate was assessed with Hochest<br />

blue/Syntox green double staining or Annexin V staining.<br />

Golgi stress response and ER stress response were analyzed<br />

with respective molecular markers. Results: RIT and LOP-induced<br />

ER stress response in the HepG2 cells was comparable<br />

to that by Tm or Tg, which was indicated by the presence of<br />

XBP-1 alternative splicing, increased phosphorylation of IREa,<br />

and increased expression of GRP94, ATF6, and EDEM1. However,<br />

RIT and LOP induced cell death was not comparable to<br />

that by Tg or Tm. RIT and LOP induced 40% cell death whereas<br />

Tg or Tm induced 15-20% cell death. Correspondingly, ER<br />

stress-related death markers, CHOP and phosphorylated JNK,<br />

were significantly higher in the RIT and LOP treated cells than in<br />

Tg or Tm treated cells. Further, RIT and LOP treatment increased<br />

three Golgi stress response markers, GCP60, HSP47 and TFE3,<br />

which was comparable to the Golgi stress response induced by<br />

SM or BADG. The Golgi stress response was not detected in<br />

Tg or Tm treated cells. The SM-induced Golgi stress led to 20%<br />

cell death based on the Annexin V analysis. Conclusion Our<br />

results suggest for the first time that both ER stress and Golgi<br />

stress contribute to the antiviral RIT and LOP induced liver cell<br />

death injury (supported by R01AA018612 NCE).<br />

Disclosures:<br />

The following authors have nothing to disclose: Hui Han, Cheng Ji<br />

589<br />

Acetaldehyde Triggers HCV-Induced Hepatitis Progression:<br />

A Proposed Mechanism<br />

Murali Ganesan 1,2 , Larisa Y. Poluektova 4,2 , Jinjin Zhang 3 , Tatiana<br />

Bronich 3 , Terrence M. Donohue 1,2 , Kusum K. Kharbanda 1,2 , Dean<br />

J. Tuma 1,2 , Natalia A. Osna 1,2 ; 1 Internal Medicine, University<br />

of Nebraska Medical Center, Omaha, NE; 2 Research, Veterans<br />

Affairs Nebraska-Western Iowa Health Care System, Omaha, NE;<br />

3 College of Pharmacy, University of Nebraska Medical Center,<br />

Omaha, NE; 4 Pharmacology and Experimental Neuroscience, University<br />

of Nebraska Medical Center, Omaha, NE<br />

Background: In HCV-infected patients, alcohol consumption<br />

accelerates hepatitis progression. For many years, the mechanism<br />

of this event was attributed to an ability of ethanol to<br />

enhance HCV RNA replication. This conclusion was mainly<br />

based on in vitro experimental data obtained from ethanol<br />

non-metabolizing hepatoma cells and was not supported by<br />

consistent results from HCV patient trials. Here, we investigated<br />

the pathogenic role of ethanol metabolism-mediated regulation<br />

of HCV RNA levels in the acceleration of liver injury. Methods<br />

and Results: HCV RNA was quantified by RT-PCR in HCV-infected<br />

CYP2E1 + Huh7.5 (RLW) cells exposed for 1-48 hrs to an<br />

external acetaldehyde (Ach) generating system (AGS). Unexpectedly,<br />

after transient up-regulation of HCV RNA levels after<br />

24 hrs of AGS exposure, we observed a decline in HCV RNA<br />

after 48 hrs of treatment with Ach. This finding was confirmed<br />

by the reduction of HCV core protein expression in these cells.<br />

The Ach-induced decrease of intracellular viral content was not<br />

related to leakage of viral particles out of cells, as since HCV<br />

infectivity in supernatants was blocked. In attempting to explain<br />

these events, we postulate that HCV infection pre-sensitizes cells<br />

to Ach-induced apoptosis, which depletes HCV + hepatocytes.<br />

Indeed, the cleavage of caspase 3 and of PARP were more<br />

robust in AGS-exposed HCV + than in HCV - cells after 48 hr.<br />

Importantly, suppression of Ach-triggered apoptosis in infected<br />

cells by the pan-caspase inhibitor prevented the decrease of<br />

HCV RNA by Ach, indicating that depletion of HCV + cells by<br />

Ach was caspase-dependent. To assess the pathogenic role of<br />

HCV-containing apoptotic bodies (AB) in liver inflammation,<br />

we generated AB from HCV + and HCV - RLW cells and incubated<br />

them with human primary macrophages. There were no<br />

differences in mRNA levels of TNFα, TGFβ and IL-12 when AB<br />

from both sources were used; however, IL-18, IL-1β and IL-10<br />

mRNAs were induced at higher levels by HCV + AB. Shaping<br />

of cytokine production in macrophages by AB from HCV +<br />

cells was virus-mediated, as the presence of HCV proteins in<br />

these AB was confirmed by Western blot. In conclusion, HCV<br />

sensitizes liver cells to pro-apoptotic effects of Ach, ultimately


502A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

generating higher levels of HCV-expressing AB. Engulfment of<br />

HCV + AB by liver macrophages increases the production of<br />

pro-inflammatory cytokines, thereby exacerbating liver injury<br />

in HCV + alcohol abusers.<br />

Disclosures:<br />

The following authors have nothing to disclose: Murali Ganesan, Larisa Y. Poluektova,<br />

Jinjin Zhang, Tatiana Bronich, Terrence M. Donohue, Kusum K. Kharbanda,<br />

Dean J. Tuma, Natalia A. Osna<br />

590<br />

Mitochondrial Fission Factor (MFF) is induced and binds<br />

to Sab (SH3BP5) in acetaminophen (APAP) and other<br />

JNK dependent hepatotoxicities<br />

Sanda Win, Tin A. Than, Neil Kaplowitz; Medicine, GI/Liver, USC<br />

KSOM, LosAngeles, CA<br />

SAB is a mitochondrial membrane binding target of JNK which<br />

mediates impaired mitochondrial function, increased ROS,<br />

and sustained JNK activation in hepatotoxicity from acetaminophen<br />

(APAP), TNF, palmitic acid, and ER stressors. We have<br />

previously shown that mitochondrial translocation of DRP-1,<br />

the mediator of mitochondrial fission and a key factor in cell<br />

death, is downstream of the interaction of P-JNK and Sab.<br />

MFF is the binding target of DRP-1 in the outer membrane of<br />

mitochondria. Therefore, our aim was to elucidate the role<br />

of Sab in mitochondrial fission. We immunoprecipitated (IP)<br />

Sab from mitochondrial extracts of livers of control and APAP<br />

treated mice (2 hours). MFF was identified in the IP and greatly<br />

increased after APAP, both by western blot and proteomic<br />

analysis. Under basal conditions, low levels of MFF are found<br />

in mitochondria but levels rapidly increased within one hour<br />

after APAP. Proteinase K treatment of intact mitochondria confirmed<br />

that MFF was exclusively in the outer membrane. The<br />

increase in MFF was not inhibited by actinomycin D suggesting<br />

that MFF is stabilized by P-JNK and Sab rather than by induction<br />

of transcription. MFF was also induced in other JNK/Sab<br />

dependent toxicity models such as tunicamycin, palmitic acid<br />

and TNF/galactosamine. Sab knockdown or liver conditional<br />

knockout of Sab abrogated the increase of MFF and prevented<br />

DRP-1 translocation to mitochondria. However, basal levels of<br />

MFF were not altered in the absence of Sab. We previously<br />

showed that knockdown of DOK4, an inner membrane Src<br />

binder, prevented mitochondrial inactivation of Src, mitochondrial<br />

dysfunction, increased ROS, sustained JNK activation,<br />

and cell death from APAP and TNF/galactosamine. Knockdown<br />

of DOK4 had no effect on Sab levels but prevented<br />

the induction of MFF suggesting that P-JNK/Sab mediated<br />

mitochondrial dysfunction leads to increased to MFF levels. In<br />

conclusion, the interaction of P-JNK and Sab leads to induction<br />

of MFF, enhancing mitochondrial fission. MFF appears to be<br />

stabilized by mitochondrial dysfunction and binding to Sab.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sanda Win, Tin A. Than, Neil<br />

Kaplowitz<br />

591<br />

Hormesis in cholestatic liver disease; preconditioning<br />

with low bile acid concentrations protects against bile<br />

acid-toxicity<br />

Esther M. Verhaag, Manon Buist-Homan, Han Moshage, Klaas<br />

Nico Faber; Gastroenterology and Hepatology, University of<br />

Groningen, Groningen, Netherlands<br />

INTRODUCTION: Cholestasis is characterized by accumulation<br />

of bile acids and inflammation, causing hepatocellular damage,<br />

eventually resulting in liver failure. Bile acids induce apoptotic<br />

and necrotic cell death, however, hepatocytes are able to<br />

adapt towards a hostile environment. This suggests a hormetic<br />

response toward bile acids limits hepatocyte cell death during<br />

cholestasis. AIM: To investigate in vitro the mechanisms that<br />

underlie the hormetic response that protect hepatocytes against<br />

experimental cholestatic conditions. METHODS: HepG2.rNTCP<br />

cells were preconditioned (24 h) with sub-apoptotic concentrations<br />

(0.1-50 mM) of various bile acids, the superoxide donor<br />

menadione, the pro-inflammatory cytokine TNF-α or the FXR<br />

agonist GW4064, followed by a challenge with the apoptosis-inducing<br />

bile acid glycochenodeoxycholic acid (GCDCA;<br />

200 mM for 4 h). Levels of caspase-3/7 activity (apoptosis),<br />

cellular LDH leakage (necrosis) and the bile acid exporter BSEP<br />

(ABCB11) mRNA were analyzed. RESULTS: Preconditioning<br />

with the pro-apoptotic bile acids GCDCA or taurocholic acid<br />

(TCA), and the protective bile acids (tauro)ursodeoxycholic<br />

acid (TUDCA, UDCA) reduced GCDCA-induced caspase-3/7<br />

activity in HepG2.rNtcp cells. None of the bile acid-preconditioning<br />

conditions induced cellular LDH leakage from GCD-<br />

CA-challenged HepG2.rNtcp cells above control (untreated<br />

cells) levels. In contrast, preconditioning with sub-apoptotic<br />

concentrations of cholic acid (CA) potentiated GCDCA-induced<br />

apoptosis and similar trends were observed for menadione<br />

and TNF-α preconditioning. CDCA preconditioning did not<br />

change GCDCA-induced levels of caspase-3/7. The hormetic<br />

effect of GCDCA preconditioning was concentration- and<br />

time-dependent, providing significant protection against the<br />

apoptotic GCDCA challenge after at least 12 h preconditioning<br />

with concentrations of GCDCA as low as 1 mM. GCDCA<br />

and CDCA preconditioning both enhanced mRNA levels of<br />

ABCB11. The FXR agonist GW4064 more potently induced<br />

ABCB11 mRNA levels, but did not cause a significant reduction<br />

in GCDCA-induced caspase-3/7 activity, though a trend<br />

towards a partial suppression was observed. CONCLUSIONS:<br />

Sub-toxic concentrations of bile acids in the range that occur<br />

under normal physiological conditions protect HepG2.rNtcp<br />

cells against GCDCA-induced apoptosis, which is largely independent<br />

of FXR signaling. Thus, low concentrations bile acids<br />

enable hepatocytes to withstand sudden increases in bile acid<br />

concentrations and this may explain why only limited levels of<br />

apoptotic cell death is observed in obstructive cholestasis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Esther M. Verhaag, Manon<br />

Buist-Homan, Han Moshage, Klaas Nico Faber<br />

592<br />

A protective role of C/EBP homologous protein in acetaminophen-induced<br />

acute hepatocyte damage in mice<br />

Tsutomu Matsubara 1 , Yuga Teranishi 2 , Kazuki Nakatani 1 , Norifumi<br />

Kawada 2 , Kazuo Ikeda 1 ; 1 Department of Anatomy and<br />

Regenerative Biology, Graduate School of Medicine, Osaka City<br />

University, Osaka, Japan; 2 Department of Hepatology, Graduate<br />

School of Medicine, Osaka City University, Osaka, Japan<br />

Background: Acetaminophen (APAP) overdose causes acute<br />

liver failure as a result of increased oxidative stress followed by<br />

hepatocellular necrosis. C/EBP homologous protein (CHOP)<br />

expression is induced the liver after exposure to APAP, but<br />

the role of CHOP in the APAP-induced liver injury is not fully<br />

understood. This study investigated whether CHOP acts on<br />

the APAP-induced liver injury using wild-type and Chop-null<br />

mice. Methods: Acute liver injury was induced in wild-type<br />

and Chop-null mice by administrating acetaminophen (APAP,<br />

300 mg/kg). The mice were sacrificed at six hours after the<br />

APAP injection. Liver damage was evaluated by measuring<br />

serum ALT levels and liver histology. Hepatic mRNA and pro-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 503A<br />

tein levels were determined by quantitative PCR and western<br />

blot analysis. Results: Six hours after 300 mg/kg of APAP<br />

injection, serum ALT levels were 3831 and 7897 U/L and<br />

relative hepatic necrosis area was 22.7 and 47.1% of view<br />

field in wild-type and Chop-null mice, respectively. The Chopnull<br />

mice showed a severer liver injury than the wild-type mice.<br />

When phosphorylations of AKT, ERK, JNK, and p38 proteins<br />

in the livers were investigated, APAP-enhanced JNK phosphorylation<br />

was reduced in the Chop-null mice, compared to the<br />

wild-type. In addition, MKK4 and JUN phosphorylations were<br />

also reduced in Chop-null mice, suggesting that JNK signal<br />

was attenuated in the Chop-null livers. Increased mRNA levels<br />

of TIFA, an NF-kB activator, and enhanced phosphorylation of<br />

FOS, an NF-kB-target protein, were observed in the livers of<br />

the Chop-null mice, suggesting that NF-kB signal was accelerated<br />

in the Chop-null livers. Hepatic ratio of LC3-II protein<br />

to LC3-I protein, an indicator of autophagy, was increased in<br />

the wild-type mice after the APAP injection (2.4 fold), but the<br />

increase was diminished in the Chop-null mice. Furthermore<br />

protein levels of hepatic p62 (also called SQSTM1), protein<br />

decomposed under autophagy, were higher in the Chop-null<br />

mice than that in the wild-type mice. There results strongly indicated<br />

a view that autophagy signal was attenuated in the livers<br />

of Chop-null mice Conclusions: In this study, molecular linkage<br />

of CHOP-JNK-autophagy was identified as a protective factor<br />

in APAP-induced liver injury. Although further study is required,<br />

accelerated NF-kB signal in Chop-null livers may interfere with<br />

the protective role of the JNK-autophagy.<br />

Disclosures:<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

The following authors have nothing to disclose: Tsutomu Matsubara, Yuga Teranishi,<br />

Kazuki Nakatani, Kazuo Ikeda<br />

593<br />

The Effect of Treatment with Chemical Chaperones and<br />

Zinc Acetate on Hepatocytes Treated with Excess Copper<br />

Shinji Oe, Koichiro Miyagawa, Yuichi Honma, Masaru Harada;<br />

School of Medicine, University of Occupational and Environmental<br />

Health, Japan., Fukuoka, Japan<br />

[Purpose] Copper is an essential trace element in human activity.<br />

But excess copper gives harm to human activity. Wilson<br />

disease is a genetic disorder characterized by excess copper<br />

deposition in various organs. Excess copper-derived oxidants<br />

contribute to progression of liver disease in Wilson disease.<br />

Oxidative stress induces accumulation of abnormal proteins.<br />

The endoplasmic reticulum (ER) is subcellular organelle which<br />

plays the role of proper protein folding. Accumulation of misfolded<br />

proteins disturbs ER homeostasis resulting in ER stress.<br />

Zinc acetate has been commonly used to treat Wilson disease,<br />

because it increases metallothionein production and inhibits<br />

absorption of copper in the intestinal epithelia. The effects of<br />

chemical chaperones and zinc acetate on copper-induced hepatotoxicity<br />

have not been examined. In this study, we evaluated<br />

the effects of chemical chaperone and zinc acetate on copper-induced<br />

disruption of cell homeostasis. [Methods] We used<br />

human hepatoma cell line (Huh7) and immortalized human<br />

hepatocyte cell line (OUMS29). The following materials were<br />

used: copper sulfate; acetyl-leucyl-leucyl-norleucinal and epoxomicin<br />

as proteasome inhibitors (PIs); bathocuproine disulfonate<br />

as copper chelator; n-acetyl-l-cysteine; zinc acetate; 4-phenylbutyrate<br />

(PBA) and ursodeoxycholic acid (UDCA) as chemical<br />

chaperones. We detected the reactive oxygen species using<br />

2’,7’-dichlorodihydrofluorescein .<br />

We analyzed copper-induced<br />

ER stress by immunoblotting for ER stress markers, including<br />

phospho α-subunit of eukaryotic initiation factor 2 and X-box<br />

binding protein 1. Hepatocyte apoptosis treated with copper<br />

with or without PI was determined by detection of cleaved poly<br />

ADP-ribose polymerase and cleaved caspase 3 by immunoblotting<br />

analysis and immunofluorescence staining, respectively.<br />

Furthermore, we examined cell proliferation on excess copper<br />

by immunofluorescence staining using Ki67. [Results] Excess<br />

copper induced oxidative stress and ER stress. Chemical chaperones<br />

and zinc acetate reduced copper-induced oxidative<br />

stress and ER stress. Co-treatment of copper and PI exacerbated<br />

copper-induced apoptosis. Although excess copper leaded to<br />

reduction in cell proliferation, chemical chaperones rescued<br />

the proliferation block in cells treated with copper. However<br />

zinc acetate did not exert this effect. [Conclusions] Excess copper<br />

induced hepatotoxicity. Chemical chaperones, PBA and<br />

UDCA, and zinc acetate reduced hepatotoxicity. These results<br />

suggest that chemical chaperones may have beneficial effects<br />

for the treatment of Wilson disease. On the other hand, zinc<br />

acetate used as the treatment of Wilson disease may inhibit<br />

liver injury directly.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shinji Oe, Koichiro Miyagawa,<br />

Yuichi Honma, Masaru Harada<br />

594<br />

Sumoylation Regulates Lipopolysaccharide-induced<br />

Macrophage Activation Modulating Ubiquiting conjugating<br />

enzyme 9 Phosphorylation in Mouse Liver and<br />

Kupffer Cells<br />

Maria Lauda Tomasi 2,1 , Komal Ramani 2 , Minjung Ryoo 2 ; 1 Medicine,<br />

University of Southern California, Los Angeles, CA; 2 Medicine,<br />

Cedars-Sinai Medical Center, Los Angeles, CA<br />

Propose of study: Activation of macrophages by lipopolysaccharide<br />

(LPS), a conserved component of the Gram-negative<br />

bacterium’s outer membrane, leads to production of a large<br />

number of immunoregulatory molecules including tumor necrosis<br />

factor α (TNF-α), interleukin 1β (IL-1β), interleukin 6 (IL-6),<br />

arachidonic acid, as well as reactive oxygen species (ROS).<br />

LPS level is increased in the blood of alcohol abusers and ethanol-fed<br />

rodents. The small ubiquitin-related modifier (SUMO)<br />

family of proteins is a more recently discovered posttranslational<br />

modification. Cellular stress arising from ROS is a robust stimulus<br />

for protein SUMOylation. SUMOylation is widely deemed<br />

as a protective mechanism; loss of SUMO conjugation reduces<br />

cell and organism viability. Ubiquitin conjugating enzyme 9<br />

(UBC9) is the only SUMO E2 enzyme. We demonstrated that<br />

UBC9 is phosphorylated by Cyclin-dependent kinase 1 (CDK1)<br />

and is more stable when phosphorylated. Furthermore, we<br />

recently found that LPS lowers UBC9 and SUMO expression<br />

in RAW cells, nevertheless the mechanism is still unknown. The<br />

aim of this study was to examine the role of Ubc9-mediated<br />

sumoylation in LPS-activated macrophages and elucidate the<br />

molecular mechanism(s). Methods: Studies were done using in<br />

vivo LPS-treated mice, primary mouse Kupffer and RAW cells.<br />

Real-time PCR and Western blotting measured mRNA and protein<br />

levels, respectively. UBC9 phosphorylation was assessed<br />

by phos-tagÔ SDS-PAGE. Results: LPS treatment increased ROS<br />

production (H 2<br />

O 2<br />

) in RAW cells, while UBC9 knockdown alone<br />

or co-treatment with LPS completely prevented LPS-induced<br />

ROS production. UBC9 siRNA alone or with LPS co-treatment<br />

induced apoptosis and increased IL-1β, iNOS, IL-6 and TNF-α<br />

mRNA levels in RAW cells; however, when LPS and UBC9<br />

siRNA were combined, they further raised these inflammatory<br />

signals. We found that LPS treatment lowers UBC9 protein


504A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

level in the LPS-treated mouse liver, primary Kupffer and RAW<br />

cells as compared with controls, whereas UBC9 mRNA level<br />

decreased minimally. Phos-tagÔ SDS-PAGE showed that LPS<br />

increased UBC9 phosphorylation in mouse liver and primary<br />

Kupffer cells. These results support a potential key mechanism<br />

for LPS that induces pro-inflammatory cytokines through lowering<br />

UBC9 expression, raising its phosphorylation. Conclusions:<br />

The novel finding that LPS lowers UBC9 protein level has<br />

important implications since this facilitates the pro-inflammatory<br />

response. A better understanding of molecular mechanisms on<br />

how LPS treatment influences sumoylation-regulated cytokines<br />

in Kupffer cells, will open a crucial area of investigation providing<br />

new targets for therapy in alcoholic liver diseases.<br />

Disclosures:<br />

The following authors have nothing to disclose: Maria Lauda Tomasi, Komal<br />

Ramani, Minjung Ryoo<br />

595<br />

A comparative analysis of the Spanish and Latin-American<br />

prospective drug-induced liver injury (DILI) networks<br />

Fernando Bessone 1 , Nelia Hernández 2 , Adriana Sanchez Ciceron<br />

2 , Maria Di Pace 2 , Gisela Gualano 3 , Marco Arrese 4 , Alex Ruiz 4 ,<br />

Aurora Loaeza del Castillo 5 , Marcos A. Girala 6 , Enrique Carrera 7 ,<br />

Maribel Lizarzábal 8 , Edgardo Mengual 8 , Javier Brahm 9 , Juan P.<br />

Arancibia 9 , Federico C. Tanno 1 , Milagros B. Davalos-Moscol 10 ,<br />

Raymundo Paraná 11 , M I. Schinoni 11 , Nahum Méndez-Sanchéz 12 ,<br />

Miguel E. Garassini 13 , Rodrigo L. Zapata 14 , I. Medina-Cáliz 15 ,<br />

Andrés González-Jiménez 15 , Mercedes Robles Diaz 15 , C. Stephens<br />

15 , A. Ortega 15 , Judith Sanabria 15 , Miren García-Cortés 15 ,<br />

B. Garcia-Muñoz 15 , M. I. Lucena 15 , Raul J. Andrade 15 ; 1 Hospital<br />

Provincial del Centenario, Rosario, Argentina; 2 Hospital de<br />

Clínicas, Montevideo, Uruguay; 3 Hospital Posadas, Buenos Aires,<br />

Argentina; 4 Pontificia Universidad Católica de Chile, Santiago,<br />

Chile; 5 Instituto Nacional de Ciencias Médicas y Nutrición “Salvador<br />

Zubirán”, Hospital General de México, Ciudad de Mexico,<br />

Mexico; 6 Hospital de Clínicas, Asunción, Paraguay; 7 Hospital de<br />

Especialidades Eugenio Espejo, Quito, Ecuador; 8 Hospital Universitario<br />

de Maracaibo, Maracaibo, Venezuela, Bolivarian Republic<br />

of; 9 Hospital Clínico Universidad de Chile, Santiago, Chile;<br />

10 Hospital Rebagliati, Facultad de Medicina, Universidad San<br />

Martin de Porres, Lima, Peru; 11 Hospital Universitario Prof. Edgard<br />

Santos, Universidad Federal da Bahía, Salvador de Bahía, Brazil;<br />

12 Fundación Clínica Médica Sur, Ciudad de Mexico, Mexico;<br />

13 Hospital Universitario de Caracas, Caracas, Venezuela, Bolivarian<br />

Republic of; 14 Universidad de Chile, Clínica Alemana de<br />

Santiago, Santiago, Chile; 15 Instituto de Investigación Biomédica<br />

de Málaga (IBIMA), Hospital Universitario Virgen de la Victoria,<br />

Universidad de Málaga, CIBERehd, Málaga, Spain<br />

Background: DILI characteristics concerning phenotype and<br />

involved drugs or other toxic compounds can vary between<br />

individuals and possibly between different geographic populations.<br />

We aimed to compare all DILI cases included in the<br />

ongoing Spanish and Latin-American DILI Network that share<br />

the same inclusion criteria and operational procedures. Material<br />

and methods: Demographics, clinical parameters and<br />

causative agents were compared between 200 Latin-American<br />

and 867 Spanish DILI cases. Results: The mean age of DILI<br />

development differed between the two registries with 51 years<br />

in LatinAmerica and 54 years in Spain (p=0.02). Females predominated<br />

among the LatinAmerican cases (59%) compared<br />

to the Spanish cases (49%) (p=0.01). Duration of treatment<br />

and time to onset were higher in LatinAmerican cases (127 vs<br />

88 days, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 505A<br />

Drug properties (Liver Toxicity Knowledge Base, Drug Discovery<br />

Today 2011) and host factors were compared in DO vs.<br />

NDO. Results: Among the 680 cases, 159 cases (22%) manifested<br />

DILI with 2 to 82 days’ delay after the treatment cessation<br />

(i.e., DO) while 521 (77%) manifested during the drug treatment<br />

(i.e., NDO). 57% of DO cases and 13% of NDO were<br />

Amoxicillin-clavulanate (AMOX/CA) cases. After excluding the<br />

AMOX/CA cases, DO cases had shorter treatment duration<br />

(median: 11 vs. 45 days, p=50% were less prevalent in<br />

DO cases (67% vs. 84%, p=0.003) also drugs with un-metabolized<br />

drug excretion >=50% were more prevalent in DO cases<br />

(25% vs. 9%, p=0.0011). Regarding host manifestations,<br />

eosinophilia was more prevalent in DO cases (30% vs. 19%,<br />

p=0.036), while positive autoantibody was more prevalent in<br />

NDO cases (25% vs. 11%, p=0.024). NDO cases were more<br />

associated with chronic underlying diseases (82% vs. 60%,<br />

p 50% loss, n=14]. Results: The 28 subjects<br />

with bile duct loss included 14 men [5 severe] and 14 women<br />

[9 severe]. The median age was 52.5 [range 11-87] years<br />

those with severe duct loss were younger [47.4 vs 59.7 years,<br />

p=0.04]. Cases were attributed to 20 different agents: 3 to<br />

amoxicillin/clavulanic acid; 3 to temozolomide; 3 to botanicals<br />

[Cactus nopal; Artemisia annua; unknown]; 2 to azithromycin;<br />

and 1 each due to cefalexin, cefazolin, levofloxacin, moxifloxacin,<br />

allopurinol, enalapril, hydralazine, infliximab, lenalidomide,<br />

thalidomide, montelukast, olanzapine, quetiapine,<br />

lansoprazole, omeprazole, metoclopramide and mesalazine.<br />

Median latency to onset was 35 [IQR 19-91] days. The typical<br />

clinical presentation was immuno-allergic, rather than autoimmune-like<br />

disease and was usually cholestatic or mixed, with<br />

median R-value= 2.1 [range 0.1-8.0]. In addition to bile duct<br />

paucity, hepatic pathology showed reactive bile duct injury<br />

and mild portal lympho-histiocytic inflammatory infiltrates.<br />

Eosinophils were noted in 9 biopsies, but peripheral eosinophilia<br />

at disease onset was noted in only 6 patients. 7 of 14<br />

subjects with mild and 8 of 14 with severe injury transitioned<br />

to chronic ongoing injury [laboratory, imaging or clinical evidence<br />

of persisting liver damage > 6 months after onset]. The<br />

likelihood of dying or requiring liver transplantation within 6<br />

months of onset was greater in those with severe bile duct<br />

loss (4/14, 29%) than in mild duct loss (0/14, 0%, p=0.05]<br />

and also greater than in those without duct loss on biopsy<br />

(22/246, 8.9%, p=0.04). Conclusions: Paucity of bile ducts<br />

is an uncommon (6%) histologic feature of acute drug-induced<br />

liver injury, can be caused by diverse agents, both drugs and<br />

botanicals, and often leads to evidence of chronic liver injury,<br />

which, when severe, may lead to premature death or liver<br />

transplantation. Pathogenesis remains imperfectly understood,<br />

but host immune reactions, perhaps influenced by genetic or<br />

environmental factors, appear likely; however, the key antigen[s]<br />

remain unknown.<br />

Disclosures:<br />

Herbert L. Bonkovsky - Advisory Committees or Review Panels: Recordati Rare<br />

Chemicals, Clinuvel, Inc.; Consulting: Alnylam, Inc, Clinuvel, Inc., Clinuvel, Inc.;<br />

Grant/Research Support: Gilead Sciences<br />

Mark W. Russo - Grant/Research Support: Merck, Salix; Speaking and Teaching:<br />

janssen, Gilead, ABBVIE, Salix<br />

The following authors have nothing to disclose: David E. Kleiner, Jiezhun Gu,<br />

Joseph A. Odin, Victor J. Navarro, Maricruz Vega, Jay H. Hoofnagle<br />

598<br />

Caspase Inhibition Switched TNF-α-Induced Apoptosis<br />

to Necroptosis But Not Autophagic Cell Death In<br />

Hepatocytes and Mouse Livers<br />

Hong-Min Ni, Xiaojuan Chao, Mitchell R. McGill, Hartmut<br />

Jaeschke, Wen-Xing Ding; Pharmacology, Toxicology and Therapeutics,<br />

The University of Kansas Medical Center, Kansas City, KS<br />

Hepatocyte death (apoptosis and necrosis) and survival (such<br />

as autophagy) are integrated and play important roles in many<br />

liver diseases such as endotoxin and alcohol-induced liver<br />

injury. How these different cell death and survival pathways<br />

regulate each other remains elusive. We previously demonstrated<br />

that when the cell survival NF-κb pathway was blocked<br />

by a general transcription inhibitor actinomycin D (ActD), tumor<br />

necrosis factor-a (TNF-a) induced apoptosis in primary cultured<br />

mouse hepatocytes, which was completely suppressed by a<br />

general caspase inhibitor, ZVAD. Surprisingly, in the present<br />

study, we found that TNF-a/ActD-treated hepatocytes died<br />

by a necrotic-like cell death evidenced by abundant cellular


506A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

vacuoles, propidium iodide positive stained intact nuclei and<br />

release of nuclear HMGB1 after a pro-longed treatment for up<br />

to 48 hours. TNF-a/ActD induced cleavage of receptor interacting<br />

protein kinase 1 (RIP1) and RIP3 as well as Beclin 1, key<br />

molecules that regulate necroptosis and autophagy, respectively.<br />

These cleavages were mediated by caspases because<br />

they were blocked by ZVAD. Necrostatin 1 (a RIP1 inhibitor)<br />

or hepatocytes isolated from RIP3 knockout mice partially<br />

protected against TNF-a/ActD+ZVAD-induced cell death, but<br />

antioxidants (N-acetylcysteine or MnSOD) and a JNK inhibitor<br />

(SP600125) had no effect. Using adeno-RFP-GFP-LC3 infected<br />

hepatocytes, we found that TNF-a/ActD+ZVAD decreased<br />

autophagic flux. Moreover, TNF-a/ActD+ZVAD-induced cell<br />

death in hepatocyte isolated from liver-specific Atg5 knockout<br />

mice was similar to wild type hepatocytes. These data suggest<br />

that TNF-a/ActD+ZVAD induced RIP1/3-mediated necroptosis<br />

but not autophagic cell death in vitro. Similar to primary<br />

hepatocytes, we found that ZVAD also protected against LPS/<br />

galactosamine (GalN)-induced apoptosis and liver injury at 6<br />

hours in mice. However, this protection was lost at 24 hours<br />

and these mice developed severe hepatic necrosis, neutrophil<br />

infiltration and liver injury. Intriguingly, while pharmacological<br />

inhibition of RIP1 or RIP3 knockout mice suppressed LPS/Gal-<br />

N+ZVAD-induced necrosis, they did not protect against LPS/<br />

GalN+ZVAD-induced liver injury due to increased caspase-independent<br />

apoptosis in these mouse livers. In conclusion, our<br />

results suggest that different cell death and cell survival pathways<br />

are closely integrated during TNF-a-induced liver injury<br />

when both caspase and NF-κb are blocked. Blocking caspase<br />

switched TNF-a-induced apoptosis to RIP1/3-dependent necroptosis<br />

but not autophagic cell death. Thus, results from our<br />

study also raise concerns on the safety of current ongoing clinical<br />

trials using caspase inhibitors.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hong-Min Ni, Xiaojuan Chao,<br />

Mitchell R. McGill, Hartmut Jaeschke, Wen-Xing Ding<br />

599<br />

Mitochondrial-shaping proteins as specific biomarkers<br />

to distinguish alcohol from fat-induced liver toxicity<br />

Elena Palma 1 , Antonio Riva 1 , Satvinder Mudan 2 , Nikolai Manyakin<br />

2 , Roger Williams 1 , Shilpa Chokshi 1 ; 1 Institute of Hepatology-<br />

Foundation for Liver Research, London, United Kingdom; 2 The<br />

London Clinic, London, United Kingdom<br />

Alcoholic Liver Disease (ALD) and Non-Alcoholic Liver Disease<br />

(NAFLD) show similar pathological spectra and present common<br />

factors (lipotoxicity, oxidative stress), but differ in clinical<br />

aspects and outcomes. A clearer understanding of their<br />

pathogenesis would be helpful in a better management of these<br />

disorders. Morphological changes as abnormal enlargements<br />

in the hepatic mitochondria, particularly megamitochondria<br />

(MM) formation, have been reported in liver biopsies from ALD<br />

patients since the 1970s but are not associated with NAFLD.<br />

Mitochondrial architecture is intimately involved with their<br />

function and adapts rapidly in response to the needs of the<br />

cell with cycles of fusion (organelle binding) and fission (fragmentation).<br />

These changes are strictly regulated by the activity<br />

of the mitochondrial-shaping proteins (MSP). Our aim is to<br />

elucidate the role of mitochondrial dynamics, shape and MSP<br />

in ALD and NAFLD in order to identify early disease specific<br />

markers. Human precision cut liver slices and VL-17A cells<br />

(positive for ADH/CYP2E1) were cultured with ethanol (EtOH)<br />

and/or oleic/linoleic acid (FFA), to induce hepatotoxicity.<br />

Post-treatment cell viability, intracellular fat accumulation and<br />

mitochondrial functionality were evaluated and changes in the<br />

mitochondrial shape were assessed by confocal and electron<br />

microscopy; MSP expression and activation were analysed by<br />

Western blot. The toxic effect of EtOH and FFA was confirmed<br />

by an increase in the percentage of apoptosis or a reduction<br />

in ATP levels and an impairment of mitochondrial respiration.<br />

The two insults induced a very different mitochondrial phenotype,<br />

as well as differential intracellular fat accumulation. Fat<br />

was associated with a more pronounced toxicity and excessive<br />

mitochondrial fission, while EtOH alone showed a moderate<br />

effect. With EtOH exposure, the surviving cells showed a dramatically<br />

increased proportion of MM, and this was associated<br />

with a reduced activation of the master fission regulator<br />

Drp-1 (Dynamin related protein-1). In parallel, changes in MSP<br />

involved in mitochondrial fusion were observed with fat-induced<br />

hepatotoxicity. In conclusion, we demonstrate that hepatotoxic<br />

agents such as EtOH or fat affect the mitochondrial dynamics<br />

and shape differently, altering the fusion/fission equilibrium<br />

and MSP expression. In particular, we show MM formation in<br />

response to the chronic EtOH insult in hepatocytes that survive,<br />

suggesting this as an adaptive mechanism to avoid death and<br />

raising the potentiality for the use of MSP as therapeutic targets<br />

for ALD and disease-specific biomarkers potentially discriminating<br />

between the early toxic events observed in ALD and<br />

NAFLD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elena Palma, Antonio Riva, Satvinder<br />

Mudan, Nikolai Manyakin, Roger Williams, Shilpa Chokshi<br />

600<br />

Patients with childhood-onset primary sclerosing cholangitis<br />

harbor rare, deleterious variants in genes involved<br />

with cholestatic syndromes and generalized risk of PSC<br />

Brian D. Juran, Aliya Gulamhusein, Saurabh Baheti, Konstantinos<br />

Lazaridis; Mayo Clinic, Rochester, MN<br />

Purpose: Genome-wide association <strong>studies</strong> (GWAS) of primary<br />

sclerosing cholangitis (PSC) have linked common genetic variants<br />

to PSC risk; yet, rare genetic variants also likely impact<br />

this disease. Next-generation sequencing provides access to<br />

these variants; however, poor functional predictability and the<br />

expectation of low-penetrance among alleles not manifesting<br />

in disease until adulthood, poses a serious obstacle to meaningful<br />

<strong>studies</strong>. Here, we focus on patients with childhood-onset<br />

PSC, as they more-likely harbor highly-penetrant variants, and<br />

concentrate on protein-alterations among candidate genes with<br />

prior likelihood for PSC impact, to maximize functional prediction.<br />

This effort will advance our scant knowledge regarding<br />

childhood PSC, and lay the groundwork towards application of<br />

sequencing-based approaches to the at-large PSC population.<br />

Methods: 10 patients with PSC diagnosis prior to age 12 were<br />

identified in the PSC Resource of Genetic, Environment and Synergy<br />

Studies (PROGRESS). Exome capture was by Agilent Sure<br />

Select XT Human Exome +UTR v5 kit and sequencing was on<br />

an Illumina HiSeq2000. Alignment, calling of single nucleotide<br />

variants (SNVs), and QC was performed using an established<br />

pipeline based on the Genome Analysis Toolkit Best Practices<br />

workflow; annotation was performed using BioR. Missense or<br />

nonsense SNVs were filtered based on quality and inclusion<br />

in manually-curated lists of (a) 28 known cholestasis genes or<br />

(b) 15 PSC GWAS-implicated genes. Final filtering kept only<br />

rare SNVs (population minor allele frequency < 1.5%) and<br />

those expected to be functional by at least 1 of 3 predictive<br />

tools (SIFT, PolyPhen-2 or Mutation Taster). Results: 14 SNVs<br />

remained after filtering; 13 were single-instance alleles found<br />

in 1 patient and 1 was found in 2 patients (rs12371484 in<br />

SH2B3). 9 of the SNVs were from the list of known cholestasis


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 507A<br />

genes and located in the following genes ABCB4, ABCC2,<br />

ATP8B1, HSD3B7, NOTCH2 (1 SNV each) and SERPINA1,<br />

TJP2 (2 SNVs each). The 5 SNVs from the PSC GWAS list were<br />

located in the genes CD226, GPR35, MST1 (1 SNV each) and<br />

SH2B3 (2 SNVs). Overall, 9/10 childhood-onset PSC patients<br />

had at least one of the SNVs (4 patients with 1 SNV, 4 with<br />

2 SNVs and 1 patient with 3 SNVs). Conclusions: Patients<br />

with childhood-onset PSC harbor rare, deleterious variants in<br />

genes involved with cholestasis and risk of PSC, suggesting that<br />

disease in these patients may lie along a spectrum between<br />

syndromic forms of cholestatic disease and classical PSC. Additional<br />

<strong>studies</strong> including patients with more-typical disease onset<br />

will help to confirm this finding and may establish a role for<br />

these genes in adult disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Brian D. Juran, Aliya Gulamhusein,<br />

Saurabh Baheti, Konstantinos Lazaridis<br />

well that patients should be treated early in disease, when<br />

ALP and bilirubin levels first exceed ULN, and that patients be<br />

treated to the lowest levels possible to avoid liver transplant<br />

and premature death.<br />

Disclosures:<br />

David Jones - Consulting: Intercept, Pfizer, Novartis; Speaking and Teaching:<br />

Falk, Shire<br />

Henk R. van Buuren - Grant/Research Support: Intercept, Zambon Nederland BV<br />

The following authors have nothing to disclose: Bettina E. Hansen, Willem J.<br />

Lammers, George F. Mells, Marco Carbone<br />

601<br />

Convergence of Two Predictive Models of Risk Reduction<br />

in Patients with Primary Biliary Cirrhosis (PBC)<br />

Bettina E. Hansen 1 , Willem J. Lammers 1 , David Jones 2 , Henk R. van<br />

Buuren 1 , George F. Mells 3 , Marco Carbone 4 ; 1 Erasmus MC, University<br />

Medical Center Rotterdam, Rotterdam, Netherlands; 2 Institute<br />

of Cellular Medicine, Newcastle University Medical School,<br />

Newcastle Upon Tyne, United Kingdom; 3 Academic Department<br />

of Medical Genetics, University of Cambridge, Cambridge, United<br />

Kingdom; 4 Division of Gastroenterology and Hepatology, Department<br />

of Medicine, Addenbrooke’s Hospital, Cambridge, United<br />

Kingdom<br />

Background: Two recent analyses on prognostic algorithms<br />

in large PBC cohorts provides a unique opportunity to test the<br />

replicability of effect. Objective: The goal of this study was to<br />

apply common assumptions to algorithms developed by the<br />

Global PBC and UK PBC cohorts to predict liver transplant-free<br />

survival, and to determine if the models are comparable in<br />

terms of quantified risk. Methods: The Global PBC cohort<br />

consists of 4,565 PBC patients treated with ursodeoxycholic<br />

acid (UDCA), recruited from 15 centers in 8 countries. An<br />

algorithm predicting liver transplant and all-cause mortality<br />

was created using Cox proportional hazards regression on<br />

a derivation sample (n=2488), and validated in a matched<br />

sample (n=1631). Predictor variables included: age, and<br />

serum bilirubin, alkaline phosphatase, albumin and platelet<br />

count measured 1 year after UDCA initiation. The UK PBC<br />

cohort (n=4022) was recruited from 150 National Health Service<br />

(NHS) Trusts across the UK and includes both patients<br />

treated and not treated with UDCA. An algorithm predicting<br />

liver transplant and liver-related death was derived from 2414<br />

patients and validated in 1608 patients. The predictor algorithm<br />

includes serum bilirubin, alkaline phosphatase, albumin,<br />

platelet count and ALT. The absolute risk for liver transplant and<br />

survival was calculated using equivalent fixed inputs for albumin<br />

(40 g/L or 1.143 of LLN=35) and platelets (120 X 10 9 /L);<br />

age set at 65 and ALT at 84 U/L. Bilirubin varied as a function<br />

of ULN: 1X = 17 mmol/L; 2X = 34 mmol/L; 3X = 51 mmol/L;<br />

4X = 68 mmol/L and ALP by 10 U/L, with range as a function<br />

of ULN: 1X = 150 U/L to 4X = 600 U/L. Results are shown in<br />

the figure below. Conclusion: The calculations showed comparability<br />

at bilirubin > 2ULN and near comparability at bilirubin<br />

at ULN, providing further compelling evidence that both ALP<br />

and bilirubin are reliable predictors of PBC outcomes. The convergence<br />

of data from these models argues strongly for a reassessment<br />

of treatment goals. These results indicate that patients<br />

with advanced disease are at highest need of treatment, as<br />

602<br />

Differentially Expressed Plasma miRNAs as Novel Biomarkers<br />

for the Early Detection of Primary Sclerosing<br />

Cholangitis and Cholangiocarcinoma<br />

Francesca Bernuzzi 2 , Francesco Marabita 4 , Ana Lleo 2 , Ilaria Bianchi<br />

2 , Massimiliano Mirolo 5 , Marco Marzioni 6 , Gianfranco Alpini 7 ,<br />

Domenico Alvaro 8 , Kirsten M. Boberg 9 , Massimo Locati 5 , Guido<br />

Torzilli 10 , Lorenza Rimassa 11 , Fabio Piscaglia 12 , Xiao-Song He 1 ,<br />

Patrick S. Leung 1 , Christopher L. Bowlus 3 , Guo-Xiang Yang 1 , M.<br />

Eric Gershwin 1 , Pietro Invernizzi 2 ; 1 Internal Medicine, University<br />

of California, Davis, CA; 2 Liver Unit and Center for Autoimmune<br />

Liver Diseases, Humanitas Clinical and Research Center, Milan,<br />

Italy; 3 Division of Gastroenterology and Hepatology, University of<br />

California at Davis, Davis, CA; 4 Unit of Computational Medicine,<br />

Department of Medicine, Karolinska Institute, Stockholm, Sweden;<br />

5 Department of Medical Biotechnologies and Translational Medicine,<br />

University of Milan, Milan, Italy; 6 Department of Gastroenterology,<br />

Universita Politecnica delle Marche, Ancona, Italy;<br />

7 Department of Medicine, Division of Gastroenterology, Texas<br />

A&M University Health Science Center, Temple, TX; 8 Department<br />

of Clinical Medicine, Division of Gastroenterology, University of<br />

Rome, Rome, Italy; 9 Medical Department, Rikshospitalet, Oslo,<br />

Norway; 10 Liver Surgery Unit, Department of Surgery, Humanitas<br />

Clinical and Research Center, Milan, Italy; 11 Medical Oncology<br />

and Hematology Unit, Humanitas Clinical and Research Center,<br />

Milan, Italy; 12 Department of Medical and Surgical Sciences<br />

DIMEC, Alma Mater Studiorum, University of Bologna, Bologna,<br />

Italy<br />

Background: Cholangiocarcinoma (CCA) is the second most<br />

common primary hepatic malignancy and a frequently fatal<br />

complication of primary sclerosing cholangitis (PSC) with the<br />

highest incidence within the first 2.5 years following the diagnosis<br />

of PSC. The lack of sensitive and specific biomarkers<br />

poses a major challenge in the early diagnosis of CCA in the<br />

setting of PSC. Specific plasma microRNAs (miRNAs) have<br />

been reported in a number of inflammatory and neoplastic<br />

conditions. The aim of the study was an intensive discovery<br />

phase to identify specific plasma miRNAs as diagnostic biomarkers<br />

for PSC and CCA. Methods: Ninety human plasma<br />

samples (30 PSC, 30 CCA, and 30 healthy subjects) were<br />

surveyed for the levels of 667 miRNAs by qRT-PCR to identify<br />

disease-associated candidate miRNAs (discovery phase).<br />

miRNAs that demonstrated an absolute log 2<br />

(fold change) ><br />

1, mean Ct < 30, and p < 0.05 were selected for validation.<br />

Candidates were validated in an independent cohort of<br />

130 samples (40 PSC, 40 CCA, 10 primary biliary cirrhosis


508A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(PBC) and 40 healthy subjects). To evaluate the discriminating<br />

effect of plasma miRNAs, ROC curve were established for<br />

each validated miRNAs, and AUC was calculated from the<br />

logistic regression. Only miRNAs with AUC>0.70 were considered<br />

useful as biomarkers. Results: In the discovery phase,<br />

we identified 21 miRNAs with differential expression in PSC,<br />

33 in CCA, and 26 in both, when compared with healthy controls,<br />

of which 24 miRNAs demonstrated significantly diverse<br />

expression between PSC and CCA. Data from validation phase<br />

demonstrated that miR-200c was differentially expressed in<br />

PSC versus healthy controls, whereas miR-483-5p, and miR-<br />

194 were differentially expressed in CCA compared with<br />

healthy controls. Importantly, we also found 3 miRNA that were<br />

specifically deregulated in CCA when compared to PSC: miR-<br />

222 was upregulated and miR-16 and miR-195 were down<br />

regulated in CCA when compared to PSC. The combined ROC<br />

analysis of identified miRNAs from each disease group significantly<br />

increases the AUC value and the combination of<br />

these markers further improved the specificity and accuracy of<br />

diagnosis. Conclusions: This study is unique and an important<br />

discovery program that emphasizes the potential identification<br />

of plasma miRNAs as biomarkers of PSC and CCA. A panel<br />

of differentially expressed plasma miRNAs is identified in PSC<br />

and CCA when compared with healthy controls. Furthermore,<br />

our data provides a basis for the use of miRNAs as biomarkers<br />

not only for the diagnosis of PSC and for the early detection of<br />

CCA, but also novel therapeutic targets. Future analyses will<br />

depend on larger cohorts of patients.<br />

Disclosures:<br />

Lorenza Rimassa - Board Membership: Eli Lilly, Merck Serono<br />

Fabio Piscaglia - Advisory Committees or Review Panels: Bayer ; Consulting:<br />

Bracco, Eisai; Grant/Research Support: Esaote<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

The following authors have nothing to disclose: Francesca Bernuzzi, Francesco<br />

Marabita, Ana Lleo, Ilaria Bianchi, Massimiliano Mirolo, Marco Marzioni,<br />

Gianfranco Alpini, Domenico Alvaro, Kirsten M. Boberg, Massimo Locati, Guido<br />

Torzilli, Xiao-Song He, Patrick S. Leung, Guo-Xiang Yang, M. Eric Gershwin,<br />

Pietro Invernizzi<br />

603<br />

Bezafibrate: A novel and effective alternative for relieving<br />

pruritus in patients with primary biliary cirrhosis<br />

Anna Reig, Pilar Sese, Albert Pares; Liver Unit, Hospital Clinic,<br />

University of Barcelona, IDIBAPS, CIBERehd, Barcelona, Spain<br />

Background and Aims: Pruritus is a common and distressing<br />

symptom in patients with primary biliary cirrhosis (PBC), and<br />

when uncontrollable it is an indication for liver transplantation.<br />

Different therapeutic approaches for pruritus have been<br />

proposed including resins, rifampicin, naltrexone, sertraline<br />

and albumin dialysis. Recent observations have indicated that<br />

fibrates may improve cholestatic itching, although no specific<br />

<strong>studies</strong> have been carried out. Therefore, we have assessed<br />

the effects of fibrates on pruritus in patients PBC who had a<br />

suboptimal response to ursodeoxycholic acid (UDCA). Patients<br />

and Methods: 46 PBC patients (43 females, age 54.3 ± 1.5<br />

years) with suboptimal biochemical response to UDCA were<br />

treated with bezafibrate (400 mg/d). Apart from clinical and<br />

biochemical changes, pruritus severity was assessed by a specific<br />

questionnaires (PBC-40 and pruritus score) and with a<br />

visual analogue scale (VAS) (form 0 to 10), at baseline and<br />

after a mean of 29 ± 4 months. Moreover, bezafibrate therapy<br />

was discontinued in 13 patients to evaluate the course<br />

of this symptom after bezafibrate withdrawal. Results: Twenty<br />

seven patients (58.7%) experienced pruritus at baseline (mean<br />

VAS: 4.4 ± 0.5). No significant differences regarding to prior<br />

duration of UDCA therapy, age, gender and severity of biochemical<br />

cholestasis were observed between patients with and<br />

without pruitus at the beginning of bezafibrate therapy, except<br />

for ALT which were significantly higher in patients with pruritus.<br />

Triglyceride and cholesterol levels as well as transient elastography<br />

(8.9 ± 0.8 vs 9.3 ±2.7 kPa, p:n.s) were also similar in<br />

patients with and without pruritus. Bezafibrate therapy resulted<br />

in a significant mitigation of pruritus (VAS from 4.4 ± 0.5 to<br />

0.8 ± 0.2,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 509A<br />

were comparable between PSC, PBC and HC (0.26 [0.37],<br />

0.23 [0.27] and 0.18 [0.18] ng/mL, resp., p=0.27), baseline<br />

fasted TBS were elevated in PSC compared to PBC and HC<br />

(29.0 [62.2], 3.2 [7.0] and 2.8 [3.4] mmol/L, resp., p


510A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

nostic groups. To evaluate each analyte as a classifier, receiver<br />

operating characteristic (ROC) curves were constructed and the<br />

area under the curve (AUC), sensitivity, and specificity were<br />

calculated as classifier performance metrics. Multinomial logistic<br />

regression coupled with a forward stepwise selection procedure<br />

and a leave-one-out cross validation analysis was then<br />

used to select analytes for inclusion into composite classifiers.<br />

RESULTS: Importantly, this methodology is strikingly more sensitive<br />

and indeed there were 30 glycoconjugates differentially<br />

expressed between PBC and controls, 27 between PSC and<br />

controls, and 20 between PBC and PSC. Age contributed to<br />

the expression of 25 glycoconjugates. Multi-analyte classifiers<br />

designed as disease-specific (i.e. PBC vs control, PSC vs control,<br />

and PBC vs PSC) diagnostic tests were able to perform their<br />

designated tasks perfectly (AUCs, sensitivities and specificities<br />

of 1). CONCLUSIONS: Glycosylation is significantly altered in<br />

patients with PBC. Glycoconjugates can serve as sensitive and<br />

specific biomarkers capable of distinguishing between specific<br />

autoimmune diseases. Indeed, future work can be used to focus<br />

not only on diagnosis, but also as biomarkers for disease severity<br />

and prognosis.<br />

Disclosures:<br />

L. Renee Ruhaak - Employment: Intel<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

The following authors have nothing to disclose: Kyoungmi Kim, Sandra L. Taylor,<br />

Qiuting Hong, Forum Patel, Carol Stroble, Patrick S. Leung, Michiko Shimoda,<br />

M. Eric Gershwin, Carlito B. Lebrilla, Emanual Maverakis<br />

607<br />

Phenotypic Variations in Primary Sclerosing Cholangitis<br />

Across the Age Spectrum<br />

John E. Eaton 1 , Brian D. Juran 1 , Elizabeth J. Atkinson 2 , Bryan<br />

McCauley 2 , Erik M. Schlicht 1 , Mariza de Andrade 2 , Velimir A.<br />

Luketic 8 , Joseph A. Odin 3 , Ayman A. Koteish 4 , Kris V. Kowdley 5 ,<br />

Kapil B. Chopra 6 , Gideon Hirschfield 7 , Naga P. Chalasani 9 , Konstantinos<br />

Lazaridis 1 ; 1 Gastroenterology and Hepatology, Mayo<br />

Clinic, Rochester, MN; 2 Biomedical Statistics and Informatics,<br />

Mayo Clinic, Rochester, MN; 3 Gastroenterology and Hepatology,<br />

Icahn School of Medicine at Mount Sinai, New York, NY; 4 Gastroenterology<br />

and Hepatology, Johns Hopkins, Baltimore, MD; 5 Liver<br />

Care Network, Swedish Medical Center, Seattle, WA; 6 Gastroenterology<br />

and Hepatology, University of Pittsburgh, Pittsburgh, PA;<br />

7 Centre for Liver Research and NIHR Biomedical Research Unit,<br />

University of Birmingham, Birmingham, United Kingdom; 8 Gastroenterology<br />

and Hepatology, Virginia Commonwealth University,<br />

Richmond, VA; 9 Gastroenterology and Hepatology, Indiana University<br />

School of Medicine, Indianapolis, IN<br />

Background & Aims: Primary sclerosing cholangitis (PSC) typically<br />

develops in middle-aged men. However, it is unknown if<br />

phenotypic differences exist among patients diagnosed with<br />

PSC at various ages. To this end, we compared the clinical<br />

characteristics of a PSC cohort based on the age when PSC<br />

was diagnosed. Methods: We performed a patient survey<br />

and a medical record review to compare the phenotypic features<br />

of PSC patients (n=967) who were diagnosed between<br />

1-19 years (yrs) (early onset, n=108), 20-59 yrs (middle age<br />

onset, n=763), and 60-79 yrs (late onset, n=96). Patients were<br />

recruited prospectively from 8 academic medical centers in<br />

North America. Results: There were no demographic differences<br />

between groups. However, those with early onset-PSC<br />

were more likely to have ulcerative colitis (UC) (61.1%) compared<br />

to the middle age onset (54.1%) and late onset (40.6%)<br />

groups (p=0.01). Similarly, concomitant autoimmune hepatitis<br />

was more common in early onset-PSC when compared to<br />

the other groups (13.9% vs. 4.8% for middle age onset and<br />

7.3% for late onset, p=0.001). Cholangiocarcinoma (CCA)<br />

was diagnosed in 54 PSC patients (early onset 0%, middle age<br />

onset 7% and late onset 8.3%). While CCA in early onset-PSC<br />

appears to be rare, we did not detect a statistically significant<br />

difference between groups and the overall development<br />

of CCA (log rank p value=0.32). However, those with late<br />

onset-PSC were more likely to be diagnosed with CCA within a<br />

year of their PSC diagnosis when compared to the other groups<br />

(early onset 0%, middle age onset 2.5% and late onset 6.2%,<br />

p = 0.02). Two-hundred and eleven individuals underwent a<br />

liver transplantation (early onset 14%, middle age onset 25%<br />

and late onset 6.2%) and survival free from liver transplant was<br />

similar between groups after adjusting for PSC duration and<br />

gender (late onset vs. early onset: [Hazard ratio] HR; 2.1, 95%<br />

Confidence Interval [CI]: 0.8-5.7, p: 0.15; middle age onset<br />

vs. early onset: HR; 0.8, 95% CI: 0.5-1.5, p=0.50). Lastly,<br />

early onset-PSC was associated with a higher proportion of<br />

second-degree relatives with inflammatory bowel disease (IBD)<br />

when compared to the other groups (19.4% vs. 6.8% for middle<br />

age onset and 2.1% for late onset, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 511A<br />

cholangitis (ASC) – and 16 healthy subjects (HS) were studied.<br />

Peripheral blood cell phenotype was determined by flow<br />

cytometry; ability to suppress was evaluated as inhibition of<br />

cell proliferation/effector cytokine production; ectoenzymatic<br />

activity by thin layer chromatography; expression of adenosine<br />

receptor, adenosine deaminase (ADA) and phosphodiesterases<br />

(PDE) by quantitative real-time PCR or Western Blot.<br />

Results: As compared to their CD39 - counterpart, Th17 CD39+<br />

cells retained expression of CCR6, IL-23R and RORC, classical<br />

Th17 cell markers; displayed higher frequencies of CD69 + ,<br />

CD44 + and CD25 + activated cells; contained higher numbers<br />

of CD73 + , CD161 + and FOXP3 + lymphocytes; and were more<br />

frequently positive for IL-22, IL-10 and TGF-b producing cells.<br />

The proportion of Th17 CD39+ cells was markedly lower in AILD<br />

than HS, with ASC displaying lower proportions than AIH<br />

patients. In AILD, Th17 CD39+ numbers are greatly decreased<br />

and have impaired ability to generate AMP/adenosine and<br />

to control both target cell proliferation and IL-17 production.<br />

When compared to HS, Th17 cells from AILD patients had a<br />

markedly decreased expression of A2A adenosine receptor<br />

while displaying similar expression levels of PDE4A, PDE4B<br />

and ADA. Conclusions: In AILD, Th17 CD39+ cells are numerically<br />

decreased and defective in their ability to generate adenosine<br />

and exert suppression. Failure of these Th17 cells to upregulate<br />

CD39 appears linked to reduced A2A adenosine receptor levels.<br />

Low CD39 and A2A expression may contribute, at least<br />

in part, to the perpetuation of Th17 effector potential in AILD.<br />

Disclosures:<br />

Simon C. Robson - Grant/Research Support: Pfizer, NIH, Dainippon; Independent<br />

Contractor: Biolegend, EMD Millipore, Mersana; Management Position:<br />

eBioscience; Speaking and Teaching: ACP, Elsevier, ATC; Stock Shareholder:<br />

Nanopharma, Puretech<br />

The following authors have nothing to disclose: Rodrigo Liberal, Charlotte R.<br />

Grant, Yun Ma, Michael A. Heneghan, Giorgina Mieli-Vergani, Diego Vergani,<br />

Maria Serena Longhi<br />

OCA on the markers of cholestasis and safety. Study inclusion<br />

criteria: PBC diagnosis, ALP ≥1.67x ULN and/or total bilirubin<br />

>ULN to


512A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Mitchell L. Shiffman - Advisory Committees or Review Panels: Merck, Gilead,<br />

Boehringer-Ingelheim, Bristol-Myers-Squibb, Abbvie, Janssen, Acchillion; Consulting:<br />

Roche/Genentech; Grant/Research Support: Merck, Gilead, Boehringer-Ingelheim,<br />

Bristol-Myers-Squibb, Abbvie, Beckman-Coulter, Achillion, Lumena,<br />

Intercept, Novartis, Gen-Probe; Speaking and Teaching: Roche/Genentech,<br />

Merck, Gilead, Abbvie, Janssen, Bayer<br />

Karel J. van Erpecum - Advisory Committees or Review Panels: Bristol Meyers<br />

Squibb, Abbvie, Janssen Cilag, Gilead<br />

Roya Hooshmand-Rad - Employment: Intercept pharmaceuticals Inc.<br />

Shawn Sheeron - Employment: Intercept Pharmaceuticals<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: Giuseppe Mazzella, Pietro Invernizzi,<br />

Joost Drenth, Jaroslaw Regula, Annarosa Floreani, Velimir A. Luketic,<br />

Victor Vargas, Catherine Vincent, Bettina E. Hansen<br />

610<br />

Patients with Primary Biliary Cirrhosis Experience Progressive<br />

Bone Loss Over Time Beyond that Expected by<br />

Age and Despite Routine Clinical Care<br />

Aliya Gulamhusein 1 , Bryan McCauley 2 , Elizabeth J. Atkinson 2 ,<br />

Brian D. Juran 1 , Erik M. Schlicht 1 , Mariza de Andrade 2 , Konstantinos<br />

Lazaridis 1 ; 1 Center for Basic Research in Digestive Diseases,<br />

Mayo Clinic, Rochester, Rochester, MN; 2 Health Science<br />

Research, Mayo Clinic, Rochester, MN<br />

Background and Aims: Osteopenia and osteoporosis are<br />

characterized by low bone mass and increased fracture risk.<br />

Previous reports suggest that patients with primary biliary cirrhosis<br />

(PBC) are at increased risk of osteoporosis with estimates<br />

varying between 15-50%, but study populations and testing<br />

approaches have been heterogeneous. Furthermore, the rate<br />

of bone loss in these patients is poorly understood. We sought<br />

to estimate the prevalence of low bone mass and characterize<br />

the change in bone mineral density (BMD) over time in<br />

a large cohort of patients with PBC. Methods: We identified<br />

patients recruited to the Mayo Clinic PBC Genetic Epidemiology<br />

Registry and Biorepository who had available dual X-ray<br />

absorptiometry (DXA) scans for assessment of BMD. Those with<br />

comorbid autoimmune hepatitis were excluded and data was<br />

censored at the time of liver transplant. Femoral neck (FN) BMD<br />

(g/cm 2 ) and T-score were used to assess bone loss. Per WHO<br />

criteria, osteopenia was defined as a T-score of < -1 but > -2.5<br />

and osteoporosis as a T-score of ≤ -2.5. Linear mixed effects<br />

models were fit on subjects with multiple scans to estimate the<br />

changes in BMD over time using gender, age, and time since<br />

diagnosis as covariates. Results: 383 patients with 885 DXA<br />

scans were included. There were 813 scans on 348 women<br />

and 71 scans on 35 men. The median number of scans per<br />

subject was 2 (25%-75% IQR=1-3) and follow-up was over 9.2<br />

years (25%-75% IQR=3.4-16.4). The prevalence of low bone<br />

mass (osteopenia or osteoporosis) in all patients aged


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 513A<br />

Disclosures:<br />

David Jones - Consulting: Intercept, Pfizer, Novartis; Speaking and Teaching:<br />

Falk, Shire<br />

The following authors have nothing to disclose: Claire Hardie, Kile J. Green,<br />

Sarah Pagan, Jessica K. Dyson, Lucy J. Walker, Laura Jopson, John G. Brain<br />

612<br />

Vitamin A deficiency promotes hepatic expansion of<br />

resident (Kupffer cells) and peripheral (Gr hi ) macrophages<br />

leading to excessive liver damage in rats with<br />

obstructive cholestasis<br />

Ali Saeed, Mark Hoekstra, Martijn O. Hoeke, Janette Heegsma,<br />

Han Moshage, Klaas Nico Faber; Department of Gastroenterology<br />

and Hepatology, University Medical Center Groningen, Groningen,<br />

University of Groningen, The Netherlands, Groningen, Netherlands<br />

Vitamin A deficiency (VAD) is associated with chronic liver diseases<br />

(CLD), including biliary atresia, primary biliary cirrhosis,<br />

primary sclerosing cholangitis and non-alcoholic fatty liver disease.<br />

Serum and/or hepatic retinol levels negatively correlate<br />

with disease progression. Previously, we found that VAD rats<br />

shown much more severe liver damage when exposed to bile<br />

duct ligation (BDL) compared to rats with sufficient vitamin A<br />

(VAS) levels, but the mechanisms involved remain unresolved<br />

so far. Vitamin A has potent anti-inflammatory properties and<br />

therefore we investigated its effect on resident (Kupffer cells)<br />

and peripheral (Gr hi ) macrophages in vitro and in vivo. Weaning<br />

male Wister rats were fed either a VAS or a VAD diet for<br />

14-16 weeks, followed by BDL for 1, 2, 4, 7 days. One group<br />

subjected to 7-day BDL received retinyl-palmitate therapy. Liver<br />

damage marker were analysed in serum, and liver tissue was<br />

analysed with Q-PCR, Western blotting and immunohistochemistry.<br />

Primary rat Kupffer cells and THP-1 (human Leukemic<br />

monocyte cell line) cells were exposed to VAS and VAD conditions<br />

and cytokine expression was analysed by Q-PCR. CD68<br />

(general macrophage marker) and CD11b (peripheral macrophages)<br />

positive cells were significantly increased in livers<br />

of VAD rats after 4- and 7-day BDL as compare to VAS-BDL<br />

control livers, and was associated with increased plasma γGT<br />

levels and expression of liver fibrosis makers, such as α-SMA<br />

and Collagen1A1. All these features were strongly suppressed<br />

in VAD-BDL rats receiving vitamin A therapy. In line, expression<br />

of the macrophage chemoattractant Ccl2 was significantly<br />

increased in VAD-BDL livers and suppressed by vitamin A therapy.<br />

VAD-BDL livers showed increased expression of proinflammatory<br />

and profibrotic markers, such as iNos, Tnf-α, Il-1,<br />

Il-6, Cox-2, Tgf-β and vitamin A therapy decreased this inflammatory<br />

response. In vitro, VAD (RPMI with lipid-stripped serum)<br />

caused an inflammatory polarization in primary Kupffer cells<br />

and PMA-activated-THP-1 cells as compare to control cells and<br />

increased the expression of inflammatory markers, including<br />

iNos, Tnf-α, Il-1, Il-6, Cox-2. The vitamin A metabolites, 9-cis<br />

retinoic acid and all trans retinoic acid, reduced expression<br />

of these inflammatory markers. In conclusion, vitamin A deficiency<br />

promotes hepatic macrophage expansion in obstructive<br />

cholestasis, which show an aggressive pro-inflammatory and<br />

pro-fibrotic phenotype that aggravates liver damage, but is<br />

effectively prevented by vitamin A supplementation. This study<br />

supports an direct role of vitamin A in preventing excessive<br />

liver damage in patients with chronic liver diseases.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ali Saeed, Mark Hoekstra, Martijn<br />

O. Hoeke, Janette Heegsma, Han Moshage, Klaas Nico Faber<br />

613<br />

The Modulation of Co-stimulatory Molecules by Circulating<br />

Exosomes in Primary Biliary Cirrhosis<br />

Takashi Tomiyama 1,2 , Guo-Xiang Yang 1 , Ming Zhao 4 , Weici<br />

Zhang 1 , Hajime Tanaka 1,5 , Jing Wang 4 , Patrick S. Leung 1 , Kazuichi<br />

Okazaki 2 , Xiao-Song He 1 , Quiajin Lu 4 , Ross L. Coppel 3 ,<br />

Christopher L. Bowlus 1 , M. Eric Gershwin 1 ; 1 Internal Medicine,<br />

University of California, Davis, CA; 2 Third Department of Internal<br />

Medicine, Division of Gastroenterology and Hepatology, Kansai<br />

Medical University, Osaka, Japan; 3 Department of Microbiology,<br />

Monash University, Melbourne, VIC, Australia; 4 Department of<br />

Dermatology, The Second Xiangya Hospital, Central South University,<br />

Hunan, China; 5 Department of Gastroenterology and<br />

Metabolism, Nagoya City University Graduate School of Medical<br />

Sciences, Nagoya, Japan<br />

Background: Exosomes are secreted intracellular microparticles<br />

that can induce antigen-specific immune responses and potentially<br />

trigger inflammation. The role of exosomes in autoimmunity<br />

has drawn significant attention because of their ability to<br />

modulate cell-to-cell communications in part by their ability to<br />

regulate cytokine and chemokine activation. However, the contribution<br />

of exosomes in autoimmune cholangitis has not been<br />

addressed. Herein, we studied the effect of exosomes from<br />

patients with primary biliary cirrhosis (PBC) and healthy controls<br />

(HCs) on cytokine production and co-stimulatory molecule<br />

expression in peripheral mononuclear cell populations. Further,<br />

we also determined the micro RNA (miRNA) profiles of circulating<br />

exosomes and verified with an independent cohort. Methods:<br />

Exosomes were purified from plasma from 28 patients<br />

with PBC and 25 HCs. Their effect on cytokine production and<br />

co-stimulatory molecule expression in mononuclear cell populations<br />

were examined using an ex vivo system. In addition, miR-<br />

NAs profile of the exosomes was measured by Agilent Human<br />

miRNA (8*60K) V19.0 microarray. miRNAs with a level of<br />

2-fold or larger difference between groups were considered<br />

differentially expressed and verified using isolated exosomes<br />

from additional groups of PBC and HC using a Taqman PCR<br />

assay for individual miRNAs. Further, we examined whether<br />

the levels of these exosomal miRNAs correlated with severity of<br />

disease. Results: Our data showed that circulating exosomes<br />

from PBC are taken up by antigen presenting cells (APCs) and<br />

affect the expression of cell surface co-stimulatory molecules<br />

CD80, CD86 and CD40 on APCs. Circulating exosomes in<br />

PBC contain altered patterns of miRNA expression with 9 miR-<br />

NAs significantly up-regulated and another 9 miRNAs significantly<br />

down-regulated when compared to the HC. Conclusions:<br />

Exosomes significantly alter co-stimulatory molecule expression<br />

on antigen presenting cell populations and there were differences<br />

of miRNA expression in circulating exosomes in patients<br />

with PBC. The rigorous dissection of exosomes and their constituents<br />

is an important component for analysis to understand the<br />

nature of the pathogenic effects produced during the natural<br />

history of autoimmune cholangitis.<br />

Disclosures:<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

The following authors have nothing to disclose: Takashi Tomiyama, Guo-Xiang<br />

Yang, Ming Zhao, Weici Zhang, Hajime Tanaka, Jing Wang, Patrick S. Leung,<br />

Kazuichi Okazaki, Xiao-Song He, Quiajin Lu, Ross L. Coppel, M. Eric Gershwin


514A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

614<br />

Increased expression of ORM1-like protein 3 (ORMDL3)<br />

in biliary epithelial cells may be related to the pathogenesis<br />

of primary biliary cirrhosis<br />

Motoko Sasaki 1 , Yasunori Sato 1 , Yasuni Nakanuma 2,1 ; 1 Human<br />

Pathology, Kanazawa University Graduate School of Medicine,<br />

Kanazawa, Japan; 2 Shizuoka Cancer Center, Shizuoka, Japan<br />

Background/Aims: ORM1-like protein 3 (ORMDL3) localizes<br />

to the endoplasmic reticulum (ER), mediates calcium homeostasis<br />

and facilitates the unfolded-protein response, which is<br />

related to the endogenous induction of inflammation. ORMDL3<br />

gene is located in chromosome 17q12-q21region, which is<br />

one of candidates associated with the development of primary<br />

biliary cirrhosis (PBC), asthma and Crohn’s disease. Although<br />

upregulated expression of ORMDL3 was reported in airway<br />

epithelial cells in asthma, the expression of ORMDL3 in biliary<br />

epithelial cells (BECs) in PBC and other liver diseases has not<br />

been reported, so far. We hypothesized that altered expression<br />

of ORMDL3 in BECs in PBC may be related to the deregulated<br />

autophagy and abnormal expression of mitochondria<br />

antibodies specifically seen in PBC. Methods: We examined<br />

immunohistochemically the expression of ORMDL3 in BECs in<br />

livers taken from patients with PBC (n=50), control diseased<br />

livers (n=55), such as primary sclerosing cholangitis (PSC),<br />

and normal livers (n=20). The association between the expression<br />

of ORMDL3 and the expression of mitochondrial antigen<br />

PDC-E2, autophagy-related markers (LC3, p62) or senescent<br />

markers (p16 INK4a and p21 WAF1/Cip1 ) was also examined. In<br />

addition, a correlation between the extent of biliary expression<br />

of ORMDL3 and the extent of ORMDL3-positive inflammatory<br />

cells infiltration in small bile ducts was evaluated. Furthermore,<br />

we examined the expression of ORMDL3 at mRNA level in cultured<br />

BECs treated with various cytokines (IL-4 [10ng/ml]; IL-13<br />

[10ng/ml]; IFN-γ [1000U/ml], TNF-α [10ng/ml] and TGF-β<br />

[4ng/ml]) and a protein kinase A activator, forskolin (5 mM).<br />

Results: The expression of ORMDL3 was seen in the cytoplasm<br />

and apical surface in damaged small bile ducts (SBDs) in PBC.<br />

The expression of ORMDL3 was significantly more extensive<br />

in SBDs in PBC, especially in early stage PBC, compared with<br />

control diseased livers and normal livers (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 515A<br />

616<br />

Association between elevated serum IgG4 (sIgG4) concentrations<br />

and the phenotype of patients with primary<br />

sclerosing cholangitis (PSC)<br />

Michael P. Manns 1 , Bertus Eksteen 2 , Mitchell L. Shiffman 3 , Cynthia<br />

Levy 4 , Kris V. Kowdley 5 , Aldo J. Montano-Loza 6 , Harry L. Janssen<br />

7 , Robert P. Myers 8 , Dora Ding 8 , Mani Subramanian 8 , John<br />

G. McHutchison 8 , Michael R. Charlton 9 , Christopher L. Bowlus 10 ,<br />

Andrew J. Muir 11 , Roger W. Chapman 12 ; 1 Hannover Medical<br />

School, Hannover, Germany; 2 University of Calgary, Calgary, AB,<br />

Canada; 3 Liver Institute of Virginia, Richmond, VA; 4 University of<br />

Miami, Miami, FL; 5 Swedish Medical Center, Seattle, WA; 6 University<br />

of Alberta, Edmonton, AB, Canada; 7 University of Toronto,<br />

Toronoto, ON, Canada; 8 Gilead Sciences, Inc., Foster City, CA;<br />

9 Intermountain Medical Center, Salt Lake City, UT; 10 University of<br />

California at Davis, Sacramento, CA; 11 Duke Clinical Research<br />

Institute, Durham, NC; 12 University of Oxford, Oxford, United<br />

Kingdom<br />

Background: Elevated sIgG4 has been reported in 9-15% of<br />

PSC patients and is associated with a more aggressive disease<br />

course. Our aim was to compare the characteristics<br />

of PSC patients with elevated and normal sIgG4. Methods:<br />

We measured sIgG4 (BN II System; Siemens, Malvern, PA)<br />

in PSC patients enrolled in a phase 2b trial of simtuzumab.<br />

Corticosteroid and/or anti-TNF-α therapies were prohibited.<br />

The associations between elevated sIgG4 (>140 mg/dL) with<br />

demographics, body mass index (BMI), ulcerative colitis (UC),<br />

use of ursodeoxycholic acid (UDCA), liver biochemistry, Fibro-<br />

Test, ELF, sLOXL2 (VIDAS® LOXL2; bioMérieux, Marcy L’Etoile,<br />

France), liver fibrosis staged by the Ishak classification, and<br />

Mayo risk score (MRS) were determined. MRCP data will be<br />

available at the time of presentation. Results: Among 234<br />

patients, 34 (14.5%) had elevated sIgG4. These patients were<br />

older than those with normal sIgG4, but sex, race, UC, and use<br />

of UDCA did not differ between groups (Table). Although liver<br />

biochemistry did not differ, patients with elevated sIgG4 had<br />

lower serum albumin and higher platelet levels compared with<br />

those with normal sIgG4. FibroTest, ELF and sLOXL2, the proportion<br />

of patients with bridging fibrosis or cirrhosis, and MELD<br />

and MRS were similar between groups. A sensitivity analysis<br />

examining a sIgG4 cut-off of >201 mg/dL revealed similar<br />

findings. Conclusions: A small proportion of PSC patients have<br />

elevated sIgG4. In this clinical trial cohort, the sIgG4 level<br />

does not have a significant impact on PSC phenotype including<br />

disease severity assessed biochemically, histologically, or<br />

according to conventional prognostic indices.<br />

Characteristics of PSC Patients According to sIgG4<br />

Median (IQR) or % (n).<br />

Disclosures:<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Mitchell L. Shiffman - Advisory Committees or Review Panels: Merck, Gilead,<br />

Boehringer-Ingelheim, Bristol-Myers-Squibb, Abbvie, Janssen, Acchillion; Consulting:<br />

Roche/Genentech; Grant/Research Support: Merck, Gilead, Boehringer-Ingelheim,<br />

Bristol-Myers-Squibb, Abbvie, Beckman-Coulter, Achillion, Lumena,<br />

Intercept, Novartis, Gen-Probe; Speaking and Teaching: Roche/Genentech,<br />

Merck, Gilead, Abbvie, Janssen, Bayer<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Robert P. Myers - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Michael R. Charlton - Grant/Research Support: GIlead Sciences, Merck, Janssen,<br />

AbbVie, Novartis<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

Andrew J. Muir - Advisory Committees or Review Panels: BMS, Gilead, Janssen,<br />

Merck; Consulting: Theravance; Grant/Research Support: Abbvie, Abbvie, BMS,<br />

Gilead, Janssen, Merck, Achillion, Lumena<br />

The following authors have nothing to disclose: Bertus Eksteen, Cynthia Levy,<br />

Aldo J. Montano-Loza, Dora Ding, Roger W. Chapman<br />

617<br />

Modeling biliary fluid dynamics reveals possible mechanism<br />

for dose-response and personalization of UDCA<br />

treatment in PSC<br />

Oleksandr Ostrenko 1 , Fabian Segovia-Miranda 2 , Mario Brosch 6 ,<br />

Wiebke Erhart 5 , Kirstin Meyer 2 , Georg Kretzschmar 3 , Christoph<br />

Jüngst 4 , Frank Lammert 4 , Marino Zerial 2 , Clemens Schafmayer 5 ,<br />

Lutz Brusch 1 , Jochen Hampe 6 ; 1 Center for Information Services<br />

and High Performance Computing, Technische Universität Dresden,<br />

Dresden, Germany; 2 Max Planck Institute of Molecular Cell<br />

Biology and Genetics, Dresden, Germany; 3 Molecular Cell Physiology<br />

and Endocrinology, Institute of Zoology, Technische Universität<br />

Dresden, Dresden, Germany; 4 Saarland University Medical<br />

Center, Homburg, Germany; 5 University Hospital Schleswig Holstein,<br />

Kiel, Germany; 6 TU Dresden, University Hospital Dresden,<br />

Dresden, Germany<br />

Background: Primary sclerosing cholangitis (PSC) is a progressive<br />

liver disease characterized by fibroobliterative destruction<br />

of the intra- and/or extra-hepatic bile ducts. Combined with<br />

immune-mediated insults to the bile duct, biliary flow obstruction<br />

might lead to pressure damage to the biliary epithelium<br />

and may drive further disease progression. Currently, the only<br />

and controversial medical treatment is the choleretic drug<br />

ursodeoxycholic acid (UDCA). However, while moderate doses<br />

(10-15mg/kg/day) of UDCA might improve liver function tests<br />

and histology, high-dose UDCA has been demonstrated to<br />

increase mortality in a RCT, thereby calling UDCA treatment<br />

in question. Aims: Develop a hydrodynamic model of biliary<br />

pressure and flow to assess the effect of UDCA on biliary pressure<br />

in normal liver and PSC. Methods: A recently developed<br />

theory of bile secretion and transport was applied to UDCA<br />

treatment. UDCA reduces bile viscosity (1) and increases solute<br />

concentration and consequently osmotic water influx into bile


516A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(2). Model parametrization was completed using measured<br />

viscosity from gallbladder bile of patients treated with UDCA<br />

and quantitative 3D-reconstruction of stacked high resolution<br />

microscopy of healthy human and PSC livers. Bile viscosity and<br />

osmotic water inflow are modelled as functions of UDCA dose<br />

according to effects (1) and (2) above. Results: This quantitative<br />

model allows the assessment of the mechanical consequences<br />

of UDCA treatment in a dose-dependent manner and predicts<br />

intrahepatic fluid pressure as a function of the physical properties<br />

of bile and the geometrical properties of the tissue. For<br />

healthy liver, the model reproduces the insensitivity of biliary<br />

pressure to UDCA dose. In PSC, a pronounced dose-response<br />

curve is predicted by our model: biliary pressure is decreasing<br />

at UDCA dose increments, reaches a minimum and then rapidly<br />

increases beyond baseline upon further dose increments. The<br />

optimal UDCA dose corresponding to the maximum pressure<br />

reduction can be quantified as a function of the physical and<br />

geometrical parameters that may vary from patient to patient.<br />

Conclusions: This study provides a hydrodynamic explanation<br />

of the clinical dose-response of UDCA in PSC. Further, it confirms<br />

tolerance to high doses of the compound in normal liver.<br />

Mechanistically, the UDCA-induced increase in water inflow<br />

has pressure-lowering (bile dilution) and pressure-increasing<br />

(increased fluid volume in need of drainage) consequences.<br />

Thus, the model advocates individualized dose-optimization<br />

based on the patient-specific biliary micro-geometry, thereby<br />

potentially rendering UDCA treatment safer and more widely<br />

applicable.<br />

Disclosures:<br />

The following authors have nothing to disclose: Oleksandr Ostrenko, Fabian<br />

Segovia-Miranda, Mario Brosch, Wiebke Erhart, Kirstin Meyer, Georg Kretzschmar,<br />

Christoph Jüngst, Frank Lammert, Marino Zerial, Clemens Schafmayer, Lutz<br />

Brusch, Jochen Hampe<br />

618<br />

In PSC with dominant bile duct stenosis, multi-resistant<br />

bacteriobilia is associated with reduced survival<br />

Christian Rupp 1 , Hannah Salzer 1 , Konrad A. Bode 2 , Karl Heinz<br />

Weiss 1 , Wolfgang Stremmel 1 , Peter Sauer 1 , Daniel Gotthardt 1 ;<br />

1 Internal Medicine IV, University Hopsital Heidelberg, Heidelberg,<br />

Germany; 2 Department of Infectious Diseases, University of Heidelberg,<br />

Heidelberg, Germany<br />

Background & Aims: Primary sclerosing cholangitis (PSC) is<br />

a chronic cholestatic liver disease, characterized by sclerosis<br />

and destruction of the biliary system. The course of diseases is<br />

often complicated by biliary infections. We aimed to analyze<br />

the frequency and influence of cholangitis with multi-resistant<br />

bacteria in PSC patients. Methods: Patients who were admitted<br />

to our Department during the time period of 1987 and<br />

2014 with well-defined diagnosis of PSC were analyzed in<br />

regard to multi-resistant bacteria in bile cultures. Comparison<br />

between frequencies was done using Chi-Square-test or Fisher’s<br />

exact test where appropriate. Continuous data were compared<br />

with the nonparametric Wilcoxon rank-sum test. Actuarial<br />

transplantation-free survival was estimated using Kaplan-Meier<br />

product limit estimator. Differences between the actuarial estimates<br />

were tested with the log rank test. Results: We identified<br />

20/244 (8.3) patients with multi-resistant bacteriobilia.<br />

8 patients had vancomycin resistant enterococcus (VRE) and<br />

12 had multi-resistant gram negative bacteria (MRGN) in bile<br />

culture. Baseline characteristic including laboratory values<br />

(AST, ALT, GGT, AP, Bilirubin), age at onset of PSC, concomitant<br />

inflammatory bowel disease, overlap with autoimmune<br />

hepatitis, dominant stenosis and Mayo Risk Score were not<br />

different between patients with or without MR bacteria. There<br />

was no difference in frequency of MR bacteriobilia between<br />

patients with or without DS (14/159, 8.8% vs. 5/66, 7.6%;<br />

p=0.8). We performed Kaplan-Meier analysis showing a markedly<br />

reduced transplantation-free survival in PSC patients with<br />

MR bacteriobilia (17.2 vs. 11.0 years; p=0.014). Stratification<br />

for presence of DS revealed reduced survival in presence<br />

of MR bacteriobilia only in patients with DS (16.7 vs. 9.9<br />

years; p=0.001), whereas no influence was detectable in PSC<br />

patients without DS (21.0 vs. 17.8 years; p=0.8). In multivariate<br />

analysis, including gender, age, AIHOL, IBD, DS, number<br />

of endoscopic interventions, MRS and MR bacteriobilia, only<br />

IBD (HR 3.3, 95% CI 1.1-11.1; p=0.04), MRS (HR 1.5, 95%<br />

CI 1.1-2.1; p=0.04) and MR bacteriobilia (HR 3.2, 95% CI<br />

1.1-9.5; p=0.03) were independent risk factors for reduced<br />

transplantation-free survival. Conclusion: In PSC patients with<br />

dominant stenosis MR bacteriobilia is associated with reduced<br />

survival, independent of number of endoscopic interventions.<br />

Multi-resistant bacteria may play a role in the progression of<br />

PSC. Optimal treatment strategies need to be established in<br />

order to achieve eradication of MR bacteriobilia.<br />

Disclosures:<br />

The following authors have nothing to disclose: Christian Rupp, Hannah Salzer,<br />

Konrad A. Bode, Karl Heinz Weiss, Wolfgang Stremmel, Peter Sauer, Daniel<br />

Gotthardt<br />

619<br />

Innate Immunity Drives the Initiation of Autoimmune<br />

Cholangitis in a Xenobiotic Murine Model of Primary<br />

Biliary Cirrhosis<br />

Chao-Hsuan Chang 2 , Ying-chun Chen 2 , Weici Zhang 1 , Patrick S.<br />

Leung 1 , M. Eric Gershwin 1 , Ya-Hui Chuang 2 ; 1 Internal Medicine,<br />

University of California, Davis, CA; 2 Department of Clinical Laboratory<br />

Sciences and Medical Biotechnology, National Taiwan<br />

University, Taipei, Taiwan<br />

Background: Invariant natural killer T (iNKT) cells are important<br />

in bridging innate and adaptive immunity by engaging<br />

with glycolipids presented by CD1d. α-galactosylceramide<br />

(α-GalCer) is a specific ligand for iNKT cell activation. Modification<br />

of α-GalCer lipid chain results in the generation of<br />

glycolipids with predominant Th1 or Th2 cytokine skewing<br />

profiles. (2s,3s,4r)-1-O-(α- D<br />

-galactopyranosyl)-N-tetracosanoyl-2-amino-1,3,4-nonanetriol<br />

(OCH) is a synthetic analog<br />

of α-GalCer which stimulates iNKT cells to predominantly Th2.<br />

Previous work in our group has demonstrated that iNKT cell<br />

activation by α-GalCer resulted in profound exacerbation of<br />

portal inflammation, bile duct damage, and hepatic fibrosis<br />

in the 2 octynoate-BSA (2OA-BSA) xenobiotic murine model<br />

of PBC. Here, we examined whether iNKT cells activated by<br />

a Th2-biasing agonist OCH, can influence the development of<br />

PBC in 2OA-BSA immunized mice. Methods: Groups of mice<br />

were treated with either OCH or α-GalCer and serially followed<br />

for cytokine production, markers of T cell activation,<br />

liver histopathology and anti-mitochondrial antibodies (AMA).<br />

In addition, groups of CD1d deleted mice and wild type mice<br />

were immunized with 2OA-BSA and monitored for liver infiltrates<br />

including CD4+ T cells, CD8+ T cells, NKT, NK and B<br />

cells. IFN-γ production by liver mononuclear cells were also<br />

compared. Results: 2OA-BSA mice treated with OCH exhibited<br />

a Th2 biased cytokine; α-GalCer initiated a rapid IL-4<br />

and prolonged IFN-γ production, whereas OCH induced predominant<br />

IL-4 production. Both OCH and α-GalCer induced<br />

high levels of AMA in 2OA-BSA immunized mice and had<br />

exacerbated portal inflammation and hepatic fibrosis. However,<br />

at 12 weeks post immunization, 2-OA-BSA/OCH immunized<br />

mice had significantly higher liver total mononuclear<br />

cell infiltrates, increased numbers of T, B and NK cells, and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 517A<br />

increased CD4 + and CD8 + T cells compared to 2-OA-BSA/PBS<br />

immunized mice. The frequencies of CD44 expressing CD8 +<br />

T cells and CD69 expressing CD8 + T cells were significantly<br />

increased in 2-OA-BSA/OCH immunized mice compared to<br />

2-OA-BSA/PBS immunized mice. Moreover, the frequency of<br />

CD44 expressing CD4 + T cells was significantly increased in<br />

2-OA-BSA/OCH immunized mice. Furthermore, the levels of<br />

AMA, cell infiltrates, and IFN-γ production of liver mononuclear<br />

cells were decreased in CD1d -/- mice when compared with<br />

wild type mice. Conclusion: Activation of iNKT cells is critical<br />

for the natural history of autoimmune cholangitis in this model<br />

and highlights the role of innate immunity in human PBC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chao-Hsuan Chang, Ying-chun<br />

Chen, Weici Zhang, Patrick S. Leung, M. Eric Gershwin, Ya-Hui Chuang<br />

620<br />

Serum metabolomic profile of patients with primary<br />

biliary cirrhosis treated with bezafibrate and changes<br />

following itching relief<br />

Albert Pares 1 , Anna Reig 1 , Miriam Perez-Cormenzana 2 , Rebeca<br />

Mayo 2 , Pilar Sese 1 , Azucena Castro 2 ; 1 Liver Unit, Hospital Clinic,<br />

University of Barcelona, IDIBAPS, CIBERehd, Barcelona, Spain;<br />

2 OWL, Derio, Spain<br />

Background and aims: Long-term fibrate treatment is an effective<br />

therapy for patients with primary biliary cirrhosis with<br />

suboptimal biochemical response to ursodeoxycholic acid.<br />

Bezafibrate therapy has further effects, thus decreasing pruritus<br />

in this cholestatic condition. The metabolomic consequences<br />

of fibrates in PBC are not known, and some <strong>studies</strong> indicate<br />

that their action may result from being a PPAR alpha agonist<br />

and also increasing the expression of some bile acid transporters.<br />

Therefore, we have assessed the metabolomic profiling of<br />

patients under bezafibrate, and particularly in those with pruritus,<br />

to define the potential mechanisms of action of bezafibrate<br />

in this cholestatic disease. Patients and Methods: Serum samples<br />

before and after bezafibrate therapy were taken from 29<br />

patients with PBC. Moreover in 14 of these patients with pruritus,<br />

samples were obtained before, during and after bezafibrate<br />

discontinuation. Metabolite extraction was accomplished<br />

by fractionating the samples into pools of species with similar<br />

physicochemical properties. Three different UPLC-MS analytical<br />

platforms were used for the analysis of fatty acyls, bile acids,<br />

steroids and lysoglycerophospholipids; amino acids; glycerolipids,<br />

glycerophospholipids, sterol lipids and sphingolipids.<br />

Results: More than 530 metabolites were identified. Regarding<br />

individual species 93 metabolites changed during bezafibrate<br />

therapy: PE(0:0/16:1) and ChoE(16:1) were in higher concentrations,<br />

while PC(17:0/18:2), PC(17:0/0:0), PC(15:0/22:6),<br />

PC(18:3/18:3) and PC(18:0/20:4) decreased with bezafibrate.<br />

Pruritus disappeared or was minimized by bezafibrate.<br />

In these patients 38 metabolites decreased (p


518A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

iary expression of SCT/SR/CFTR/AE2 compared to WT. SCT<br />

levels in biliary supernatants was decreased, but increased in<br />

serum in dnTGFβRII mice compared to WT, suggesting that SCT<br />

comes from other sources such as S cells. There was increased<br />

fibrosis and SCT/SR/CFTR/AE2 in early stage PBC samples,<br />

but decreased expression of SCT/SR/CFTR/AE2 in advanced<br />

PBC samples compared to controls. SCT levels increased in<br />

serum but decreased in bile in PBC compared to control. Conclusion:<br />

Loss of biliary bicarbonate secretion may contribute<br />

to disease progression, which may be partly counteracted by<br />

peripheral sources of SCT.<br />

Disclosures:<br />

The following authors have nothing to disclose: Lindsey Kennedy, Shannon S.<br />

Glaser, Francesca Bernuzzi, Fanyin Meng, Julie Venter, Heather L. Francis,<br />

Domenico Alvaro, M. Eric Gershwin, Antonio Franchitto, Paolo Onori, Sharon<br />

DeMorrow, Marco Marzioni, Eugenio Gaudio, Pietro Invernizzi, Gianfranco<br />

Alpini<br />

622<br />

Colectomy is associated with development of hepatobiliary<br />

malignancy but not cholangiocarcinoma in patients<br />

with Primary Sclerosing Cholangitis.<br />

Kate D. Williamson 1,2 , Ladislav Kozak 3 , Said Al Mamari 1 , John<br />

Halliday 1,2 , Roger W. Chapman 1,2 ; 1 Translational Gastroenterology<br />

Unit, John Radcliffe Hospital, Oxford, United Kingdom; 2 Nuffield<br />

Department of Medicine, University of Oxford, Oxford, United<br />

Kingdom; 3 Department of Sociology, University of Oxford, Oxford,<br />

United Kingdom<br />

Introduction: Primary Sclerosing Cholangitis (PSC) is associated<br />

with increased risk of malignancy – including hepatobiliary<br />

malignancies and colorectal cancer. Colectomy is often<br />

performed in patients due to either active inflammatory bowel<br />

disease (IBD) or development of dysplasia or colorectal cancer.<br />

Recently, the Mayo Clinic in US presented an abstract demonstrating<br />

an association between colectomy in PSC patients and<br />

the development of cholangiocarcinoma (CCA). We sought to<br />

validate this association. Methods: A database of 272 patients<br />

with PSC exists at John Radcliffe Hospital, Oxford UK, containing<br />

information on baseline demographics and clinical course.<br />

We performed a Cox regression analysis to identify variables<br />

associated with development of firstly CCA, and secondly any<br />

hepatobiliary malignancy (hepatocellular cancer, gallbladder<br />

cancer, other cancers arising within the liver, and cholangiocarcinoma).<br />

Variables examined were age at PSC diagnosis,<br />

gender, presence of IBD, and whether the patient underwent<br />

colectomy (including whether the indication was for active disease<br />

or for a composite of dysplasia or colorectal cancer).<br />

Results: Data were available for 228 patients. Median age at<br />

PSC diagnosis was 46.3 years (12.5 – 83.2y), and 61% were<br />

male. 4/228 (1.8%) had PSC/overlap with autoimmune hepatitis<br />

and were included in the analysis. The median follow up<br />

was 8.9 years (0.1 – 29.9y). 13/228 (5.7%) had a diagnosis<br />

of CCA, whilst 6 had a diagnosis of an alternative hepatobiliary<br />

malignancy, hence 19/228 (8.3%) had a diagnosis of any<br />

hepatobiliary malignancy. 18/19 patients with any hepatobiliary<br />

malignancy diagnosis had died at time of last follow up.<br />

18% patients had undergone a colectomy, the indication being<br />

dysplasia or cancer in 15% cases. On multivariate Cox proportional<br />

hazard models, presence of IBD, age at PSC diagnosis<br />

and gender were not associated with development of CCA,<br />

nor the composite endpoint of any hepatobiliary malignancy.<br />

Likewise, colectomy was not associated with development of<br />

CCA (HR 0.71, 95%CI 0.15-3.35, p=ns). However, colectomy<br />

was associated with development of any hepatobiliary<br />

malignancy (HR of 70.39, 95% CI 4.00-1238.52, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 519A<br />

<strong>studies</strong> in PSC. Furthermore, ALP may be an important factor to<br />

include in prognostic modeling for PSC.<br />

Disclosures:<br />

Ulrich Beuers - Consulting: Intercept via University of Amsterdam, Novartis via<br />

University of Amsterdam; Grant/Research Support: Falk, Zambon; Speaking and<br />

Teaching: Falk Foundation, Gilead, Roche, Shire<br />

Cyriel Y. Ponsioen - Advisory Committees or Review Panels: Takeda; Consulting:<br />

AbbVIE; Grant/Research Support: AbbVIE, Schering Plough, Dr. Falk Pharma,<br />

Tramedico Netherlands, Takeda<br />

The following authors have nothing to disclose: Elisabeth M. de Vries, Junfeng<br />

Wang, Mariska M. Leeflang, Kirsten Boonstra, Ronald Geskus<br />

Diagnostic Performance of Serum Fibrosis Markers in PSC<br />

624<br />

Validation of serum fibrosis marker panels in patients<br />

with primary sclerosing cholangitis (PSC) in a randomized<br />

trial of simtuzumab<br />

Christopher L. Bowlus 1 , Keyur Patel 2 , Indra Neil Guha 3 , Roger W.<br />

Chapman 4 , Olivier Chazouillères 5 , Naga P. Chalasani 6 , John M.<br />

Vierling 7 , Robert P. Myers 8 , Dora Ding 8 , Raul E. Aguilar Schall 8 ,<br />

Mani Subramanian 8 , John G. McHutchison 8 , Andrew J. Muir 2 ,<br />

Zachary D. Goodman 9 , Cynthia Levy 10 ; 1 University of California at<br />

Davis, Sacramento, CA; 2 Duke Clinical Research Institute, Durham,<br />

NC; 3 NIHR Nottingham Digestive Diseases Biomedical Research<br />

Unit, Nottingham, United Kingdom; 4 Oxford University, Oxford,<br />

United Kingdom; 5 Hôpital Saint Antoine, Paris, France; 6 Indiana<br />

University, Indianapolis, IN; 7 Baylor College of Medicine, Houston,<br />

TX; 8 Gilead Sciences, Inc., Foster City, CA; 9 Inova Fairfax<br />

Hospital, Falls Church, VA; 10 University of Miami, Miami, FL<br />

Background: Our objective was to assess the diagnostic performance<br />

of serum fibrosis marker panels in patients with PSC.<br />

Methods: We calculated FibroTest, ELF, APRI, and FIB-4 and<br />

measured serum LOXL2 (sLOXL2) by immunoassay (VIDAS®<br />

LOXL2; bioMérieux, Marcy L’Etoile, France) in PSC patients<br />

enrolled in a phase 2b trial of simtuzumab. Liver fibrosis was<br />

staged according to the Ishak classification and their diagnostic<br />

performance for predicting bridging fibrosis (Ishak stages<br />

3-6 vs. 0-2) and cirrhosis (stages 5-6 vs. 0-4) was determined<br />

using AUROCs. Sensitivity analyses were conducted according<br />

to biopsy length (≥ vs.


520A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

trolled <strong>studies</strong> (2 Phase 2 trials and 1 Phase 3). Abnormal<br />

bilirubin was not a common feature in the individual <strong>studies</strong>,<br />

therefore, data from all three <strong>studies</strong> was pooled; since all<br />

<strong>studies</strong> included a 3-month point, integrated data was evaluated<br />

at 3 months. ALP was markedly elevated at baseline<br />

in patients with abnormal bilirubin and OCA treatment was<br />

associated with significant ALP reductions. Mean bilirubin levels<br />

increased in the placebo arm and decreased with OCA,<br />

although the difference was not statistically significant over<br />

this short 3-month timeframe. Further evaluation of a subset of<br />

subjects treated to 12 months showed a significant reduction<br />

vs. placebo (Mean±SE:-8.9±2.2 vs -0.7±1.7; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 521A<br />

28d with no exclusions for severity or treatment of PRU. Patients<br />

were stratified by presence of prurius (PRU) at BL using the 5-D<br />

Itch Scale (5D). PRU was measured by Visual Analogue Scale<br />

(VAS) and the 5D at BL and Days 14, 28 and 42. Subjects<br />

with a BL and Day 28 VAS (n=38) and/or 5D (n=41) were<br />

categorized as improved, no change or worsened based on a<br />

>20% change from BL at Day 28. Correlations between PRU<br />

and liver chemistries or serum BAs by VAS were assessed by<br />

Spearman correlation and regression analyses with a clinical<br />

anchor threshold of r >0.30 and p


522A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Hemant Shah - Advisory Committees or Review Panels: Boehringer-Ingelheim;<br />

Consulting: Hoffman La-Roche, Merck, Gilead; Speaking and Teaching: Vertex<br />

Andrew Mason - Advisory Committees or Review Panels: AbbVie, Novartis,<br />

Intercept; Grant/Research Support: Abbvie, Gilead, Astellas<br />

Richard Pencek - Employment: Intercept Pharmaceuticals; Stock Shareholder:<br />

Intercept Pharmaceuticals<br />

Tonya Marmon - Employment: Intercept Pharmaceuticals, Inc; Stock Shareholder:<br />

Intercept Pharmaceuticals, Inc<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

Roya Hooshmand-Rad - Employment: Intercept pharmaceuticals Inc.<br />

The following authors have nothing to disclose: Velimir A. Luketic<br />

629<br />

Thyroid dysfunction in primary biliary cirrhosis: a comparative<br />

study at two European centers.<br />

Chiara Mangini 1 , Ana Reig 2 , Irene Franceschet 1 , Nora Cazzagon<br />

1 , Lisa Perini 1 , Llorenç Caballeria 2 , Silvia Cocchio 3 , Vincenzo<br />

Baldo 3 , Albert Pares 2 , Annarosa Floreani 1 ; 1 Department of Surgery,<br />

Oncology and Gastroenterology, University of Padova,<br />

Padova, Italy; 2 Liver Unit, Hospital Clinic-IDIBAPS, University of<br />

Barcelona, Barcelona, Spain; 3 Department of Molecular Medicine,<br />

Laboratory of Public Health and Population Studies, University of<br />

Padova, Padova, Italy<br />

Primary Biliary Cirrhosis (PBC) is often associated with other<br />

autoimmune diseases, but there are few data about the influence<br />

of other diseases on the natural history of PBC. Thyroid<br />

diseases have mainly an autoimmune aetiology, are often diagnosed<br />

in women and can be associated with PBC. Aim: To<br />

analyze the association between PBC and thyroid diseases<br />

and the impact of these disorders on the natural history of<br />

PBC in two European centres. Methods: 921 PBC patients<br />

enrolled between 1969 and 2015 in Padova (376 patients,<br />

M:F=25:351, mean follow-up 115.2 ± 85.6 months) and Barcelona<br />

(545 patients, M:F=42:503, mean follow-up 135.0<br />

± 94.9 months) were considered. Data on histological stage<br />

at diagnosis, biochemical data, associated thyroid disorders,<br />

extrahepatic autoimmune conditions, clinical events including<br />

hepatic decompensation (ascites, spontaneous bacterial<br />

peritonitis, encephalopathy, variceal bleeding and HCC<br />

development) were recorded. Survival was analysed using<br />

Kaplan-Meier curves and Cox regression method. Results: 150<br />

patients (16.3 %) had a thyroid disorder: 94 patients (10.2 %)<br />

had Hashimoto’s thyroiditis, 15 (1.6 %) had Graves’ disease,<br />

22 (2.4 %) had a multi-nodular goitre, 7 (0.8 %) had thyroid<br />

cancer, 13 (1.4 %) had other thyroid disorders. The prevalence<br />

of different types of thyroid disease was similar in Padova and<br />

Barcelona, except for Graves’ thyroiditis and thyroid cancer<br />

that resulted more frequent in the Padova cohort (15.7% vs<br />

5.0% and 8.6% vs 1.3% respectively, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 523A<br />

64% with an average intensity of 1.7 + 0.9 (scale 1-4). Over<br />

90% of participants were willing to donate biospecimens and<br />

be contacted for clinical research. Conclusions: Participants in<br />

the PSC Partners Registry are similar to other cohorts but are<br />

majority female, more symptomatic, and are a resource for<br />

future research.<br />

for PBC, a clear and important unmet need persists, as corroborated<br />

by variable UDCA response. The ongoing unmet need<br />

for PSC is striking, despite widespread UDCA use, and in the<br />

face of rising rates of transplantation for other liver diseases.<br />

Table 1. Comparison of Confirmed versus Unconfirmed PSC<br />

Cases<br />

Disclosures:<br />

Gregory T. Everson - Advisory Committees or Review Panels: Roche/Genentech,<br />

Abbvie, Galectin, Boehringer-Ingelheim, Eisai, Bristol-Myers Squibb, HepC<br />

Connection, BioTest, Gilead, Merck; Board Membership: HepQuant LLC, PSC<br />

Partners, HepQuant LLC; Consulting: Abbvie, BMS, Gilead, Bristol-Myers Squibb;<br />

Grant/Research Support: Roche/Genentech, Pharmassett, Vertex, Abbvie, Bristol-Myers<br />

Squibb, Merck, Eisai, Conatus, PSC Partners, Vertex, Tibotec, GlobeImmune,<br />

Pfizer, Gilead; Management Position: HepQuant LLC, HepQuant LLC;<br />

Patent Held/Filed: Univ of Colorado; Speaking and Teaching: Abbvie, Gilead<br />

Christopher L. Bowlus - Advisory Committees or Review Panels: Gilead Sciences,<br />

Inc; Grant/Research Support: Gilead Sciences, Inc, Intercept Pharmaceuticals,<br />

Bristol Meyers Squibb, Takeda, Lumena, Merck; Speaking and Teaching: Gilead<br />

Sciences, Inc<br />

The following authors have nothing to disclose: Rachel Gomel, Ricky Safer, Jesse<br />

A. King, Estella M. Geraghty, Keith D. Lindor<br />

631<br />

Cholestatic liver disease and liver transplantation in the<br />

UK: a twenty year review<br />

Gwilym Webb 1,2 , James W. Ferguson 2 , David Jones 4 , James Neuberger<br />

2,3 , Gideon Hirschfield 1,2 ; 1 NIHR Centre for Liver Research,<br />

University of Birmingham, Birmingham, United Kingdom; 2 Hepatology,<br />

Queen Elizabeth Hospital, Birmingham, United Kingdom;<br />

3 NHS Blood and Transplant, Bristol, United Kingdom; 4 Newcastle<br />

University, Newcastle, United Kingdom<br />

Background: Primary biliary cirrhosis (PBC) and primary sclerosing<br />

cholangitis (PSC) represent major indications for orthotopic<br />

liver transplantation (OLT). Ursodeoxycholic acid (UDCA) has<br />

been prescribed in the UK for most patients with PBC and PSC,<br />

with consensus for efficacy only in PBC. Aim: To analyse transplant<br />

listing and transplantation for PBC and PSC over the last<br />

20 years of UK-wide practice. Results: We analysed national<br />

transplant registry data for 1995 to 2014 inclusive. Records for<br />

12,935 listings and 10,512 OLTs were reviewed. Listings for<br />

all indications increased from 494 to 976/year (21.2±2.3/<br />

year; r 2 =0.82; p


524A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

were immunoreactive for both markers with a wide range of<br />

expression (MMP9: median HS, 11.6; range 2.4-122; and<br />

p-P38: median HS, 54; range 0.9-185). The most intense<br />

MMP9 expression was in areas of portal inflammation and<br />

interface hepatitis where immunoreactivity was associated with<br />

macrophages or within the extracellular matrix. Also, random<br />

lobular macrophages and Kupffer cells were strongly MMP9<br />

positive irrespective of the extent of portal inflammation. MMP9<br />

expression was associated with bile canalicular cholestasis<br />

(P=0.001) and negatively with periductal fibrosis (P=0.03), but<br />

was not associated with Ishak stage, the grade of portal, lobular<br />

or interface inflammation, duct injury, CK7 staining, liver<br />

biochemistry, or UC. Within inflamed portal areas, nuclear<br />

p-P38 localized to small mononuclear cells and less frequently<br />

to reactive duct epithelial cells. p-P38 expression was not associated<br />

with any histologic feature, liver biochemistry, or UC.<br />

Conclusions: In PSC patients, hepatic MMP9 and p-P38 expression<br />

is increased. Additional study is necessary to validate<br />

that MMP9 and ASK1 pathway activation may contribute to<br />

disease pathogenesis.<br />

Disclosures:<br />

Dorothy French - Employment: Gilead Sciences; Stock Shareholder: Genentech<br />

- Roche<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

Erik G. Huntzicker - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Satyajit Karnik - Consulting: Monsoon Dx; Employment: Gilead Sciences<br />

Victoria Smith - Employment: Gilead Sciences Inc<br />

Raul E. Aguilar Schall - Employment: Gilead Sciences, Inc.<br />

Bittoo Kanwar - Employment: Gilead Sciences<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Andrew J. Muir - Advisory Committees or Review Panels: BMS, Gilead, Janssen,<br />

Merck; Consulting: Theravance; Grant/Research Support: Abbvie, Abbvie, BMS,<br />

Gilead, Janssen, Merck, Achillion, Lumena<br />

Robert P. Myers - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

The following authors have nothing to disclose: David Newstrom<br />

to the modified Amsterdam score (mAm score). Results: In total<br />

154 PSC patients with adequate bile sample were included.<br />

The mean age of patients was 40.9±12.7 years. The median<br />

(IQR) disease duration was 4 (3 - 9) years. Concomitant IBD<br />

was present at 71 %. The mean (±SD) S-ALP was 137±123<br />

IU/l (≤ 105 IU/l), S-bilirubin 14.6±12.8mmol/l, (≤ 20 IU/l.<br />

The mean UDCA dose was 17.1±4.4 mg/kg/day. UDCA<br />

therapy was associated with markedly increased proportion<br />

of UDCA in bile (49.1± 22.2 %). The presence or absence<br />

of IBD had no impact on biliary UDCA content. S-ALP IU/ml<br />

correlated positively with mAm score, p for linearity 0.005, but<br />

S-ALP could be < UNL even in advanced biliary disease. No<br />

correlation was found between UDCA dose and S-ALP level,<br />

r=0.13 (95%CI: -0.07-0.27). S- ALP level did not correlate with<br />

the UDCA content (mg%) in bile, fig 1. Conclusions. S-alkaline<br />

phosphatase in not an optional surrogate marker for disease<br />

progression or monitoring the treatment response of UDCA in<br />

PSC patients.<br />

Figure 1.<br />

Disclosures:<br />

The following authors have nothing to disclose: Martti A. Färkkilä, Kalle Jokelainen,<br />

Hannu Kautiainen<br />

633<br />

S-Alkaline phosphatase is not a reliable surrogate<br />

marker for ursodeoxycholic acid therapy in patients<br />

with primary sclerosing cholangitis<br />

Martti A. Färkkilä 1,2 , Kalle Jokelainen 2 , Hannu Kautiainen 1 ; 1 Helsinki<br />

University, Helsinki, Finland; 2 Gastroenterology, Helsinki University<br />

Hospital, Helsinki, Finland<br />

Background. Primary sclerosing cholangitis (PSC) is chronic<br />

cholestatic liver disease of unknown origin, leading to inflammation<br />

and strictures of intra- and extrahepatic bile ducts.<br />

Ursodeoxycholic acid (UDCA is widely used to treat PSC,<br />

demonstrating amelioration of cholestatic liver enzymes. S-ALP<br />

is the most used surrogate marker for treatment response for<br />

UDCA therapy. Objectives of the study.To analyze the effect<br />

of UDCA therapy 20mg/kg/day on biliary UDCA content and<br />

S-ALP-levels in PSC. Patients and methods. The study population<br />

consists of PSC patients on UDCA therapy (n= 154) and<br />

referred for ERCP for confirmation of the diagnosis of PSC or<br />

for biliary dysplasia surveillance. After cannulation the common<br />

bile duct a bile sample was aspirated using balloon catheter,<br />

immersed immediately in liquid nitrogen and stored in -20<br />

o C. The content of UDCA in bile was analyzed. The conten of<br />

UDCA was expressed as molar percentages (mM%) of total<br />

bile acids. Cholangiographic findings were scored according


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 525A<br />

634<br />

Elevation of alkaline phosphatase during follow-up is<br />

an early predictor of hyperbilirubinaemia and of clinical<br />

endpoints in primary biliary cirrhosis – an international<br />

study<br />

Willem J. Lammers 1 , Henk R. van Buuren 1 , Cyriel Y. Ponsioen 2 ,<br />

Harry L. Janssen 3 , Annarosa Floreani 4 , Gideon Hirschfield 5 ,<br />

Christophe Corpechot 6 , Marlyn J. Mayo 7 , Pietro Invernizzi 8 ,<br />

Pier Maria Battezzati 9 , Albert Pares 10 , Frederik Nevens 11 , Douglas<br />

Thorburn 12 , Andrew Mason 13 , Kris V. Kowdley 14 , Angela<br />

C. Cheung 3 , Teru Kumagi 15 , Palak J. Trivedi 5 , Raoul Poupon 6 ,<br />

Ana Lleo 8 , Llorenç Caballeria 10 , Keith D. Lindor 16,17 , Maren H.<br />

Harms 1 , Bettina E. Hansen 1 ; 1 Gastroenterology & Hepatology,<br />

Erasmus MC, University Medical Center, Rotterdam, Netherlands;<br />

2 Academic Medical Center, Amsterdam, Netherlands; 3 Toronto<br />

Western & General Hospital, Toronto, ON, Canada; 4 University<br />

of Padua, Padua, Italy; 5 University of Birmingham, Birmingham,<br />

United Kingdom; 6 Hôpital Saint-Antoine, APHP, Paris, France; 7 UT<br />

Southwestern Medical Center, Dallas, TX; 8 Humanitas Clinical and<br />

Research Center, Rozzano, Italy; 9 Università degli Studi di Milano,<br />

Milan, Italy; 10 , University of Barcelona, Barcelona, Spain; 11 University<br />

Hospitals Leuven, KULeuven, Leuven, Belgium; 12 The Royal<br />

Free Hospital, London, United Kingdom; 13 University of Alberta,<br />

Edmonton, AB, Canada; 14 Swedish Medical Center, Seattle, WA;<br />

15 Ehime University graduate School of Medicine, Ehim, Japan;<br />

16 Mayo Clinic, Rochester, MA; 17 Arizona State University, Phoenix,<br />

AZ<br />

Background and aim: Earlier reports have shown that patients<br />

with primary biliary cirrhosis (PBC) and low alkaline phosphatase<br />

(ALP) and bilirubin levels after 1 year of ursodeoxycholic<br />

acid (UDCA) treatment have favourable liver transplant-free survival<br />

. However, ALP and bilirubin may increase later during follow-up.<br />

The aim of the study was to identify patients within this<br />

group at risk of adverse outcome during follow-up. Methods:<br />

Patient data were obtained from the international Global PBC<br />

study group database, comprising long-term follow-up data<br />

of 4845 patients from 15 centers across North America and<br />

Europe. UDCA-treated patients with ALP ≤2.0xULN or bilirubin<br />

≤1.0xULN at 1 year follow-up were included in this analysis.<br />

Time-dependent Cox regressions analysis models was applied.<br />

Results: During a median follow-up of 7.11 years (IQR, 3.34-<br />

11.18) 350/2527 patients underwent liver transplantation or<br />

died. Patients who developed elevation of ALP (>2xULN) during<br />

follow-up had an increased risk of liver transplantation/death<br />

compared with those who remained ≤2.0xULN (HR 2.20, 95%<br />

CI 1.73-2.81, P value 1xULN) during follow-up were also at<br />

higher risk of reaching an endpoint (HR 5.51, IQR, 4.40-6.88,<br />

P value 2xULN, 25% had an<br />

event within 6.4 years; for elevated bilirubin levels this was<br />

1.5 years. Elevated ALP proofed to be an early predictor of<br />

abnormal bilirubin levels during follow-up; of patients with both<br />

normal bilirubin levels and ALP levels ≤2.0xULN 407 developed<br />

elevated bilirubin. Those patients who reached ALP levels<br />

>2.0xULN during follow-up were at higher risk of developing<br />

hyperbilirubinaemia (HR 3.79, 2.95-4.86, P value 2.0xULN) during<br />

follow-up is an early predictor of hyperbilirubinaemia, and an<br />

important early marker of a possible clinical event.<br />

Disclosures:<br />

Henk R. van Buuren - Grant/Research Support: Intercept, Zambon Nederland BV<br />

Cyriel Y. Ponsioen - Advisory Committees or Review Panels: Takeda; Consulting:<br />

AbbVIE; Grant/Research Support: AbbVIE, Schering Plough, Dr. Falk Pharma,<br />

Tramedico Netherlands, Takeda<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Marlyn J. Mayo - Grant/Research Support: Intercept, Salix, NGM, Lumena,<br />

Gilead<br />

Albert Pares - Consulting: Lumena Pharmaceuticals, Intercept Pharmaceuticals, Inc<br />

Frederik Nevens - Consulting: MSD, CAF, Intercept, Gore, BMS, Abbvie, Novartis,<br />

MSD, Eumedica, Janssen, Promethera Biosciences; Grant/Research Support:<br />

Ferring, Roche, Astellas, Novartis, Janssen-Cilag, Abbvie<br />

Andrew Mason - Advisory Committees or Review Panels: AbbVie, Novartis,<br />

Intercept; Grant/Research Support: Abbvie, Gilead, Astellas<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Palak J. Trivedi - Grant/Research Support: Wellcome Trust<br />

The following authors have nothing to disclose: Willem J. Lammers, Annarosa<br />

Floreani, Gideon Hirschfield, Christophe Corpechot, Pietro Invernizzi, Pier Maria<br />

Battezzati, Douglas Thorburn, Angela C. Cheung, Teru Kumagi, Raoul Poupon,<br />

Ana Lleo, Llorenç Caballeria, Keith D. Lindor, Maren H. Harms, Bettina E. Hansen<br />

635<br />

Evaluation of the Posology of Obeticholic Acid (OCA) in<br />

Patients with PBC<br />

Richard Pencek, Karen Lutz, Tonya Marmon, Leigh MacConell;<br />

Intercept Pharmaceuticals, San Diego, CA<br />

Background: Obeticholic acid (OCA) is a potent and selective<br />

FXR agonist developed for treatment of primary biliary cirrhosis<br />

(PBC). 216 patients were enrolled in the Phase 3, double-blind,<br />

placebo-controlled trial and randomized to: Placebo (PBO), 10<br />

mg OCA (maximally efficacious dose in PBC), or 5 mg OCA<br />

titrating to 10 mg after Month (M) 6 based on tolerability and<br />

clinical response. Titration of OCA at M6 mitigated tolerability<br />

(pruritus, the most common AE) without compromising efficacy.<br />

While both 5 mg and 10 mg resulted in clinically meaningful<br />

and significant reductions in ALP compared to PBO, efficacy<br />

was greater at the 10 mg dose. This analysis evaluated the tolerability<br />

and efficacy Phase 3 data to support titration of OCA<br />

earlier than M6 after initiating OCA treatment. Methods: Time<br />

to first onset of treatment-emergent pruritus included the number<br />

of patients with pruritus (first onset), without pruritus (censored),<br />

and the minimum and maximum times in days. Kaplan-Meier<br />

estimates were plotted as a “survival curve” for each treatment<br />

group. Liver biochemistry measures (ALP, total bilirubin, GGT,<br />

ALT, and AST) were evaluated by LS mean change over time<br />

during the 12-month treatment period. Results: A Kaplan-Meier<br />

survival curve demonstrated that median time to first pruritus<br />

event was within the initial month of starting OCA. The median<br />

time to first onset of any pruritus event occurred within 2 weeks<br />

(OCA 10 mg), M1 (OCA 5 mg), and M2 (PBO). Mean BL<br />

ALP (U/L) was 327 (PBO) and 316 (OCA 10 mg). Clinically<br />

meaningful and significant improvements in ALP were achieved<br />

within 3M and maintained through the 12M period (Figure).<br />

Similar trends were observed in other liver biochemistry markers.<br />

Conclusions: The timing of onset of pruritus is early for both<br />

OCA doses. The majority of efficacy is observed within the<br />

first 3M of initiating OCA therapy. Thus, 3M after initiation of<br />

OCA is a reasonable timeframe to assess if a PBC patient may<br />

benefit from uptitration of OCA 5 mg to 10 mg.


526A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

around these tools particularly amongst non-specialists. Implementing<br />

a stratified approach to management requires these<br />

gaps to be addressed.<br />

Disclosures:<br />

David Jones - Consulting: Intercept, Pfizer, Novartis; Speaking and Teaching:<br />

Falk, Shire<br />

Gideon Hirschfield - Advisory Committees or Review Panels: Intercept Pharma;<br />

Consulting: Dignity Sciences, GSK, NGM Bio, Lumena, J & J; Grant/Research<br />

Support: BioTie; Speaking and Teaching: Falk Pharma<br />

The following authors have nothing to disclose: Margaret Corrigan, Luke Vale,<br />

Diarmuid Coughlan<br />

Disclosures:<br />

Richard Pencek - Employment: Intercept Pharmaceuticals; Stock Shareholder:<br />

Intercept Pharmaceuticals<br />

Tonya Marmon - Employment: Intercept Pharmaceuticals, Inc; Stock Shareholder:<br />

Intercept Pharmaceuticals, Inc<br />

The following authors have nothing to disclose: Karen Lutz, Leigh MacConell<br />

636<br />

Clinician confidence in stratifying risk in primary biliary<br />

cirrhosis – a UK-PBC survey<br />

Margaret Corrigan 1 , Luke Vale 2 , Diarmuid Coughlan 2 , David<br />

Jones 3 , Gideon Hirschfield 1 ; 1 NIHR Liver Biomedical Research<br />

Unit, University of Birmingham, Birmingham, United Kingdom;<br />

2 Institute of Health and Society, Newcastle University, Newcastle,<br />

United Kingdom; 3 Institute of Cellular Medicine, Newcastle Uiniversity,<br />

Newcastle, United Kingdom<br />

Background: Primary biliary cirrhosis (PBC) has only one present<br />

licensed therapy, ursodeoxycholic acid (UDCA). Future<br />

therapies will be offered to patients at high risk of disease progression.<br />

Assessment of biochemical response to UDCA allows<br />

identification of high risk patients. Aim: To survey current<br />

understanding of UDCA response criteria in the UK. Methods:<br />

A survey of current clinical practice was created by UK-PBC<br />

and distributed to clinicians via the British Society of Gastroenterology<br />

(BSG) and British Association for the Study of the<br />

Liver (BASL) mailing lists and newsletters. 206 responses were<br />

received from 1900 invites. Questions covered diagnosis and<br />

management of the condition with four questions specifically<br />

covering UDCA response assessment. Results: Respondents<br />

came from a variety of clinical backgrounds - consultant hepatologists<br />

in tertiary centres – 14 (7%), consultant hepatologists<br />

in non tertiary centres – 32 (15.5%), consultant gastroenterologists<br />

– 75 (36.4%), trainees – 78 (37.9%), others including specialist<br />

nurses - 7 (3.4%). Whilst 90% of respondents reported<br />

routine use of UDCA in clinical practice, only 20% reported<br />

always assessing UDCA response once patients have been<br />

on treatment for 12 months whilst 50% never assess response.<br />

Looking at rates of assessment of UDCA response between the<br />

specialist groups: 64% of gastroenterologists and 47% trainees<br />

never assess response compared to 25% of non-tertiary<br />

hepatologists and 14% of tertiary hepatologists. The number of<br />

patients seen appeared to affect UDCA response assessment:<br />

64% of those who saw fewer than 10 patients/year never<br />

assess response compared to 10% of those who saw more than<br />

50 patients/year. 40% of respondents reported themselves as<br />

‘not all confident’ in assessing response with 58% stating they<br />

were unaware that criteria were available and 27% unsure<br />

which criteria were best to use. Conclusion: The majority of<br />

patients with PBC are cared for in non-tertiary centres and most<br />

are managed by non-specialist clinicians. The application of<br />

emerging therapies for patients with PBC requires appropriate<br />

use of risk stratification tools in routine clinical practice.<br />

Our results demonstrates gaps in knowledge and confidence<br />

637<br />

The IgG/IgG4 mRNA ratio by quantitative PCR accurately<br />

diagnoses IgG4-related disease and predicts<br />

treatment response<br />

Lowiek M. Hubers 1 , Marieke E. Doorenspleet 2,3 , Emma L. Culver<br />

4,5 , Lucas Maillette de Buy Wenniger 1 , Paul L. Klarenbeek 2,3 ,<br />

Roger W. Chapman 4,5 , Stan F. van de Graaf 1 , Joanne Verheij<br />

6 , Thomas van Gulik 7 , Frank Baas 3 , Eleanor Barnes 4,5 , Niek<br />

de Vries 2 , Ulrich Beuers 1 ; 1 Gastroenterology & Hepatology and<br />

Tytgat Institute of Liver and Intestinal Research, Academic Medical<br />

Center, Amsterdam, Netherlands; 2 Clinical Immunology &<br />

Rheumatology and Amsterdam Rheumatology and Immunology<br />

Center, Academic Medical Center, Amsterdam, Netherlands;<br />

3 Genome Analysis, Academic Medical Center, Amsterdam, Netherlands;<br />

4 Translational Gastroenterology Unit, John Radcliffe Hospital,<br />

Oxford, United Kingdom; 5 NDM Oxford University, Peter<br />

Medawar, Oxford University, Oxford, United Kingdom; 6 Pathology,<br />

Academic Medical Center, Amsterdam, Netherlands; 7 Surgery,<br />

Academic Medical Center, Amsterdam, Netherlands<br />

Introduction: IgG4-associated cholangitis (IAC) and autoimmune<br />

pancreatitis (AIP) are major manifestations of IgG4-related<br />

disease (IgG4-RD). Misdiagnosis and inadequate<br />

treatment are common since IAC and AIP mimic other inflammatory<br />

and malignant pancreatobiliary diseases, and accurate<br />

diagnostic biomarkers are lacking. Moreover, since relapse<br />

after tapering of immunosuppressive therapy occurs in 50%<br />

of patients, there is a need for biomarkers monitoring disease<br />

activity. Recently, using Next-Generation Sequencing,<br />

we observed that dominant IgG4+ B-cell receptor clones in<br />

peripheral blood distinguish patients with active IAC/AIP<br />

from primary sclerosing cholangitis (PSC) and pancreatobiliary<br />

malignancies (CA) [Hepatology 2013;57:2340]. Here,<br />

we report on a simple quantitative PCR (qPCR) protocol for<br />

diagnosing IAC/AIP and monitoring disease activity. Patients<br />

and Methods: 15 patients with IAC and/or AIP according to<br />

HISORt criteria, 7 patients with PSC and 8 with CA formed the<br />

test cohort. Intra- and extramural replication cohorts consisted<br />

of 16 IAC/AIP, 5 PSC and 13 CA patients (Dutch cohort), and<br />

8 IAC/AIP and 8 PSC patients (British cohort). In 20 Dutch<br />

IAC/AIP patients, follow-up samples after 4 and 8 weeks of<br />

corticosteroid therapy were available. RNA was isolated and<br />

the constant region of the B-cell receptor was amplified using a<br />

generic forward IgG primer together with either a generic IgG<br />

or a IgG4-specific reverse primer. The ratio total IgG/IgG4<br />

mRNA was calculated and expressed as ΔC T<br />

. Results: ΔC T<br />

as<br />

measure of IgG/IgG4 mRNA expression in peripheral blood<br />

of the test cohort was 2.8±1.1 (mean+SD) in IAC/AIP patients,<br />

compared to 6.8±1.6 in PSC and 7.6±1.4 in CA (Figure 1A,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 527A<br />

sion: IgG4-RD of the biliary tree and pancreas can be accurately<br />

distinguished from PSC or pancreatobiliary malignancies<br />

by ΔC T<br />

based on an affordable qPCR test. ΔC T<br />

can also be<br />

used as a marker for treatment response and disease activity<br />

in IAC and AIP.<br />

Figure 1.<br />

Disclosures:<br />

Ulrich Beuers - Consulting: Intercept via University of Amsterdam, Novartis via<br />

University of Amsterdam; Grant/Research Support: Falk, Zambon; Speaking and<br />

Teaching: Falk Foundation, Gilead, Roche, Shire<br />

The following authors have nothing to disclose: Lowiek M. Hubers, Marieke E.<br />

Doorenspleet, Emma L. Culver, Lucas Maillette de Buy Wenniger, Paul L. Klarenbeek,<br />

Roger W. Chapman, Stan F. van de Graaf, Joanne Verheij, Thomas van<br />

Gulik, Frank Baas, Eleanor Barnes, Niek de Vries<br />

638<br />

Emperipolesis mediated by CD8+ T cells correlated with<br />

biliary epithelia cell injury in primary biliary cirrhosis<br />

Suxian Zhao, Rong-qi Wang, Yuguo Zhang, Huijuan Du, Yuemin<br />

Nan; Third Hospital of Hebei Medical University, Shijiazhuang,<br />

China<br />

Introduction: Primary biliary cirrhosis (PBC) is thought to be an<br />

autoimmune disease characterized by chronic destruction of<br />

bile ducts. A major unanswered question regarding the pathogenesis<br />

of PBC is the precise mechanisms of small duct injury<br />

and destruction. Emperipolesis is one of cell-in-cell structures<br />

that have been observed in chronic viral hepatitis. However,<br />

it remains unknown whether emperipolesis is involved in biliary<br />

epithelia cell injury and it diagnostic value in PBC. Aims:<br />

To clarify the pathogenesis and histological characteristics of<br />

emperipolesis in the liver injury of PBC. Methods:Sixty PBC<br />

patients, diagnosed by liver biopsy, were divided into two<br />

groups, early PBC (stages I and IIn=40) and late PBC (stages III<br />

and IV, n=20). The degrees of hepatic inflammation and fibrosis<br />

were assessed with Scheuer Scoring System. Emperipolesis<br />

was observed in liver sections stained with hematoxylin-eosin.<br />

The expression of CK19 (cholangiocyte), CD8 (T cell), CD20 (B<br />

cell), CD56 (NK cell), CD68 (macrophage cell) and MPO (neutrophil)<br />

was evaluated by mmunofluorescence double labeling<br />

combined with confocal microscopic observation. The apopotosis<br />

of BECs was detected by TUNEL assay under laser confocal<br />

microscope. Results: Emperipolesis was observed in 73.3%<br />

of patients with PBC, and BECs were predominantly host cells.<br />

In PBC patients with emperipolesis, the number of emperipolesis<br />

structures correlated with lymphoid follicles (r=0.4394,<br />

P


528A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Guo-Xiang Yang, Takashi Tomiyama,<br />

Ying Sun, Weici Zhang, Patrick S. Leung, Xiao-Song He, Sandeep S.<br />

Dhaliwal, M. Eric Gershwin<br />

640<br />

Natural killer cells regulate T cell immunity in primary<br />

biliary cirrhosis<br />

Shinji Shimoda 1 , Minoru Nakamura 2 , M. Eric Gershwin 3 ; 1 Kyushu<br />

University, Fukuoka, Japan; 2 Nagasaki University, Nagasaki,<br />

Japan; 3 UC Davis, Davis, CA<br />

BACKGROUND: Dissection of the multi-lineage response that<br />

leads to biliary epithelial destruction is essential for future therapeutic<br />

efforts. Indeed, a variety of work has already reflected<br />

that there are autoreactive T and B cells that do target biliary<br />

epithelial cells (BEC), but there is also an emerging set of data<br />

that suggests innate immunity is critical not only for the initiation,<br />

but also for a variety of interactions that alter the natural<br />

history of cholangitis. Our previous work has demonstrated that<br />

biliary cell cytotoxity is dependent on the initiation of innate<br />

responses followed by chronic adaptive as well as bystander<br />

mechanisms which are the interactions between natural killer<br />

(NK) cells and BEC. OBJECTIVE: We have taken advantage<br />

of our ability to isolate NK, BEC and endothelial cells from<br />

explanted liver samples, and studied the interactions between<br />

NK cells and BEC and focused on the mechanisms that activated<br />

autoreactive T cells in the presence or absence of antigen<br />

presenting cells (APC) and their dependence on IFN-g as<br />

well as expression of BEC HLA class I and class II molecules.<br />

Explanted liver and spleen from a total of 7 patients included<br />

2 with PBC and 5 with hepatitis C virus infection were used.<br />

There were no significant differences in the results derived from<br />

the liver mononuclear cells or the BEC population from either<br />

PBC or hepatitis C and ultimately all data were combined. This<br />

is consistent with our thesis that the BEC is an innocent victim<br />

as well as the established data that PBC reoccurs following<br />

transplantation. RESULTS: At a high NK/BEC ratio, NK cells<br />

attack autologous BEC in either a PBC-specific nor auto antigen-specific<br />

manner, and as glutathiolation does not occur in<br />

apoptotic BEC, autoantigen PDC-E2 that exists in the mitochondrial<br />

membrane is not degenerated by caspases in PBC, PDC-<br />

E2 from lysed BEC activate autoreactive CD4 positive T cells<br />

in the presence of APC. In contrast, at a low NK/BEC ratio,<br />

BEC were not lysed whereas IFN-g production was induced<br />

from NK cells, which in turn induced expression of HLA class<br />

I and II molecules on BEC and protected them from lysis upon<br />

exposure to autoreactive NK cells. In addition, IFN-g secretion<br />

from NK cells after exposure to autologous BEC enabled autoreactive<br />

CD4 positive T cells to lyse BEC in the presence of<br />

PDC-E2 antigen. CONCLUSION: Our data suggest that NK cell<br />

mediated innate immune responses are critical for the breach<br />

of tolerance at the initial stage of PBC, but also maintained the<br />

cytotoxic effect of autoantigen-specific T cells, which are critical<br />

for disease progression.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shinji Shimoda, Minoru<br />

Nakamura, M. Eric Gershwin<br />

641<br />

Combination Anti-Retroviral Therapy Provides Reduction<br />

in Human Betaretrovirus Load and Durable Biochemical<br />

Responses in Patients with Primary Biliary Cirrhosis<br />

Ellina Lytvyak, Aldo J. Montano-Loza, Lynora Saxinger, Andrew<br />

Mason; University of Alberta, Edmonton, AB, Canada<br />

Purpose/Background: A human betaretrovirus has been characterized<br />

in primary biliary cirrhosis (PBC). Reverse transcriptase<br />

inhibitors lamivudine and zidovudine have shown limited<br />

efficacy. Whereas combination lopinavir/ritonavir (LPRr) and<br />

tenofovir/emticitabine (TDF/FTC) has been reported to be<br />

efficacious in normalizing liver tests (Lancet 2011). Methods:<br />

PBC patients unresponsive to UDCA were randomized into<br />

a crossover study with daily TDF/FTC 300/200mg and LPRr<br />

800/200mg versus placebo for 6 months; followed by an<br />

open label study for a total of 24 months. HBRV DNA was<br />

assessed in PBMC using digital droplet PCR. Results: (A) In the<br />

6 month RCT, a significant reduction in Alk Phos was observed<br />

in patients on TDF/FTC and LPRr (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 529A<br />

642<br />

Long term impact of fibrates in primary sclerosing cholangitis<br />

with incomplete biochemical response to ursodeoxycholic<br />

acid : the French-Spanish experience<br />

Sara Lemoinne 2,1 , Christophe Corpechot 2,1 , Astrid Donald D.<br />

Kemgang Fankem 2,1 , Farid Gaouar 2 , Raoul Poupon 2,1 , Chantal<br />

Housset 2,1 , Albert Pares 3 , Olivier Chazouillères 2,1 ; 1 INSERM,<br />

UMRS_938, F-75012 Paris, France UPMC Univ Paris 06, F-75012<br />

Paris, France, Paris, France; 2 Service d’Hépatologie, Centre de<br />

Référence des Maladies Inflammatoires des Voies Biliaires, Hôpital<br />

Saint-Antoine, Assistance-Publique Hôpitaux de Paris, Paris,<br />

France; 3 Liver Unit, Hospital Clinic, University of Barcelona,<br />

IDIBAPS, CIBERehd, Barcelona, Spain<br />

Background : In patients with primary sclerosing cholangitis<br />

(PSC), ursodeoxycholic acid (UDCA) improves serum livers tests<br />

and some surrogate markers of prognosis. In the absence of<br />

medical treatment with proven efficacy, UDCA is widely used<br />

in European PSC patients. However, newer therapies are obviously<br />

needed. Fibrates, PPAR agonists that exert anti-inflammatory<br />

properties in several experimental models of autoimmunity,<br />

seem to have a beneficial effect on liver biochemistries in primary<br />

biliary cirrhosis. In 2014, we reported improvement of<br />

liver tests under fibrates in 15 French PSC patients with an<br />

incomplete biochemical response to UDCA. Aim : To confirm<br />

the safety and efficacy of fibrates in an extended PSC population<br />

including patients from another center. Methods : This<br />

retrospective study included patients with PSC treated with<br />

fibrates (fenofibrate 200mg/d or bezafibrate 400mg/d) for<br />

at least 6 months in addition to UDCA, after an incomplete<br />

biochemical response (ALP>1ULN) to UDCA (15-20mg/kg/d)<br />

for at least 1 year. Patients with associated liver diseases, especially<br />

auto-immune hepatitis were not included. Changes in biochemical<br />

parameters have been assessed using a linear mixed<br />

model and Wilcoxon tests. Results : 26 patients were included<br />

(19 from Paris and 7 from Barcelona) : 17 males, median age<br />

49.7 years, 16 with inflammatory bowel disease, median liver<br />

stiffness 10.6 kPa (corresponding to fibrosis ≥ F3). Median<br />

duration of treatment with fibrates was 36.8 months (6.7-<br />

61.3). Under treatment with fibrates, ALP and ALT decreased<br />

significantly (p= 0.04 and 0.01 respectively) and this decrease<br />

was maintained at 48 months of treatment. GGT tended to<br />

decrease (p=0.10). AST, total bilirubin and albumin remained<br />

unchanged. No serious adverse event related to fibrates<br />

occured. In 54% of patients, treatment by fibrates was stopped<br />

because of various reasons : development of biliary stones,<br />

gastro-intestinal bleeding, decision by cardiologist, liver tests<br />

worsening. Excluding this latter group, interruption of fibrates<br />

was followed by a significant increase of ALP. During follow-up<br />

of a median duration of 4.1 years, 2 cholangiocarcinoma, 5<br />

liver transplantations and 1 death were recorded. Interestingly,<br />

despite a biochemical improvement, liver stiffness assessed by<br />

transient elastography increased significantly under fibrates.<br />

Conclusion : Addition of fibrates induces a significant biochemical<br />

improvement in PSC patients with incomplete biochemical<br />

response to UDCA but seems to be insufficient to control the<br />

progression of advanced PSC. This raises again the critical<br />

issue of surrogate markers of prognosis in PSC clinical trials.<br />

Disclosures:<br />

Albert Pares - Consulting: Lumena Pharmaceuticals, Intercept Pharmaceuticals, Inc<br />

Olivier Chazouillères - Consulting: APTALIS, MAYOLY-SPINDLER<br />

The following authors have nothing to disclose: Sara Lemoinne, Christophe<br />

Corpechot, Astrid Donald D. Kemgang Fankem, Farid Gaouar, Raoul Poupon,<br />

Chantal Housset<br />

643<br />

Predictors of recurrence and subsequent outcomes for<br />

patients admitted with a sentinel vs recurrent episode of<br />

acute cholangitis<br />

Amanda M. Lynn 1 , James H. Tabibian 3 , Todd H. Baron 2 ; 1 Internal<br />

Medicine, Mayo Clinic, Rochester, MN; 2 Gastroenterology and<br />

Hepatology, University of North Carolina, Chapel Hill, NC; 3 Gastroenterology<br />

and Hepatology, Mayo Clinic, Rochester, MN<br />

Introduction: Acute cholangitis (AC) is well-recognized as a<br />

morbid and potentially lethal condition. Previous <strong>studies</strong> have<br />

investigated predictors of disease severity and short-term outcomes;<br />

however, data on long-term outcomes such as risk of<br />

recurrence and subsequent out-of-hospital mortality are lacking.<br />

We investigated potential differences in short and longterm<br />

outcomes of patients admitted with a sentinel episode<br />

of AC compared to those with a history of AC. Methods: We<br />

retrospectively identified all patients admitted to Mayo Clinic<br />

hospitals in Rochester, MN for AC from 2009-2011. Patients<br />

were categorized into two groups based on the presence or<br />

absence of prior AC. Data on pertinent clinico-demographic<br />

variables and outcomes including hospital length-of-stay (LOS),<br />

AC recurrence, and 1-year mortality were abstracted using a<br />

standardized data collection form. Tests of significance were<br />

2-tailed, and p


530A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

644<br />

Long-Term Safety of OCA in Patients with PBC<br />

Yvette Peters, Roya Hooshmand-Rad, Richard Pencek, Janet<br />

Owens-Grillo, Tonya Marmon, Leigh MacConell, David Shapiro;<br />

Intercept Pharmaceuticals, San Diego, CA<br />

Background: Obeticholic Acid (OCA) is a potent and selective<br />

FXR agonist developed for treatment of primary biliary cirrhosis<br />

(PBC). 216 patients with PBC were treated in a randomized,<br />

doubleblind (DB), placebo (PBO)controlled Phase 3 clinical<br />

study to evaluate the efficacy and safety of OCA. 198 patients<br />

completed the DB phase of the study and 193 enrolled in a<br />

longterm extension (LTSE) phase. Exposure to OCA during the<br />

DB phase resulted in statistically significant liver biochemistry<br />

improvements and was safe. Methods: All patients enrolled in<br />

the LTSE first met the inclusion criteria for the DB study, which<br />

included PBC diagnosis, ALP ≥1.67x ULN and/or total bilirubin<br />

>ULN to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 531A<br />

646<br />

Significant non–adherence to DAA HCV therapy in<br />

decompensated patients - a tertiary hepatology centre<br />

assessment<br />

Aisling B. Considine 1,2 , Suman Verma 3 , Kath Oakes 3 , Kate E.<br />

Childs 3 , Sarah Knighton 1,2 , Andrew Ayers 3 , Abid Suddle 3 , Kosh<br />

Agarwal 3 ; 1 Pharmacy, Kings College Hospital, London, United<br />

Kingdom; 2 Kings College London, Institute of Pharmaceutical Sciences,<br />

London, United Kingdom; 3 Kings College Hospital, Institute<br />

of Liver Studies, London, United Kingdom<br />

Background: The advent of directly acting anti-virals (DAA) has<br />

led to simplified HCV regimens with increased efficacy rates<br />

and better tolerability. The high adherence seen in clinical<br />

trials may not translate to real life populations. Optimal adherence<br />

is critical to protect patients from treatment failure and<br />

resistance. Aim: To identify potential factors which may contribute<br />

to sub-optimal adherence in a population with decompensated<br />

cirrhosis. Methods: HCV patients eligible for access<br />

to a 12 week treatment regimen under a UK NHSE mandated<br />

access scheme were consented to complete a baseline questionnaire<br />

capturing data on sociodemographics, clinical status<br />

and perceptions of illness.Adherence assessments at treatment<br />

weeks 4, 8 and 12 with pill counts and a Morisky Medication<br />

Adherence Scale were undertaken. Suboptimal adherence<br />

(SA) was defined as any report or pill count which indicated<br />

a delayed (>2hrs) or missed dose of anti-viral. Descriptive<br />

statistic and multivariate regression methods were utilised for<br />

analysis. Results: 80% (n=47) of the cohort had reached SVR<br />

12 and are included in the adherence analysis.85% received<br />

sofosbuvir/ledipasvir and 15% sofosbuvir with daclatasvir. All<br />

patients received ribavirin and had been attending specialist<br />

hepatology services for >12 months.43% (n=20) of patients<br />

demonstrated SA during their treatment. Multivariate analysis<br />

demonstrated that any non-attendance during treatment<br />

and a ‘limited support network’ predicted higher rates of SA<br />

(p3 medications<br />

(vs 74% in adherent group). The average MELD score was<br />

15 in the SA group (vs 13 in adherent group). 80% of SA<br />

was associated with symptoms secondary to disease stage not<br />

reflected by the MELD. SA secondary to side effects of therapy<br />

was not identified. 85% of the SA group achieved SVR12.Failure<br />

to achieve SVR was seen in those with SA adherence at all<br />

assessment points (weeks 4, 8 and 12). Conclusions: Our initial<br />

results demonstrate that despite an intensive adherence focused<br />

multi-disciplinary approach, SA occurs in decompensated HCV<br />

cirrhotic patients. Acceptable SVR rates were achieved in this<br />

population. Further research is warranted to develop strategies<br />

which maximise adherence in all HCV populations.<br />

Disclosures:<br />

Kosh Agarwal - Advisory Committees or Review Panels: Gilead, BMS, Novartis,<br />

Janssen, AbbVie, Gilead; Consulting: MSD, Janssen; Grant/Research Support:<br />

Roche, Gilead, BMS, BMS; Speaking and Teaching: Astellas<br />

The following authors have nothing to disclose: Aisling B. Considine, Suman<br />

Verma, Kath Oakes, Kate E. Childs, Sarah Knighton, Andrew Ayers, Abid Suddle<br />

647<br />

Hepatology nursing in Canada: Current practices and<br />

future challenges<br />

Colina Yim 1 , Cheryl Dale 2 , Geri Hirsch 3 , Jo-Ann E. Ford 4 , Carolyn<br />

Klassen 4 ; 1 University Health Network, Toronto, ON, Canada;<br />

2 London Health Sciences Centre, London, ON, Canada; 3 Halifax<br />

Capital District Health Authority, Halifax, NS, Canada; 4 Vancouver<br />

General Hospital, Vancouver, BC, Canada<br />

Background: Hepatitis C treatment landscape is changing.<br />

The rapid therapeutic advancements may impact the roles and<br />

responsibilities of some of the hepatology nurses whose practice<br />

focused solely on hepatitis C. This study aims to describe<br />

the current practices of hepatology nurses in Canada and<br />

to gain insights into nurses’ perceptions and concerns about<br />

their future roles and values. Methods: An 18-item survey was<br />

designed. Survey was sent in November 2014 to all active<br />

members of the Canadian Association of Hepatology Nurses<br />

(CAHN) via email for their voluntary participation. A second<br />

email was sent 8 weeks later as a reminder to complete survey.<br />

Demographics collected include age, length of practice,<br />

practice settings and focuses. CAHN Members were asked to<br />

rate their levels of concerns in regards to job security, sustainability<br />

of current positions, change of role and the future of<br />

hepatology nursing. Results: Of a total of 136 active CAHN<br />

members, 94 (69%) completed the survey. Eighty-one percent<br />

were employed full time, 78% age > 40 years, 64% have at<br />

least 6 years or more work experience in hepatology. Hepatitis<br />

C treatment was the sole practice focus in 52% while 38% of<br />

them managed more than 50 hepatitis C patients on treatment<br />

in last 6 months. A majority of CAHN members (78%) work<br />

in academic institutions and community settings. Salary support<br />

varies with 35% being paid from industry grants and the<br />

remainder either by government (health authority, provinces,<br />

academic centres and community agencies) or by research<br />

grants. Up to 85% expressed satisfaction with their current<br />

positions and the support they received from their physician<br />

partners but at least 60% are concerned about their job security,<br />

job description change and current position sustainability.<br />

A lack of funding for hepatology nurse positions remain the<br />

major challenge, 69% of members predict a positive hepatology<br />

nursing future. Conclusions: Hepatology nurses in Canada<br />

are experienced professionals in the field. A majority have a<br />

focused practice on hepatitis C treatment management. We<br />

are satisfied with our current roles and responsibilities and<br />

feel most supported by our physician colleagues. Hepatology<br />

nurses acknowledge challenges exist for role sustainability but<br />

remain optimistic about our values in future.<br />

Disclosures:<br />

Colina Yim - Advisory Committees or Review Panels: Janssen; Consulting: Gilead;<br />

Speaking and Teaching: Abbvie<br />

Jo-Ann E. Ford - Advisory Committees or Review Panels: Gilead, Roche, Janssen,<br />

Merck Canada, Vertex; Speaking and Teaching: Gilead, Roche, Janssen, Merck<br />

Canada, Vertex<br />

The following authors have nothing to disclose: Cheryl Dale, Geri Hirsch, Carolyn<br />

Klassen<br />

648<br />

Pharmacists mitigate effect of simeprevir on blood pressure<br />

in patients with hepatitis C infection receiving concurrent<br />

calcium channel blocker therapy<br />

Heather Johnson 1,2 , Emily Graham 2 , Michael A. Dunn 1 , Kapil B.<br />

Chopra 1 ; 1 University of Pittsburgh, PIttsburgh, PA; 2 UPMC Presbyterian,<br />

Pittsburgh, PA<br />

Simeprevir (SMV) an NS3/4A inhibitor for the treatment of<br />

Hepatitis C infection (HCV), has a number of drug-drug inter-


532A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

actions mediated by CYP3A4 or p-glycoprotein. Current product<br />

labeling indicates increased clinical monitoring of patients<br />

receiving SMV with oral calcium channel blockers (CCB), but<br />

no specific guidance is available. Pharmacists complete a<br />

medication review for all patients receiving therapy for HCV<br />

to identify potential drug-drug interactions and opportunities<br />

to mitigate interactions. We examined blood pressure control<br />

in patients on SMV and simultaneous CCB, regardless of<br />

mitigating interventions, between December 2014 and June<br />

2015 to determine if concurrent treatment with SMV and a<br />

CCB resulted in hypotension leading to adjustments in antihypertensive<br />

regimens or if hypertension resulted in patients who<br />

had preemptive CCB dose reductions before SMV initiation.<br />

Twenty-three patients were included. Of those, 65.2% were<br />

male and 95.6% were Caucasian. Patients were receiving an<br />

average of 5.2mg amlodipine equivalents at initiation of SMV.<br />

Doses of CCB were preemptively reduced in 9 (39%) patients<br />

prior to SMV initiation. Prior to preemptive dose reduction, the<br />

mean amlodipine equivalent in patients was 8.2 mg . All 9<br />

patients who received preemptive dose reduction of their CCB<br />

were concurrently taking other antihypertensive medications.<br />

Most were receiving one additional antihypertensive (7) with 1<br />

each receiving 2 and 3 other medications. In patients without<br />

CCB dose reduction, 6 were on one and 2 were on 2 other<br />

antihypertensive medications. Only one patient without preemptive<br />

CCB dose reduction discontinued CCB during SMV<br />

treatment. Two patients discontinued other antihypertensive<br />

therapies due to hypotension during concurrent treatment with<br />

SMV and a CCB. The average mean arterial pressure (MAP) in<br />

all patients prior to initiating SMV was 95 mmHg. The average<br />

MAP in all patients mid-treatment with SMV and after discontinuation<br />

was 93 and 90 mmHg, respectively. In patients who<br />

had preemptive dose reductions of their CCB, the MAP at baseline,<br />

mid-treatment, and after discontinuation were 94.3, 93.2,<br />

and 90.6 mmHg, respectively. Preemptive CCB dose reduction<br />

is a safe strategy in patients receiving SMV and did not lead to<br />

uncontrolled hypertension. Most patients receiving a CCB with<br />

SMV did not experience hypotension during therapy, however,<br />

further dose reduction may be warranted in select patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Heather Johnson, Emily Graham,<br />

Michael A. Dunn, Kapil B. Chopra<br />

649<br />

FGF21 facilitates normal liver regeneration and restores<br />

impaired liver regeneration in steatotic liver<br />

Hui-Xin Liu, Ying Hu, Yu-Jui Yvonne Wan; Medical Pathology and<br />

Laboratory Medicine, UC DAVIS, Sacramento, CA<br />

Fibroblast growth factor 21 (FGF21) is a master metabolic<br />

regulator that has a remarkable ability to reverse obesity,<br />

repair injured tissue, and prolong lifespan. The current study<br />

examined the effect of FGF21 in regulating 2/3 partial hepatectomy<br />

(PHx)-induced liver regeneration in normal diet and<br />

Western diet-fed mice. In vitro, adenovirus expressing FGF21<br />

(Ad-FGF21) was used to overexpress FGF21 in mouse primary<br />

hepatocytes. In the absence of serum, exogenous FGF21<br />

increased hepatocyte viability by 160% indicating a mitogenic<br />

effect. In mice, PHx resulted in transient induction of FGF21<br />

and activation of ERK1/2, which was associated with induction<br />

of cell cycle genes 1 Day after liver resection. Forced<br />

expression of FGF21 through tail vein injection with Ad-FGF21<br />

accelerated hepatocyte proliferation that was accompanied<br />

by earlier activation of ERK1/2 and induction of cell cycle<br />

genes compared to mice injected with control Ad-CMV. Mild<br />

steatosis, which is considered as an energy source for hepatocyte<br />

proliferation, was absent in Ad-FGF21-injected mice<br />

1-1.5 Days after PHx. Conversely, FGF21 knockdown using<br />

siRNA (siFGF21) increased the severity of steatosis, delayed<br />

the induction of cell cycle genes, and decreased the number<br />

of Ki67-positive cells. These findings suggested a close<br />

relationship between FGF21-regulated lipid metabolism and<br />

hepatocyte proliferation. Consistently, liver regeneration was<br />

delayed in steatotic livers induced by Western diet feeding.<br />

The impaired liver regeneration, accompanied by decreased<br />

Ki67-positive cells, reduced cell cycle gene expression, and<br />

steatosis, was further exacerbated by FGF21 knockdown.<br />

However, forced expression of FGF21 in Western diet-induced<br />

fatty livers rescued impaired liver regeneration as indicated<br />

by normalized cell cycle gene and lipid metabolizing gene<br />

expression patterns. Conclusion: FGF21 plays a pivotal role in<br />

regulating metabolism and liver regeneration. Forced expression<br />

of FGF21 facilitates normal regeneration and restores the<br />

regeneration program in Western diet-induced fatty liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hui-Xin Liu, Ying Hu, Yu-Jui<br />

Yvonne Wan<br />

650<br />

HNF4α reprogramming during initiation and termination<br />

of liver regeneration<br />

Ian Huck, Michael W. Manley, Jr., Udayan Apte; Pharmacology,<br />

Toxicology and Therapeutics, University of Kansas Medical Center,<br />

Kansas City, KS<br />

Hepatocyte nuclear factor 4 alpha (HNF4α) is an orphan<br />

nuclear receptor essential for hepatocyte differentiation and<br />

function. Recent <strong>studies</strong> have indicated that HNF4α also<br />

inhibits hepatocyte proliferation and may be tumor suppressor.<br />

However, the role of HNF4α in liver regeneration is not<br />

known. We investigated changes in HNF4α expression and<br />

function in liver regeneration after partial hepatectomy (PH)<br />

using three-month-old male C57BL/6 mice. Real Time PCRbased<br />

analysis revealed presence of mRNA for both P1 and P2<br />

isoforms after PH but only the P1 (adult) form of HNF4α protein<br />

was detected in the liver by Western blot. A rapid decline in<br />

total and nuclear levels of HNF4α was observed within 1 hr<br />

after PH, which further decreased at 6 hr post-PH. Whereas,<br />

HNF4α protein increased at 12 hr post PH as compared to 6<br />

hr, it remained significantly lower than 0 hr and continued to<br />

be lower till 5 days after PH, at which time it returned to control<br />

(0 hr) levels. A coinciding decline in HNF4α function as measured<br />

by 44 select targets genes was observed within 12 hr<br />

after PH, which recovered substantially by 5 days post PH. We<br />

further studied the role of HNF4α is liver regeneration after PH<br />

using conditional hepatocyte specific HNF4α knockout mice<br />

(HNF4α-KO). Cell proliferation measured by PCNA analysis<br />

started 24 hr earlier in the HNF4a-KO mice as compared to<br />

WT and continued beyond 7 days after PH. HNF4α deletion<br />

resulted in an accelerated liver regeneration and induced a<br />

defect in termination of regeneration after PH. In conclusion,<br />

these data indicate that HNF4α undergoes significant reprogramming<br />

during liver regeneration after PH and is involved in<br />

both initiation and termination of liver regeneration.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ian Huck, Michael W. Manley,<br />

Jr., Udayan Apte


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 533A<br />

651<br />

Skeletal muscle mitochondrial fragmentation results in<br />

functional abnormalities during hyperammonemia of<br />

cirrhosis<br />

Gangarao Davuluri 2 , Avinash Kumar 2 , Samjhana Thapaliya 2 ,<br />

Allawy Allawy 2,5 , Dharmvir Singh 2 , Julie H. Rennison 2 , Rafaella<br />

Nascimento e Silva 2 , David R. Van Wagoner 3 , Sathyamangla V.<br />

Naga Prasad 3 , Xin Qi 4 , Hoppel Charles 4 , Takhar Kasumov 2 , Srinivasan<br />

Dasarathy 1 ; 1 Department Of Gastroenterology and Hepatology,<br />

Cleveland Clinic, Cleveland, OH; 2 Pathobiology, Cleveland<br />

Clinic, Cleveland, OH; 3 Molecular Cardiology, Cleveland Clinic,<br />

Cleveland, OH; 4 Pharmacology and Medicine, Case Western<br />

Reserve University, Cleveland, OH; 5 Department Of Internal Medicine,<br />

Cleveland Clinic, Cleveland, OH<br />

Background. Hyperammonemia is a consistent abnormality<br />

in cirrhosis and contributes to sarcopenia. Skeletal muscle is<br />

a major alternate organ for ammonia disposal by the initial<br />

reaction of conversion to glutamate in the mitochondria. We<br />

have observed skeletal muscle mitochondrial functional abnormalities<br />

in cirrhosis and hyperammonemia. It is not known if<br />

the functional abnormalities are due to structural alterations<br />

in the skeletal muscle mitochondria. Methods. Mitochondrial<br />

mass was measured by immunoblots for mitochondrial proteins<br />

citrate synthase, voltage dependent anion channels and cyclooxygenase<br />

IV during hyperammonemia in C2C12 myotubes.<br />

Additional methods to quantify mitochondrial mass included<br />

electron microscopy and immunofluorescence images using<br />

mitochondrial dye, MitoTracker Red. Mitochondrial fragmentation<br />

was quantified using electron microscopy and immunofluorescence<br />

images with Mitotracker red. Membrane polarization<br />

generated by proton pumping by components of the electron<br />

transport chain (ETC) was quantified using the dye TMRE and<br />

uncoupler, FCCP was used as a positive control. The mechanism<br />

of altered mitochondrial dynamics during hyperammonemia<br />

was determined by quantifying genes regulating mitochondrial<br />

fission and fusion by real time PCR and immunoblots of DRP1<br />

protein expression and polymerization polymerization. Finally,<br />

siRNA to knockdown of DRP1 and P110 peptide to block<br />

DRP1 function in myotubes were used to show that ammonia<br />

promoted DRP1 polymerization that resulted in mitochondrial<br />

fragmentation and dysfunction. Results. Despite unaltered mitochondrial<br />

mass with hyperammonemia, increased mitochondrial<br />

fragmentation was observed on both immunofluorescence<br />

and electron microscopy. These changes were accompanied<br />

by consistent alterations in expression of genes regulating mitochondrial<br />

dynamics, Fis 1, OPA1, MFN1, and MFN2. We<br />

also observed increased polymerization of dynamin related<br />

protein-1 (DRP1) during hyperammonemia in the myotubes.<br />

Genetic and pharmacologic inhibition of DRP1 in restored<br />

ATP content and decreased reactive oxygen species in myotubes<br />

regardless of hyperammonemia. These changes were<br />

accompanied by loss of membrane potential during hyperammonemia.<br />

Conclusions. Increased mitochondrial fragmentation<br />

was observed with unaltered mitochondrial mass and was due<br />

to polymerization of dynamin related protein-1 (DRP1). Since<br />

blocking mitochondrial fragmentation reversed the low ATP<br />

and elevated ROS during hyperammonemia, our observations<br />

provide a novel approach to reverse skeletal muscle bioenergetic<br />

dysfunction and sarcopenia of cirrhosis by targeting<br />

mitochondrial fragmentation using anti DRP1 therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gangarao Davuluri, Avinash<br />

Kumar, Samjhana Thapaliya, Allawy Allawy, Dharmvir Singh, Julie H. Rennison,<br />

Rafaella Nascimento e Silva, David R. Van Wagoner, Sathyamangla V. Naga<br />

Prasad, Xin Qi, Hoppel Charles, Takhar Kasumov, Srinivasan Dasarathy<br />

652<br />

Thr505 RelA phosphorylation controls liver proliferative<br />

response and supresses NF-κB tumor-promoting activities<br />

Anna Moles 1 , Jacqueline Butterworth 2 , Jill E. Hunter 2 , Ana M.<br />

Sanchez 2 , Dina Tiniakos 1 , Derek Mann 1 , Fiona Oakley 1 , Neil<br />

D. Perkins 2 ; 1 Institute of Cellular Medicine, Newcastle University,<br />

Newcastle Upon Tyne, United Kingdom; 2 Institute for Cell and<br />

Molecular Biosciences, Newcastle University, Newcastle Upon<br />

Tyne, United Kingdom<br />

Background and aims: NF-κB is a family of transcription factors<br />

that are key regulators of the immune and inflammatory<br />

responses. NF-κB regulation is complex and occurs at many<br />

levels, including direct post-translational modifications (PTMs)<br />

of the subunits. Phosphorylation of RelA (p65) at Thr505 regulates<br />

proliferation, migration and autophagy in in vitro cellular<br />

systems; however its role and relevance in vivo still remains<br />

unclear. Thus, the aim of this study was to analyse the role<br />

of RelA Thr505 phosphorylation in vivo by mutating this site<br />

through creating a knock in mouse. Methods: Knock in mutant<br />

mice where generated where Thr505 was substituted by Alanine,<br />

thus preventing phosphorylation. Partial hepatectomy<br />

(PhX), acute CCl 4<br />

administration and 15 days DEN injection<br />

was performed in WT and knock in mice. Animals were culled<br />

at 36, 72 hours and 5 days for PhX; at 24, 48 and 72 hours<br />

for acute CCl 4<br />

and 30 weeks post-DEN. Paraffin staining for<br />

BrdU and PCNA were used as markers of proliferation and<br />

γH2AX for DNA damage. Western blot was used to determined<br />

p53 and p-JNK expression. Gene set enrichment analysis<br />

(GSEA) was performed in WT and RelA T505 livers after 36<br />

hours of PhX. Tumor assessment was performed by an expert<br />

pathologist. Results: Thr505 mutation of RelA resulted in an<br />

increased liver to body weight ratio after PhX at 36, 72 hours<br />

and 5 days. Hepatocyte proliferation was increased after both<br />

injuries, PhX and acute CCl 4,<br />

at all the time points in RelA<br />

T505 mice versus WT, as assessed by PCNA and/or BrdU<br />

staining. Furthermore, mitotic body counts were also higher<br />

in RelA T505 mice after both injuries. Of note, proliferation<br />

after PhX was sustained for longer in RelA T505 mice, pointing<br />

towards an inefficient resolution of the proliferative response.<br />

Increased DNA damage, as determined by γH2AX staining,<br />

was seen in both WT and RelA T505 after 36 hours of PhX.<br />

However, the DNA damage was significantly higher in RelA<br />

T505 mice in comparison with WT. Western blot analysis of<br />

these samples revealed increased expression of the p53 tumour<br />

suppressor and active JNK in RelA T505 mice. Finally, chronic<br />

DEN administration resulted in a significant increased tumor<br />

number, mostly HCC, in RelA T505 versus WT. Conclusions<br />

RelA Thr505 phosphorlyation provides an important and novel<br />

regulatory mechanism to control the liver proliferative response<br />

and is a critical pathway suppressing NF-κB tumour-promoting<br />

activities of RelA.<br />

Disclosures:<br />

The following authors have nothing to disclose: Anna Moles, Jacqueline Butterworth,<br />

Jill E. Hunter, Ana M. Sanchez, Dina Tiniakos, Derek Mann, Fiona<br />

Oakley, Neil D. Perkins


534A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

653<br />

Lysophosphatidic acid receptor 3 (LPAR3) is expressed<br />

by cancer progenitor cells in human HCC tissue at the<br />

tumor:non-tumor margin and regulates cell migration.<br />

Valentina Zuckerman 1 , Eugene Sokolov 1 , Jacob H. Swet 1 , Victor<br />

C. Showalter 1 , William A. Ahrens 2 , David A. Iannitti 1 , Iain H.<br />

McKillop 1 ; 1 Department of Surgery, Carolinas Medical Center,<br />

Charlotte, NC; 2 Department of Pathology, Carolinas Medical Center,<br />

Charlotte, NC<br />

Background The poor prognosis for hepatocellular carcinoma<br />

(HCC) patients can be attributed, in part, to late detection,<br />

rapid tumor expansion, and metastases. LPAR3 is detected<br />

at low levels in the liver/ hepatocytes, and previous <strong>studies</strong><br />

report LPAR3 expression increases in HCC and diseased/cirrhotic<br />

non-tumor liver (NTL) from HCC patients compared to<br />

healthy normal liver (NL). A growing interest in stem cells in<br />

tumor biology, and the role of LPAR3 in other cancers, led us<br />

to hypothesize that changes in LPAR3 expression in HCC may<br />

occur in a subset of hepatic progenitor cells involved in tumor<br />

progression/expansion. Methods Sections from human HCC-<br />

NTL, and lysates from SKHep1 cells (liver cancer progenitor<br />

cells) and NL were analyzed for LPAR3, CD44 (mesenchymal<br />

stem cell marker), EpCAM (cancer progenitor cell marker),<br />

and Hepar1 (hepatocyte marker) expression-localization. The<br />

effect of LPA on cell function was determined in SKHep1 using<br />

2D and 3D cell migration analysis. Cell signaling (AKT/MEK<br />

activity) and migration following LPA (10μM) treatment was<br />

analyzed in conjunction with inhibition of Gi-protein dependent<br />

signaling (PI3K inhibitor (LY294002 [LY], 40μM) and MEK<br />

inhibitor (U0126, 5μM)). In parallel shRNA was used to inhibit<br />

endogenous LPAR3 expression. Results Extensive co-localization<br />

of LPAR3 with CD44 and EpCAM was detected in the<br />

HCC-NTL margin, but not the NTL or tumor mass, data mirrored<br />

in SKHep1 cells (LPAR3, CD44 and EpCAM detected). Conversely,<br />

Hepar1 was readily detected in NL and the HCC mass,<br />

but not SKHep1 cells, or cells in the HCC margin that stained<br />

positive for LPAR3. In vitro analysis demonstrated LPA-dependent<br />

PI3K and ERK activation in SKHep1 cells resulting in<br />

significant cell migration. Pretreatment with LY and U0126<br />

abrogated LPA-dependent PI3K and MEK activation. However,<br />

inhibition of PI3K (LY) failed to significantly alter LPA-dependent<br />

migration whereas ERK inhibition (U0126) abolished LPA-dependent<br />

migration. The effects of pharmacological inhibition<br />

were mirrored by inhibiting LPAR3 expression (shRNA), in<br />

which exogenous LPA failed to stimulate ERK activation and<br />

cell migration. Summary These data identify a unique subset<br />

of LPAR3 positive cancer progenitor cells at the tumor–NTL<br />

margin in HCC patients. Using an in vitro model of hepatic<br />

tumor progenitor (SKHep1) cells, with similar LPAR profiles to<br />

in vivo tissue, we report LPA regulates tumor cell migration via<br />

Gi-MEK-ERK dependent signaling. These data suggest targeting<br />

LPA-LPAR3 signaling within this progenitor cell subset in the<br />

liver may slow tumor expansion and/or metastasis and limit<br />

disease progression.<br />

Disclosures:<br />

David A. Iannitti - Consulting: Davol, Covidien, Ethicon Endosurgery, Davol,<br />

Covidien, Ethicon Endosurgery, Davol, Covidien, Ethicon Endosurgery, Davol,<br />

Covidien, Ethicon Endosurgery; Speaking and Teaching: Davol, Covidien,<br />

Ethicon Endosurgery, Davol, Covidien, Ethicon Endosurgery, Davol, Covidien,<br />

Ethicon Endosurgery, Davol, Covidien, Ethicon Endosurgery<br />

The following authors have nothing to disclose: Valentina Zuckerman, Eugene<br />

Sokolov, Jacob H. Swet, Victor C. Showalter, William A. Ahrens, Iain H. McKillop<br />

654<br />

Hepatitis B virus X protein promotes HCC metastasis<br />

through activation of CD44v6 splicing form switching by<br />

regulation of splicing factor ESRP1/ESRP2<br />

Ya-Ju Chang 1 , Li-Chen Wu 2 , Ja-an Annie Ho 1 ; 1 Department of<br />

Biochemical Science and Technology, National Taiwan University,<br />

Taipei, Taiwan; 2 Department of Applied Chemistry, National Chi<br />

Nan University, Puli, Taiwan<br />

Purpose Hepatitis B virus x (HBx) has been implicated in the<br />

development of hepatocellular carcinoma (HCC) through the<br />

promotion of malignant transformation, upregulation of the<br />

properties of liver cancer stem cells (CSC), and manipulation<br />

of cell cycle progression, but it is not clear how HBx promotes<br />

HCC metastasis. Among CSC markers, CD44 is associated<br />

with migration, proliferation, and invasion, especially CD44v6,<br />

which is critical for metastasis. Herein the effects of HBx on<br />

CD44v6-dependant HCC metastasis was investigated. Methods<br />

Stable tet-on inducible HBx overexpression cells were obtained<br />

by transfection of pLV-CMVTO-GFP-HBx plasmids to HepG2<br />

for studying HCC metastasis. The expression of CD44v6 was<br />

induced by doxycycline (Doxy.) and was determined by immunoblotting,<br />

quantitative real-time PCR (Q-PCR), and flow cytometry.<br />

The metastatic ability of HBx-induced HCC cell lines was<br />

investigated by real-time cell analysis. The effect of HBx/snail<br />

complex on the expression of CD44 splicing factor ESRP1 and<br />

ESRP2 was analyzed. The inhibitory effects of siESRP1 and<br />

siESRP2 were also studied. Xenograft tumor model was used for<br />

in vivo study for the investifation of the association of CD44v6<br />

in metastasis and the splicing factor on HBx expressed tumors.<br />

Results HBx significantly induced the expression of CD44v6<br />

as identified by immunoblotting, Q-PCR, and flow cytometry.<br />

Through interaction with snail, HBx promoted the expression<br />

of downstream signaling CD44-related splicing factor ESRP2,<br />

and restricted ESRP1. Inhibition of snail and ESRP2 but not<br />

ESRP1 markedly reduced the mobility of HBx-expressed cells.<br />

Addtionally, overexpressed N-cadherin induced by HBx but not<br />

E-cadherin was detected, indicating the undergoing of EMT.<br />

MMP2 and 9 were also evaluated for the potential of metastasis.<br />

HBx enhanced the expression of these two metalloproteinases.<br />

Futher metastasis was confirmed by enhanced RhoA-GTP,<br />

a ROCK family protein participating in cell migration/invasion,<br />

determined by IP/IB in HBx expressed HepG2, and cell metastatic<br />

ability as determined by real-time cell monitoring. Based<br />

on these findings, HBx appears to significantly promote HCC<br />

invasion through snail modulated ESRP1/ESRP2, which might<br />

be a driven force to enhance CD44v6-dependent metastatic<br />

ability of HCC. Conclusion HBx was able to promote metastatic<br />

ability in HCC cell line through epigenetic regulation. The metastatic<br />

modification of CD44 variant was through the ESPR1<br />

and ESRP2, that function as specific modulators in CD44 splicing<br />

to maneuver metastatic ability. Inhibition of CD44 splicing<br />

seems to be a potential approach to arrest the progression of<br />

HBV-related HCC metastasis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ya-Ju Chang, Li-Chen Wu, Ja-an<br />

Annie Ho


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 535A<br />

655<br />

Regulation of the Hepatic Drug-metabolizing Enzymes<br />

by Probiotics and Conventionalization in Germ-free<br />

Mice<br />

Felcy pavithra Selwyn samraj, Sunny Lihua Cheng, Curtis Klaassen,<br />

Julia Yue Cui; Environmental and Occupational Health Sciences,<br />

University of Washington, Seattle, WA<br />

The use of probiotics and antibiotics has raised concerns of<br />

potential adverse “drug-bacteria” interactions in humans. VSL3<br />

is a probiotic medication that delivers 8 live strains of bacteria<br />

to the host, and has been used to treat ulcerative colitis. The<br />

present study utilized conventional (CV) and germ-free (GF)<br />

mice to determine the effects of VSL3 and conventionalization<br />

of GF mice by exposing them to the CV housing environment<br />

on the expression of approximately 90 critical Phase-I<br />

and –II drug-metabolizing enzymes (DMEs) in liver. Starting at<br />

2-months of age, CV and GF mice (male, n=6-8/group) were<br />

treated with or without VSL3 (4.5×10 6 CFU/ml) in drinking<br />

water for 28 days. In a separate study, 1-month old GF mice<br />

(male, n=4/group) were taken out of the GF isolator and were<br />

housed in the conventional environment for 2 months. Regarding<br />

the bacteria profiles, VSL3 increased each VSL3 component<br />

in the large intestinal content of CV mice, and increased<br />

these bacteria even more in GF mice, likely due to less competition<br />

for growth in the GF-environment (16S rRNA qPCR).<br />

Conventionalization of GF mice increased the total bacteria<br />

and selected bacteria strains with known important functions<br />

for host metabolism in the large intestinal content. Regarding<br />

the effect of VSL3 on the expression of DMEs in livers of CV<br />

mice, VSL3 increased the mRNAs of the Phase-I enzymes cytochrome<br />

P450 (Cyp) 4v3, Alcohol dehydrogenase (Adh) 1,<br />

and Carboxyesterase (Ces) 2a, but decreased the mRNAs<br />

of Cyp3a44 and 3a11; VSL3 also decreased the mRNAs<br />

of multiple Phase-II glutathione-S-transferases (Gstm1-3 and<br />

Gsto1). In livers of GF mice, VSL3 decreased the mRNAs of<br />

the UDP-glucuronosyl transferases (Ugt) 1a9 and 2a3. Regarding<br />

the effects of GF and conventionalization on the hepatic<br />

expression of DMEs, GF mouse livers had increased mRNA of<br />

Ugt1a9, but decreased mRNAs of Gstpi and sulfotransferase<br />

5a1. GF conditions down-regulated many genes in the Cyp3a<br />

cluster but up-regulated many genes in the Cyp4a cluster. Conventionalization<br />

normalized the expression of these genes back<br />

to CV levels. The protein expression of Cyp3a and Cyp4a was<br />

confirmed by Western blot. ChIP-qPCR indicated that changes<br />

in the hepatic PXR- and PPARα-DNA binding to the Cyp3a<br />

and Cyp4a gene clusters during GF- and exGF-conditions<br />

appeared to be responsible for the changes in the Cyp3a and<br />

4a gene expression. In conclusion, alterations in the intestinal<br />

bacteria by VSL3, GF condition, and conventionalization<br />

impact the expression of many DMEs in the host liver, suggesting<br />

the importance of considering “bacteria-drug” interactions<br />

in human populations. (Supported by NIH GM111381,<br />

ES019487, and P30 ES0007033)<br />

Disclosures:<br />

The following authors have nothing to disclose: Felcy pavithra Selwyn samraj,<br />

Sunny Lihua Cheng, Curtis Klaassen, Julia Yue Cui<br />

656<br />

HMGB1 interacts with AIM2 to upregulate hepatocyte<br />

autophagy and prevent cell death after redox stress<br />

Qian Sun, Timothy R. Billiar, Melanie Scott; Surgery, university of<br />

pittsburgh, Pittsburgh, PA<br />

The release of damage-associated molecular patterns such as<br />

double-stranded DNA (dsDNA) after oxidative stress leads<br />

to maturation of the inflammasome and caspase-1 in both<br />

immune and non-immune cells. We have previously shown<br />

that activation of AIM2 inflammasome and caspase-1 in<br />

mouse hepatocytes (HCs) upregulates mitochondrial autophagy<br />

after redox stress rather than leading to the maturation of<br />

cytokines as in immune cells. Nuclear protein HMGB1 binds<br />

dsDNA and translocates to the cytosol during redox stress.<br />

Cytosolic HMGB1 has been shown to upregulate autophagy<br />

to protect against mitochondrial dysfunction. Therefore, we<br />

hypothesized that cytosolic HMGB1 interacts with AIM2, and<br />

its activating ligand dsDNA, to regulate caspase-1 activation<br />

and caspase-1-mediated mitochondrial autophagy in HCs.<br />

Methods: HCs isolated from C57BL/6 (WT) and HC-HMGB1 -/-<br />

(hepatocyte-specific HMGB1 knockout) mice were transfected<br />

with GFP-LC3 before hypoxia (6h, 1% O 2<br />

) and reoxygenation<br />

treatment. Levels of autophagic flux were assessed by increase<br />

in the number of GFP–LC3 puncta in HCs after treatment of<br />

bafilomycin (50nM, 1h). Mitochondrial and cytosolic specific<br />

ROS production was measured by HyPer-Mito and HyPer-<br />

Cyto respectively. Cell death was determined by Annexin V/<br />

PI staining. Interaction between HMGB1 and AIM2 in HCs<br />

was assessed by immunoprecipitation. Results: Hypoxia/reoxygenation<br />

triggered enhanced association between AIM2 and<br />

HMGB1, as well as increased caspase-1 activity in WT HCs,<br />

but not in HC-HMGB1 -/- cells, suggesting a role for HMGB1<br />

in regulating caspase-1 activation. Autophagy levels were<br />

increased in WT HCs after hypoxia/reoxygenation as shown<br />

by increased beclin1 expression and autophagic flux. However,<br />

HC-HMGB1 -/- HCs showed decreased beclin1 levels compared<br />

with WT and impaired autophagic flux after hypoxia/<br />

reoxygenation, similar to our previous results in AIM2 -/- HCs.<br />

Additionally, HC-HMGB1 -/- HCs had increased mitochondrial<br />

volume and ROS production in the mitochondria in comparison<br />

with WT after hypoxia/reoxygenation, suggesting a role for<br />

HMGB1 in up-regulating mitochondrial autophagy. Consistent<br />

with our previous findings in caspase-1 -/- and AIM2 -/- HCs,<br />

HC-HMGB1 -/- HCs showed increased apoptosis and necrosis<br />

in comparison with WT after hypoxia/reoxygenation, confirming<br />

a protective role for HMGB1 in HCs. Conclusion: Our data<br />

suggest that HMGB1 is protective in HCs after redox stress by<br />

upregulating mitochondrial autophagy through its interaction<br />

with AIM2. Our study links HMGB1 with AIM2 inflammasome<br />

activation and stress-induced autophagy in non-immune cells,<br />

which may have implications for future inflammasome-targeting<br />

therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Qian Sun, Timothy R. Billiar,<br />

Melanie Scott<br />

657<br />

MYCN is A Novel Biomarker for Chemoprevention of<br />

Hepatocellular Carcinoma by Acyclic Retinoid<br />

Soichi Kojima 1 , Xian-Yang Qin 1 , Harukazu Suzuki 2 , Masao<br />

Honda 3 , Goshi Shiota 4 , Naoto Ishibashi 5 , Hisataka Moriwaki 6 ,<br />

Masahito Shimizu 7 ; 1 Micro-Signaling Regulation Technology<br />

Unit, RIKEN, Wako, Japan; 2 Division of Genomic Technologies,<br />

RIKEN CLST, Yokohama, Japan; 3 Department of Gastroenterology,<br />

Kanazawa University, Kanazawa, Japan; 4 Division of Molecular<br />

and Genetic Medicine, Tottori University, Yonago, Japan; 5 Pharmaceutical<br />

Division, KOWA Company, Higashimurayama, Japan;<br />

6 Gifu University, Gifu, Japan; 7 Department of Gastroenterology,<br />

Gifu University School of Medicine, Gifu, Japan<br />

Background & aim: Poor prognosis of hepatocellular carcinoma<br />

(HCC) is partly due to its high rate of recurrence. Acyclic<br />

retinoid (ACR) is under phase III clinical trials in Japan to


536A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

prevent the recurrence and development of HCC after surgical<br />

removal of the primary tumors. Previous in vitro experiments<br />

showed that ACR selectively suppresses the growth of human<br />

HCC cells (JHH7, HuH7, and HepG2) but not normal human<br />

hepatic cells (Hc). In the present study, the genome-wide transcriptome<br />

analysis was performed to identify a biomarker for<br />

ACR treatment of HCC. Method: Differential gene expression<br />

profiles of JHH7 and Hc cells treated with ACR were measured<br />

using the next-generation sequencing-based Cap Analysis<br />

Gene Expression (CAGE) analysis. Inhibitory effect of ACR on<br />

MYCN expression was confirmed by real-time PCR in culture<br />

cells, animal models and liver biopsy of HCC patients administered<br />

with ACR (n = 6). MYCN-dependent signaling pathways<br />

underlying the growth suppression by ACR were explored<br />

using knowledge-based ingenuity pathway analysis (IPA),<br />

RNA interference, chemical inhibitors and luciferase assays.<br />

Relationship between MYCN levels and tumor recurrence in<br />

HCC patients (n = 102) was evaluated using the log-rank test.<br />

Correlations between MYCN and liver cancer stem cell (CSC)<br />

markers were determined using Pearson correlation coefficients<br />

based on microarray database or in patients. Results: MYCN<br />

was expressed in JHH7 cells but not in Hc cells, and inhibited<br />

by ACR but not by its ethyl analogues that showed no growth<br />

suppression. ACR also selectively suppressed the growth of<br />

MYCN-positive neuroblastoma cells, but not that negative for<br />

MYCN. Four out of 6 liver biopsies of HCC patients (66.7%)<br />

who had received 8 weeks of ACR (600 mg/day) after definitive<br />

treatment showed decreased MYCN expression (< 0.5-<br />

fold). The result of IPA suggested a Sp1/MYCN/caspase-8<br />

signaling pathway underlying ACR’s anticancer effect. Upon<br />

knockdown of MYCN, HCC cells showed suppressed cell<br />

growth and increased caspase-8 activity. An Sp1 inhibitor<br />

suppressed MYCN expression, whereas a caspase-8 inhibitor<br />

restored ACR-induced growth suppression. Finally, clinical<br />

analysis indicated that MYCN expression in HCC tumors<br />

was significantly correlated with CSC markers AFP, EpCAM<br />

and CD133 but not CD90, and negatively associated with<br />

recurrence of early-stage HCC. Conclusion: MYCN is a new<br />

biomarker of eligibility for ACR treatment that correlates with<br />

liver CSC with progenitor features and poor prognosis of HCC.<br />

ACR suppressed HCC cell growth through a Sp1/MYCN/<br />

caspase-8 dependent signaling pathway.<br />

Disclosures:<br />

Naoto Ishibashi - Employment: KOWA Company, LTD.<br />

The following authors have nothing to disclose: Soichi Kojima, Xian-Yang Qin,<br />

Harukazu Suzuki, Masao Honda, Goshi Shiota, Hisataka Moriwaki, Masahito<br />

Shimizu<br />

658<br />

Farnesoid X Receptor (FXR) Activation by Bile Salts and<br />

FXR Agonists Increases the Hepatic Expression of the<br />

Unfolded Protein Response Regulator X-box Binding<br />

Protein 1 Spliced (XBP1s)<br />

Xiaoying Liu, David Hilburn, Seong W. Park, Brian E. LeCuyer,<br />

Richard M. Green; Northwestern University, Chicago, IL<br />

The Unfolded Protein Response (UPR) is an adaptive response<br />

to endoplasmic reticulum stress that occurs in cholestatic and<br />

other forms of liver disease. X-box binding protein 1 spliced<br />

(XBP1s) is a highly conserved regulator of the UPR, and is<br />

formed by an unconventional splicing mechanism that removes<br />

26 nucleotides from the unspliced XBP1 mRNA. The farnesoid<br />

X receptor (FXR) is a nuclear bile acid receptor which regulates<br />

bile acid metabolism, and FXR agonists may provide<br />

novel therapies for cholestatic and fatty liver diseases. In this<br />

study, we demonstrate a novel role of FXR in regulating liver<br />

XBP1 splicing and XBP1s expression. Methods: Male FVB mice<br />

(8-12 weeks) were fed chow supplemented with 0.3% sodium<br />

deoxycholate (DCA), an endogenous FXR bile salt ligand, for<br />

up to 7 days. Huh7/NTCP or HepG2 cells were treated with<br />

either bile salts (GCDCA, TCDCA, TDCA and TUDCA) or the<br />

synthetic FXR agonists obeticholic acid (OCA) and GW4064<br />

for 4-8 hours. These cells were also co- treated with the FXR<br />

antagonist guggulsterone, and either FXR siRNA or control<br />

siRNA. Gene and protein expression was analyzed by qPCR<br />

and Western blotting. XBP1 splicing activity was measured<br />

using an XBP1-luciferase reporter construct that assays the 26-nt<br />

atypical splicing of the XBP1 mRNA. Results: Hepatic XBP1s<br />

and its downstream target ERdj4 mRNA expression increased<br />

6.5-fold (P < 0.05) and 1.7-fold (P < 0.05) in mouse livers after<br />

DCA feeding for 1 and 3 days, respectively. GCDCA (50 μM),<br />

TCDCA (10 μM) and TDCA (10 μM) treatment increased XBP1s<br />

protein expression in Huh7/NTCP cells, while TUDCA (10<br />

μM) had no effect. The synthetic FXR agonists OCA (2.5 μM)<br />

and GW4064 (2.5 μM) also induced XBP1s mRNA and protein<br />

expression. XBP1s protein induction by FXR agonists was<br />

blocked by the FXR antagonist guggulsterone (50 μM) and by<br />

FXR siRNA. The FXR agonists TCDCA (100 μM) and GW4064<br />

(10 μM) also increased XBP1 splicing-luciferase activity 2 and<br />

4-fold, respectively compared to controls (P < 0.001). Conclusions:<br />

DCA bile salt feeding to mice increases the expression<br />

of hepatic XBP1s and its downstream target ERdj4. FXR bile<br />

salt ligands and synthetic agonists increase XBP1s expression<br />

in Huh7/NTCP or HepG2 cells, while TUDCA has no effect.<br />

These effects are due, at least in part, to enhanced XBP1 splicing<br />

into XBP1s. Increased XBP1s expression is blocked by the<br />

FXR antagonist guggulsterone and by FXR knockdown. We<br />

demonstrate the novel finding that a FXR-mediated pathway<br />

increases the expression of liver XBP1s. We speculate that this<br />

may be important in the pathogenesis of cholestatic and fatty<br />

liver diseases.<br />

Disclosures:<br />

Richard M. Green - Consulting: McNeil<br />

The following authors have nothing to disclose: Xiaoying Liu, David Hilburn,<br />

Seong W. Park, Brian E. LeCuyer<br />

659<br />

Cerium oxide nanoparticle treatment enhances liver<br />

regeneration after two-thirds partial hepatectomy in<br />

rats<br />

Manuel Morales-Ruiz 1,3 , Altamira Arce-Cerezo 1,5 , Montserrat<br />

Pauta 1,5 , Jordi Ribera 1,3 , Denise Oró 1,3 , Gregori Casals 1,3 , Guillermo<br />

Fernández-Varo 1,3 , Tetyana Yudina 4 , Víctor Puntes 4 , Wladimiro<br />

Jiménez 1,2 ; 1 Biochemistry and Molecular Genetics, Hospital<br />

Clinic, Barcelona, Spain; 2 Department of Physiological Sciences I,<br />

IDIBAPS, CIBERehd, University of Barcelona School of Medicine,<br />

Barcelona, Spain; 3 IDIBAPS, CIBERehd, Barcelona, Spain; 4 Catalan<br />

Institute of Nanoscience and Nanotechnology, UAB, Bellaterra,<br />

Spain; 5 IDIBAPS, Barcelona, Spain<br />

Introduction and aim: Several <strong>studies</strong> have shown that oxidative<br />

stress impairs hepatic regeneration in mice. Therefore,<br />

testing new anti-oxidant drugs to improve liver regeneration<br />

has clinical interest. Recently, considerable attention has been<br />

paid to cerium oxide nanoparticles (CeO 2<br />

NPs) as a treatment<br />

for oxidative stress-related diseases. This interest relies on the<br />

multi-enzyme mimetic properties of CeO 2<br />

NPs that turn these<br />

nanoparticles into an effective scavenger of reactive oxygen<br />

species. Consequently, we aimed to investigate the effect of<br />

CeO 2<br />

NPs treatment on hepatic regeneration after partial hepatectomy.<br />

Methods: CeO 2<br />

NPs (4 nm in size), were synthesized<br />

by precipitation of a cerium (III) salt by the addition of ammo-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 537A<br />

nium salt tetramethylammonium-hydroxide (TmaOH). Forty-four<br />

male Wistar rats were divided equally into two groups. In the<br />

first group, rats were treated with 0.1 mg/kg CeO 2<br />

NPs i.v.<br />

twice a week for two weeks before 2/3 partial hepatectomy<br />

(PHx). The second group received vehicle as a control treatment.<br />

PHx was performed in both experimental groups and<br />

rats were sacrificed at different time points for further analysis.<br />

Results: Rats were sacrificed at day 6 after PHx and the ratio<br />

“liver remnant weight/total body weight” was calculated to<br />

obtain the liver regenerative index. Rats treated with CeO 2<br />

NPs<br />

showed a significant 11% increase in liver regeneration, compared<br />

with vehicle treated rats (p


538A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Emma A. Kruglov, Meena Ananthanarayanan,<br />

Jittima Weerachayaphorn, Pedro Sousa, Michael H. Nathanson<br />

662<br />

Effects of Ursodeoxycholic-acid and Rifampicin on autophagy<br />

in the liver<br />

Katrin Panzitt 1 , Hanns-Ulrich Marschall 2 , Michael Trauner 3 , Peter<br />

Fickert 1 , Martin Wagner 1 ; 1 Division of Gastroenterology and<br />

Hepatology, Department of Internal Medicine, Medical University<br />

Graz, Graz, Austria; 2 Wallenberg Laboratory, Department<br />

of Molecular and Clinical Medicine, Sahlgrenska Academy, University<br />

of Gothenburg, Gothenburg, Sweden; 3 Division of Gastroenterology<br />

and Hepatology, Department of Internal Medicine III,<br />

Medical University of Vienna, Vienna, Austria<br />

Background: Bile acids and activation of the bile acid receptor<br />

FXR inhibit autophagy, a cellular self-digestion process necessary<br />

for cell homeostasis and regeneration. The effects of<br />

chronic bile acid accumulation in chronic cholestatic liver disease<br />

on autophagy have not been studied in detail. However,<br />

indirect evidence (e.g. accumulation of Mallory-Denk bodies<br />

in primary biliary cirrhosis) indicates that autophagy may be<br />

impaired in human cholestasis. Hypothesis: We aim to determine<br />

whether ursodeoxycholic-acid (UDCA) and rifampicin<br />

(Rifa), two drugs for the treatment of human cholestatic liver<br />

disease may activate autophagy as a potential mode of drug<br />

action. Methods: Markers of autophagy (LC3, p62, ATG5,<br />

ATG7, ATG12) and the upstream mTOR signaling pathway<br />

(Raptor, ULK1, pS6K) have been studied by Western blot,<br />

immunofluorescence and qPCR in liver biopsies from patients<br />

treated with UDCA and Rifa. Mechanistic details of UDCA<br />

and Rifa action have further been studied in detail in human<br />

HepG2 cells and human primary hepatocytes using shPXR<br />

and shFXR knockdown strategies as well as conventional techniques<br />

to study autophagy (e.g. chloroquine treatment, LC3-<br />

RFP-GFP transfections). Results: Both, UDCA and Rifa induce<br />

LC3 II protein levels as the main autophagy readout in human<br />

liver biopsies. More detailed <strong>studies</strong> in hepatocyte cell lines<br />

using chloroquine treatment and LC3-RFP-GFP transfections<br />

reveal that true autophagic flux is induced. UDCA appears<br />

to activate autophagy via mTOR-ULK1 signaling whereas Rifa<br />

induces autophagy on transcriptional levels (LC3C, LAMP1,<br />

ATG10) without impacting on mTOR signaling. Furthermore,<br />

knockdown of the Rifa activated transcription factor PXR significantly<br />

represses autophagy already under basal conditions on<br />

mRNA and protein levels. In addition, PXR knockdown prevents<br />

Rifa induced autophagy induction. Summary and conclusions:<br />

UDCA and Rifa induce autophagy in the liver via different<br />

mechanisms. UDCA induces autophagy fux via mTOR signaling<br />

pathways whereas Rifa induces autophagy flux mTOR independently<br />

via the transcription factor PXR. Part of the beneficial<br />

effects of UDCA and Rifa in the treatment of cholestatic liver<br />

disease may be attributed to an induction of autophagy. Both<br />

compounds, UDCA and Rifa may have additional beneficial<br />

effect by inducing autophagy on other hepatological as well as<br />

non-hepatological diseases.<br />

Disclosures:<br />

Hanns-Ulrich Marschall - Consulting: Albireo<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Peter Fickert - Consulting: Falk Foundation, Falk Foundation, Falk Foundation,<br />

Falk Foundation; Speaking and Teaching: Roche Austria, Gilead Austria, MSD<br />

Austria, MERCK Austria, Roche Austria, Gilead Austria, MSD Austria, MERCK<br />

Austria, Roche Austria, Gilead Austria, MSD Austria, MERCK Austria, Roche<br />

Austria, Gilead Austria, MSD Austria, MERCK Austria<br />

The following authors have nothing to disclose: Katrin Panzitt, Martin Wagner<br />

663<br />

Involvement of Endoplasmic Reticulum Stress Transducer<br />

BBF2H7 in Hepatocellular Carcinoma Promotion<br />

Yizhou Zhang 1,2 , Nelson Hayes 1,2 , Masataka Tsuge 1,2 , Nobuhiko<br />

Hiraga 1,2 , Takuro Uchida 1,2 , Daiki Miki 3,2 , Hiromi Abe 1,2 , Kazuaki<br />

Chayama 1,2 ; 1 Department of Gastroenterology and Metabolism,<br />

Hiroshima University, Hiroshima City, Japan; 2 Liver Research Project<br />

Center, Hiroshima University, Hiroshima City, Japan; 3 RIKEN<br />

Center for Integrative Medical Sciences, Yokohama, Japan<br />

Objectives: Internal and external environmental stress brings<br />

about dysregulation of gene expression and genomic instabilities.<br />

The fundamental role of the tumor suppressor p53 in<br />

the response to various forms of stress ensures our genomic<br />

integrity. However, p53 is reported to be functional inactivated<br />

in almost all cancers. Endoplasmic reticulum (ER) stress<br />

is indicated as a factor contributing to cancer. So far, how<br />

effectors of ER stress cause Hepatocellular Carcinoma (HCC)<br />

is not well studied. Here we show that ER stress transducer<br />

BBF2H7 promotes ubiquitin-dependent p53 degradation and<br />

diminishes p53 phosphorylation, which in turn contributes to<br />

cell proliferation and survival of HCC. Methods: Expression of<br />

ER stress transducers in paired non-tumor and HCC tissues was<br />

measured by Real-Time PCR. Immunohistochemistry of HCC<br />

tissues was performed with BBF2H7 antibody. Cell growth of<br />

HepG2 cells with BBF2H7 overexpressed or knocked down<br />

was monitored by live cell imaging. Expression of p53 and<br />

its downstream targets were checked by Real-Time PCR and<br />

western blot in BBF2H7 overexpressed HepG2 cells, with or<br />

without the addition of proteasome inhibitor MG132. Dual<br />

luciferase reporter assay was performed to test how BBF2H7<br />

modulates p53 transcriptional activities. Nuclear abundance of<br />

p53 affected by BBF2H7 was also identified by immunostaining.<br />

Results: Among all the ER stress transducers, BBF2H7, an<br />

ERresident basic leucine zipper transcription factor, was found<br />

to be upregulated in HCC. IHC showed BBF2H7 accumulated<br />

in the nucleus of tumor cells, which is different from that in<br />

adjacent non-tumor tissues. Cell growth of HepG2 cells was<br />

promoted by BBF2H7 overexpression and was suppressed by<br />

BBF2H7 siRNA. Real-Time PCR results showed that BBF2H7<br />

overexpression increased MDM2 and BCL-2 expression, but<br />

p53 mRNA level was not changed. Protein levels of both p53<br />

and phos-p53 were decreased by BBF2H7 overexpression,<br />

but only p53 protein could be restored by blocking proteasome<br />

activity. The anti-apoptosis proteins BCL-2 and BCL-xL<br />

were up-regulated by BBF2H7, while Pol.δ cofactor PCNA was<br />

not affected. Strikingly, p53 transcriptional activity was drastically<br />

reduced by BBF2H7 overexpression. Contrary to the<br />

case of overall p53 protein, p53 transcriptional activity could<br />

barely be recovered by proteasome inhibitor. The impairment<br />

of p53 nuclear abundance by BBF2H7 might be the reason<br />

behind this. Conclusions: These findings reveal a new mechanism<br />

in which p53 can be destabilized by ER stress transducer<br />

BBF2H7, and also imply that BBF2H7 has full potential to trigger<br />

HCC carcinogenesis through impairing p53 transcriptional<br />

activity.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Asuka, Bayer, Daiichi Sankyo, Dainippon Sumitomo, Eisai, Janssen,<br />

Kowa, Mitsubishi Tanabe, MSD, Eli Lily, Nippon Kayaku, Nippon Shinyaku,<br />

Otsuka, Roche, Takeda, Toray, Torii, Tsumura, Zeria; Speaking and Teaching:<br />

Eisai, Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Asuka, Bayer, BMS,<br />

Chugai, Daiichi Sankyo, Dainippon Sumitomo, J&J, Jimro, Miyarisan, MSD,<br />

Nihon Kayaku, Olympus<br />

The following authors have nothing to disclose: Yizhou Zhang, Nelson Hayes,<br />

Masataka Tsuge, Nobuhiko Hiraga, Takuro Uchida, Daiki Miki, Hiromi Abe


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 539A<br />

664<br />

Restoration of Hepatic Physiological Parameters in vitro<br />

Influences Proprotein Convertase Subtilisin Kexin type 9<br />

(PCSK9) and LDL Regulation by Atorvastatin.<br />

Banumathi K. Cole, David Manka, Brian R. Wamhoff, Brett R.<br />

Blackman, Ajit Dash; HemoShear, Charlottesville, VA<br />

Background: Therapeutic targeting of Proprotein Convertase<br />

Subtilisin Kexin type 9 (PCSK9) in hyperlipidemia is attractive<br />

due to its ability to regulate hepatocyte LDL receptor levels and<br />

thereby clearance of circulating LDL cholesterol. Conventional<br />

hepatocyte models, however, are troubled by rapid dedifferentiation,<br />

loss of liver phenotype and function in culture, and<br />

non-translational drug dosing and exposure responses, challenging<br />

the relevance of these models and the accuracy of<br />

their biological response. We previously described a primary<br />

human hepatocyte system that uses liver-derived hemodynamics<br />

and physiological transport parameters to restore in vivo-like<br />

phenotype and hormone/drug responses at exposures matching<br />

physiological/clinical levels. Aims: The goal of this study<br />

was to validate the differences in regulation of PCSK9 biology<br />

and LDL uptake in response to atorvastatin, as they relate to differences<br />

in LDL exposure and physiological and conventional<br />

model conditions. Methods: Primary human hepatocyte tissues<br />

were cultured in collagen gel configurations and applied in 3<br />

conditions; ‘static’ non-perfused state, ‘hemodynamics’ with<br />

perfusion, or ‘perfusion’ only controls to account for LDL exposure<br />

effects. Cultures were exposed to 50μg/ml nLDL and atorvastatin<br />

for up to 48 hours. Endpoints included immunostaining<br />

for E-cadherin, measurement of intracellular protein levels of<br />

PCSK9, furin and HNF1α, secreted levels of PCSK9, and LDL<br />

uptake kinetics. Results: E-cadherin staining patterns revealed<br />

peripheral localization indicative of differentiated hepatocytes<br />

under hemodynamics but not in perfused only cultures. We<br />

found that continuous loading of nLDL, reflective of physiological<br />

exposure patterns in both the perfused and hemodynamic<br />

conditions, decreased baseline LDL uptake rates and PCSK9<br />

secretion relative to static cultures. The expected effect of atorvastatin<br />

on increasing PCSK9 secretion was noted only in the<br />

perfused and hemodynamic conditions but lost in static cultures.<br />

However, atorvastatin increased intracellular PCSK9 and<br />

LDL uptake in both static and hemodynamic conditions. Hemodynamics,<br />

but not perfusion alone, was also seen to upregulate<br />

key proteins, such as furin and HNF1α, that modulate<br />

PCSK9 expression and activity, which were not impacted by<br />

atorvastatin treatment. Conclusion: The results highlight the role<br />

of a combination of physiological milieu and hemodynamic<br />

parameters in resetting baseline biology and restoring PCSK9<br />

processing and secretion, which is important in elucidating the<br />

mechanisms surrounding PCSK9 regulation and the consequent<br />

development of therapeutic strategies for lowering circulating<br />

LDL levels.<br />

Disclosures:<br />

Banumathi K. Cole - Employment: HemoShear, LLC<br />

David Manka - Employment: HemoShear Therapeutics; Stock Shareholder:<br />

HemoShear Therapeutics<br />

Brian R. Wamhoff - Stock Shareholder: HemoShear, LLC<br />

Brett R. Blackman - Board Membership: HemoShear LLC; Management Position:<br />

HemoShear LLC; Patent Held/Filed: HemoShear LLC; Stock Shareholder:<br />

HemoShear LLC<br />

Ajit Dash - Employment: HemoShear LLC<br />

665<br />

The regulation of liver volume gain and regeneration-associated<br />

steatosis through Pten<br />

Ekaterina Kachaylo 1 , Christoph Tschuor 1 , Nathalie Borgeaud 1 ,<br />

Perparim Limani 1 , Michelangelo Foti 2 , Rolf Graf 1 , Bostjan Humar 1 ,<br />

Pierre A. Clavien 1 ; 1 Visceral and Transplant Surgery, University<br />

Hospital of Zurich, Zurich, Switzerland; 2 Département de Physiologie<br />

Cellulaire & Métabolisme, Centre Médical Universitaire,<br />

Geneva, Switzerland<br />

Successful liver regeneration is associated with the transient<br />

accumulation of lipids in hepatocytes, likely for the provision<br />

of energy to fuel the regenerative process. How this transient<br />

steatosis is regulated is ill-understood. The loss of Pten from<br />

liver induces rapid redistribution of peripheral fats into liver.<br />

Given that Pten also controls the growth-promoting Akt-mTORC<br />

pathway, Pten may act after resection to facilitate lipid metabolism<br />

and to drive liver volume gain. To explore this possibility,<br />

we are analysing inducible, hepatocyte-specific Pten knockout<br />

mice following hepatectomy. Pten levels were assessed in<br />

hepatectomized wild-type mice. AlbCre tg/+ PTEN fl/fl and Alb-<br />

Cre +/+ PTEN fl/fl (control) mice were studied shortly after tamoxifen<br />

induction. Regeneration was monitored via hepatic weight<br />

gain, proliferative activity and cyclin levels. Lipid content was<br />

quantified through a sulpho-phospho-vanillin protocol. In wild<br />

type mice, Pten levels dropped along with pronounced T389-<br />

Akt phosphorylation at 16h post resection, coinciding with the<br />

peak of transient steatosis. In parallel, fatty acid import (Cd36)<br />

and β-oxidation (Cpt1a, Hadha/b) increased, whilst lipogenesis<br />

(Acc, Scd, Fasn) decreased. Inducing Pten loss In AlbCre t-<br />

g/+ PTEN fl/fl mice resulted in spontaneous lipid accumulation<br />

associated with increased expression of lipogenic enzymes<br />

and elevation of serum triglycerides. However already at 16h<br />

after hepatectomy expression levels of Acc, Scd and Fasn and<br />

serum triglycerides were similar to controls and 3 days after<br />

resection, lipid content was comparable to controls, yet liver<br />

weight gain was accelerated. Notably, the increased liver<br />

weight gain was not accompanied by elevations in proliferative<br />

markers, suggesting an involvement of hypertrophy. Western<br />

blot analysis revealed increased S474-Akt2 phosphorylation at<br />

32 hours after hepatectomy in Pten ko livers vs controls, which<br />

indicates changes in glucose and lipid metabolism upon loss of<br />

Pten. Our results demonstrate a growth-promoting function for<br />

Pten downregulation during liver regeneration. The reduction<br />

in Pten may aid liver growth through (i) the release of inhibition<br />

of the hypertrophic Akt-mTORC pathway, and through (ii) the<br />

provision of energy via modulation of metabolism.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ekaterina Kachaylo, Christoph<br />

Tschuor, Nathalie Borgeaud, Perparim Limani, Michelangelo Foti, Rolf Graf,<br />

Bostjan Humar, Pierre A. Clavien<br />

666<br />

Organopollutant Exposures Further Decrease Epidermal<br />

Growth Factor Signaling in Non-Alcoholic Steatohepatitis<br />

and Alter Hepatic Energy Metabolism<br />

Josiah Hardesty 2 , Keith C. Falkner 1 , Heather B. Clair 2 , Banrida<br />

Wahlang 1 , Russell A. Prough 2 , Matthew C. Cave 1,3 ; 1 Department<br />

of Medicine/GI, University of Louisville, Louisville, KY; 2 Biochemistry,<br />

University of Louisville, Louisville, KY; 3 Robley Rex VAMC,<br />

Louisville, KY<br />

Background: Decreased Epidermal Growth Factor Receptor<br />

(EGFR) signaling has been associated with non-alcoholic steatohepatitis<br />

(NASH). We have previously demonstrated that<br />

the ubiquitous environmental pollutants, polychlorinated biphenyls<br />

(PCBs), are a ‘second hit’ in the transition of steatosis to


540A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

steatohepatitis. PCB-induced nuclear receptor activation only<br />

partially explained these findings, although the other ‘off target’<br />

mechanisms involved are unknown. These observations<br />

taken together led us to investigate the hypothesis that PCB-mediated<br />

EGFR inhibition alters downstream targets including the<br />

mechanistic target of rapamycin (mTOR) – a master regulator<br />

of hepatic energy metabolism in NASH. Methods: Mice<br />

were fed either a control diet (CD) or 42% milk fat high fat<br />

diet (HFD) and treated with either corn oil or Aroclor1260<br />

(PCB mixture, 20 mg/kg by one time gavage) for 12 weeks.<br />

Livers were removed and lysed for western blot analysis of<br />

pEGFR Y1173, EGFR, pmTOR, mTOR, and β-actin. In vitro<br />

cell culture assays were conducted with HepG2 and AML-12<br />

cells. Cells were treated with either EGF and EGFR inhibitor,<br />

or EGF alone, or in combination with Aroclor 1260; proteins<br />

were extracted for western blot analysis. Results: CD fed mice<br />

did not develop either steatosis or hepatic necro-inflammation.<br />

While HFD treated mice developed steatosis, only those co-exposed<br />

to HFD and Aroclor 1260 developed hepatic inflammation<br />

and steatohepatitis consistent with NASH. Co-exposed<br />

mice also had significant alterations in hepatic glucose and<br />

lipid metabolism. Analysis of liver samples showed an impact<br />

of diet on EGFR expression in which HFD decreased EGFR<br />

expression relative to CD. Aroclor-exposed mice demonstrated<br />

decreased phosphorylated EGFR at Y1173 independent of<br />

diet. Mice exposed to PCBs also had decreased phosphorylated<br />

mTOR and decreased total mTOR expression. In HepG2<br />

or AML-12 cells, Aroclor 1260 exposure was found to diminish<br />

EGFR phosphorylation at Y1173 relative to the positive control.<br />

Conclusion: These data demonstrate that HFD can lead<br />

to decreased hepatic EGFR expression and organopollutant<br />

exposure can act as a ‘second hit’ further decreasing EGFR<br />

signaling and altering mTOR activity. These may be the first<br />

data on environmental chemicals altering tyrosine kinase signaling<br />

in NASH. These findings require further study but may<br />

represent a new mechanism by which environmental pollution<br />

may contribute to NASH and the impaired liver regeneration<br />

associated with steatosis. Because mTOR inhibition is generally<br />

anti-inflammatory, additional mechanisms of PCB-induced<br />

hepatic inflammation should be investigated.<br />

Disclosures:<br />

Matthew C. Cave - Advisory Committees or Review Panels: Intercept, Abbvie;<br />

Consulting: Abbvie, Diapharma; Grant/Research Support: Merck, Gilead, Intercept,<br />

Conatus, Lumena, Cepheid, Tobira, Galectin, Bayer; Speaking and Teaching:<br />

BMS, Abbvie, Gilead, Janssen, Genentech<br />

The following authors have nothing to disclose: Josiah Hardesty, Keith C. Falkner,<br />

Heather B. Clair, Banrida Wahlang, Russell A. Prough<br />

667<br />

Skeletal muscle hyperammonemia causes mitochondrial<br />

functional abnormalities in cirrhosis<br />

Gangarao Davuluri 2 , Allawy Allawy 2,6 , Samjhana Thapaliya 2 ,<br />

Dharmvir Singh 2 , Avinash Kumar 2 , Julie H. Rennison 2 , Rafaella<br />

Nascimento e Silva 2 , Cynthia Tsien 3 , David R. Van Wagoner 4 ,<br />

Sathyamangla V. Naga Prasad 4 , Hoppel Charles 5 , Takhar Kasumov<br />

2 , Srinivasan Dasarathy 1 ; 1 Department Of Gastroenterology<br />

and Hepatology, Cleveland Clinic, Cleveland, OH; 2 Pathobiology,<br />

Cleveland Clinic, Cleveland, OH; 3 Gastroenterology, Toronto<br />

General Hospital, Toronto, ON, Canada; 4 Molecular Cardiology,<br />

Cleveland Clinic, Cleveland, OH; 5 Pharmacology and Medicine,<br />

Case Western Reserve University, Cleveland, OH; 6 Department Of<br />

Internal Medicine, Cleveland Clinic, Cleveland, OH<br />

Background. Skeletal muscle is a recognized metabolic partner<br />

to the liver for ammonia disposal and tissue specific adaptive<br />

responses occur during hyperammonemia. However, it is<br />

not known if increased metabolic and cellular stress induced<br />

by elevated ammonia uptake by the skeletal muscle results in<br />

defects in mitochondrial bioenergetic function. Methods. Skeletal<br />

muscle from human cirrhosis and controls, hyperammonemic<br />

portacaval anastomosis (PCA) rat and sham operated controls<br />

and myotubes exposed to hyperammonemia were used. Flow<br />

cytometry, using the fluorophores DCFDA and MitoSox, was<br />

used to measure reactive oxygen species (ROS) in C2C12<br />

myotubes during hyperammonemia. Oxidative modification of<br />

proteins by immublots for carbonylated proteins, lipid peroxidation<br />

by measuring thiobarbituric acid reactive substances<br />

(TBARS) and ATP content by fluorometric assays. The site of<br />

electron leak that generated the ROS was determined my quantifying<br />

H 2<br />

O 2<br />

production by oxidation of fluorogenic indicator<br />

amplexred in the presence of specific inhibitors (rotenone,<br />

malonate, antimycin A,oligomycin) of the electron transport<br />

chain. Cellular NAD+ and NADH were quantified by a fluorometric<br />

assay with and without inhibitors of the ETC to determine<br />

the mechanism of impaired electron flow. Results. Hyperammonemia<br />

increased mitochondrial ROS in C2C12 myotubes<br />

that was reversed by MitoTEMPO, a specific mitochondrial<br />

ROS scavenger. ATP content was lower, and carbonylated<br />

proteins and TBARS were significantly higher in the skeletal<br />

muscle from patients with cirrhosis and the PCA rat and in<br />

myotubes during hyperammonemia compared to the respective<br />

controls. Lowering cellular ammonia concentration reversed<br />

the low ATP content observed during hyperammonemia. Using<br />

sequential blockers of the various electron transport chain complexes,<br />

we demonstrate that Complex III was the major source<br />

of electron leak and ROS production production in the skeletal<br />

muscle during hyperammonemia. Finally, NAD+/NADH ratio<br />

was decreased with lower NAD+ and increased NADH due to<br />

impaired Complex I of the ETC NADH oxidase during hyperammonemia.<br />

Conclusions. Our <strong>studies</strong> in a comprehensive<br />

array of models show that skeletal muscle ammonia disposal<br />

activates a sequence of metabolic responses that culminate in<br />

disordered mitochondrial function. Notably, the reduction in<br />

ATP content is accompanied by electron leak at complex III<br />

and consequent generation of ROS during hyperammonemia.<br />

Reduction in ammonia can potentially reverse the mitochondrial<br />

bioenergetics dysfunction. These data provide evidence of<br />

a novel mechanism of hyperammonemia-induced perturbations<br />

in skeletal muscle cellular bioenergetics.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gangarao Davuluri, Allawy<br />

Allawy, Samjhana Thapaliya, Dharmvir Singh, Avinash Kumar, Julie H. Rennison,<br />

Rafaella Nascimento e Silva, Cynthia Tsien, David R. Van Wagoner,<br />

Sathyamangla V. Naga Prasad, Hoppel Charles, Takhar Kasumov, Srinivasan<br />

Dasarathy<br />

668<br />

Hyperammonemia activates a leucine responsive amino<br />

acid starvation stress response in the skeletal muscle<br />

Gangarao Davuluri 2 , Dawid Krokowski 3 , Bo-Jhih Guan 3 , Dharmvir<br />

Singh 2 , Allawy Allawy 2,6 , Avinash Kumar 2 , Rafaella Nascimento e<br />

Silva 2 , Ashok Runkana 2 , Samjhana Thapaliya 2 , Chenyang Zhao 4 ,<br />

Thomas Hamilton 4 , Sathyamangla V. Naga Prasad 5 , Maria Hatzoglou<br />

3 , Srinivasan Dasarathy 1 ; 1 Department Of Gastroenterology<br />

and Hepatology, Cleveland Clinic, Cleveland, OH; 2 Pathobiology,<br />

Cleveland Clinic, Cleveland, OH; 3 Department of Pharmacology,<br />

Case Western Reserve University, Cleveland, OH; 4 Immunology,<br />

Cleveland Clinic, Cleveland, OH; 5 Molecular Cardiology, Cleveland<br />

Clinic, Cleveland, OH; 6 Department Of Internal Medicine,<br />

Cleveland Clinic, Cleveland, OH<br />

Backgound. Skeletal muscle ammonia concentrations are<br />

increased in cirrhosis and results in sarcopenia. Decreased<br />

plasma and muscle concentrations of leucine and increased


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 541A<br />

plasma concentration of glutamine accompany cirrhosis. Here<br />

we show that hyperammonemia activates a specific molecular<br />

program in the skeletal muscle increasing the cellular demand<br />

for leucine that is transported into the mitochondria to potentially<br />

serve as a metabolic substrate for glutamine synthesis.<br />

Methods. Studies were performed in the skeletal muscle from<br />

patients with cirrhosis, hyperammonemic portacaval anastomosis<br />

rat, differentiated murine C2C12 myotubes during hyperammonemia<br />

appropriate controls. Myotubes were exposed<br />

to either ammonium acetate (10mM) or leucine (5mM) or a<br />

combination of leucine and ammonium acetate. Expression of<br />

mTOR, its target signaling proteins, global repressor of protein<br />

synthesis eIF2α and intracellular amino acid deficiency sensor,<br />

GCN2 expression were quantified by immunoblots. System L<br />

leucine transporter (LAT1) and glultamine transporter SLC38A2<br />

were quantified by real time PCR. Rate of protein synthesis was<br />

quantified using 3 H phenylalanine incorporation. Cellular and<br />

mitochondrial leucine uptake in myotubes was quantified using<br />

3 H leucine uptake and amino acids quantified using HPLC.<br />

Results. Skeletal muscle from cirrhotic patients and the PCA<br />

rat showed increased phosphorylation of eIF2α and reduced<br />

mTOR activation compared to that from respective controls.<br />

Murine myotubes showed activation of general control derepressed-2<br />

(GCN2) and its target, eIF2α, a global repressor of<br />

protein synthesis during hyperammonemia. This was accompanied<br />

by impaired mTOR signaling and reduced protein synthesis.<br />

Both the molecular perturbations and impaired protein<br />

synthesis were reversed by L-leucine supplementation in the<br />

medium. Lack of induction of GADD34 expression, an eIF2α<br />

phosphatase during hyperammonemia resulted in prolonged<br />

phosphorylation and activation of eIF2α. Increased expression<br />

of leucine exchanger, LAT1 but not glutamine transporter,<br />

SLC38A2 in both human muscle and murine myotubes were<br />

observed during hyperammonemia. Conclusions. We demonstrate<br />

a novel amino acid starvation response in the skeletal<br />

muscle during hyperammonemia that responds specifically to<br />

leucine supplementation despite increased leucine transport<br />

and intracellular concentrations. These observations combined<br />

with mitochondrial transport of leucine suggest that increased<br />

demand due to cataplerosis results in increased mitochondrial<br />

transport and activation of a specific program of translational<br />

repression mediated by persistent activation of eIF2α and<br />

impaired mTOR1 signaling.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gangarao Davuluri, Dawid Krokowski,<br />

Bo-Jhih Guan, Dharmvir Singh, Allawy Allawy, Avinash Kumar, Rafaella<br />

Nascimento e Silva, Ashok Runkana, Samjhana Thapaliya, Chenyang Zhao,<br />

Thomas Hamilton, Sathyamangla V. Naga Prasad, Maria Hatzoglou, Srinivasan<br />

Dasarathy<br />

669<br />

Orchestration of BH3-only proteins Bim, Bid and Puma<br />

controls hepatocyte apoptosis in Bcl-xL/Mcl-1 knockout<br />

mice<br />

Yoshinobu Saito, Hayato Hikita, Takahiro Kodama, Yuto Shiode,<br />

Yasutoshi Nozaki, Yugo Kai, Yuki Makino, Tasuku Nakabori,<br />

Satoshi Tanaka, Satoshi Aono, Ryotaro Sakamori, Naoki Hiramatsu,<br />

Tomohide Tatsumi, Tetsuo Takehara; Gastroenterology<br />

and Hepatology, Osaka University Graduate School of Medicine,<br />

Suita, Japan<br />

Background and Aim: Liver humanized mouse models using<br />

Alb-uPA SCID mice and Alb-HSVtk NOG mice have been<br />

reported to provide a lot of information on human liver physiology<br />

and pathology. However, these mouse models are not<br />

appropriate for investigating chronic inflammation and liver<br />

cancer because these mice are immune-deficient and could not<br />

survive more than one year. Immune tolerance is considered<br />

to be established during embryonic or neonatal stage. In this<br />

study, we investigated the engraftment efficacy of allogeneic or<br />

xenogeneic hepatocytes into the fetal liver by in utero transplantation<br />

in immune-competent mice. Method: As immune-competent<br />

recipient mice, we used hepatocyte-specific Mcl-1 knockout<br />

(KO) mice, which cause spontaneous hepatocyte apoptosis<br />

in life (Liver Injury mice), and hepatocyte-specific Mcl-1 and<br />

Bcl-xL double KO (DKO) mice, which show a decreased number<br />

of hepatocytes on embryonic (E) 18.5 day and die within<br />

one day after birth due to hepatic failure (Liver Impairment<br />

mice). Primary hepatocytes isolated from ubiquitously green fluorescent<br />

protein (GFP) transgenic (Tg) mice or primary human<br />

hepatocytes were administered into these recipient mice on E<br />

16.5 day via vitelline vein. These embryos were given birth by<br />

cesarean section and these livers were investigated at birth or<br />

at the age of 6 weeks. Results: Hepatocyte-transplanted wildtype<br />

mice and Liver Injury mice showed scattered embolization<br />

areas at birth by HE staining of the liver sections. The embolization<br />

was considered to be caused by transplanted hepatocytes.<br />

In contrast, Liver Impairment mouse livers escaped from<br />

embolization. In livers of wild-type mice and Liver Injury mice<br />

transplanted with GFP Tg hepatocytes, GFP-positive cells were<br />

observed forming a lot of clusters. While repopulation rate was<br />

not different between livers of wild-type mice and those of Liver<br />

Injury mice at birth (35.8% (3.6-48.3) and 29.8% (20.4-39.1),<br />

respectively), that of Liver Impairment was 7.8% (3.3-16.7).<br />

Meanwhile, repopulation rate at the age of 6 weeks was 0% in<br />

wild-type mice and 4.2% (2.1-8.0) in Liver Injury mice. In livers<br />

of wild-type and Liver Injury mice transplanted with primary<br />

human hepatocytes, embolization was similarly observed at<br />

birth by HE staining, suggesting that transplanted hepatocytes<br />

might be engrafted. Conclusion: Our results suggests that transplantation<br />

of primary hepatocytes into the fetal liver of Liver<br />

Injury mice, but not Liver Impairment mice, have the possibility<br />

of generating humanized liver mice without immunodeficiency.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yoshinobu Saito, Takahiro<br />

Kodama, Yuto Shiode, Yasutoshi Nozaki, Yugo Kai, Yuki Makino, Tasuku Nakabori,<br />

Satoshi Tanaka, Satoshi Aono, Ryotaro Sakamori, Naoki Hiramatsu,<br />

Tomohide Tatsumi<br />

670<br />

Sustained Inhibition of Liver Regeneration in Mice following<br />

Combined Elimination of MET & EGFR Signaling<br />

Shirish Paranjpe, William C. Bowen, Meagan Haynes, Anne Orr,<br />

Wendy M. Mars, Jianhua Luo, George K. Michalopoulos; Pathology,<br />

University of Pittsburgh, Pittsburgh, PA<br />

Acute loss of liver mass following toxic or surgical means elicits<br />

hepatic regeneration and restoration of liver tissue to its original<br />

mass, indicating a tightly regulated growth process. PHx<br />

in mice and rats results in rapid induction of more than 100<br />

genes, (that are not expressed in resting liver) & activation<br />

of multiple signaling pathways. EGFR & MET, the two major<br />

mitogenic receptor tyrosine kinases, are activated within 15 -<br />

30 minutes following a PHx. The MET-EGFR signaling pathway<br />

plays a significant role in is activated within 15-30 minute<br />

following a partial hepatectomy (PHx) in mice and rats. The<br />

role played by these two pathways during liver regeneration<br />

was investigated by utilizing mice that had c-met deleted in the<br />

liver using a tamoxiphen inducible Cre-loxP system. Mice were


542A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

injected with Tamoxiphen (1mg/ml) and injected i.p everyday<br />

for 5 days. Controls were injected with corn oil alone.This<br />

treatment deleted exon 16 that contains a critical ATP-binding<br />

site in the intracellular tyrosine kinase (TK) domain, essential for<br />

the activation of c-met. PhosphoEGFR was inhibited by using<br />

Canertinib, an irreversible pan-pEGFR inhibitor that was administered<br />

i.p at 80 mg/kg. Mice were injected a day before PHx<br />

and then everyday for 14 days. Appropriate vehicle controls<br />

were also used. In mice treated with Tyrosine kinase inhibitors,<br />

pEGFR levels were significantly reduced compared to vehicle<br />

treated controls. Analysis of proliferation indices indicated significant<br />

drop in KI67 positive staining in treated mice compared<br />

to controls Conclusion: Combined elimination of both MET and<br />

EGFR signaling resulted in sustained inhibition of liver regeneration<br />

up to 14 days.The PI3K-AKT-mTOR and the RAS-MEK-<br />

ERK pathways associated with cellular proliferation were down<br />

regulated in MET-EGFR inhibited mice.A significant reduction<br />

in Ki67 labeling was also evident, as was a reduction in liver to<br />

body weight ratio, suggesting a liver regeneration deficit.The<br />

increase in hepatocyte size in Met and EGFR suppressed livers<br />

may represent a compensatory effect to increase liver weight<br />

in the absence of hepatocyte proliferation.Suppression of MET-<br />

EGFR caused profound alterations in expression of many cell<br />

cycle associated genes. microRNA`s and LncRNA`s.<br />

Disclosures:<br />

George K. Michalopoulos - Consulting: Vital Therapies<br />

The following authors have nothing to disclose: Shirish Paranjpe, William C.<br />

Bowen, Meagan Haynes, Anne Orr, Wendy M. Mars, Jianhua Luo<br />

671<br />

Hyperammonemia impairs skeletal muscle protein synthesis<br />

via a myostatin dependent mechanism<br />

Srinivasan Dasarathy 1 , Gangarao Davuluri 2 , Samjhana Thapaliya<br />

2 , Avinash Kumar 2 , Gabriella A. Ten Have 3 , Dharmvir<br />

Singh 2 , Stephen L. Welle 4 , Marielle Engelen 3 , Sathyamangla<br />

V. Naga Prasad 5 , Nicolaas E. Deutz 3 ; 1 Department Of Gastroenterology<br />

and Hepatology, Cleveland Clinic, Cleveland, OH;<br />

2 Pathobiology, Cleveland Clinic, Cleveland, OH; 3 Department of<br />

Health and Kinesiology, Texas A&M University, College Station,<br />

TX; 4 Medicine, University of Rochester Medical Center, Rochester,<br />

NY; 5 Molecular Cardiology, Cleveland Clinic, Cleveland, OH<br />

Hyperammonemia is a consistent abnormality in cirrhosis and<br />

contributes to sarcopenia or skeletal muscle loss. We have<br />

previously shown that myostatin expression is increased in<br />

the hyperammonemic portacaval anastomosis (PCA) rat and<br />

in C2C12 myotubes in vitro. To demonstrate that hyperammonemia<br />

induced skeletal muscle loss is mediated via an<br />

increased expression of myostatin in vivo, we induced hyperammonemia<br />

in post developmental myostatin-/- and myostatin+/+<br />

mice and the metabolic and molecular signaling<br />

responses regulating protein synthesis (mTOR1 targets p70s6k,<br />

ribosomal s6 protein, 4EBP1) in the gastrocnemius muscle were<br />

studied. The post developmental myostatin knockout mice were<br />

generated by floxing the 3rd exon of myostatin gene (codes<br />

for entire active protein), a transgene encoding tamoxifen-activated<br />

Cre recombinase. At 4 months of age, mice were fed<br />

tamoxifen for 2 weeks to deplete myostatin. Hyperammonemia<br />

was induced by daily intraperitoneal injection of ammonium<br />

acetate for 4 weeks. Muscle protein synthesis was quantified<br />

by administration of a pulse of L-ring-D5 phenylalanine<br />

(D5Phe) and protein incorporation of labeled D5-Phe quantified<br />

using LC-MS/MS. Plasma (60.1±22.2 vs 378.0±71.2<br />

μmol/l; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 543A<br />

673<br />

Effects of Targeting Wnt/β-Catenin and RAS/RAF/<br />

MAPK Pathway on Hepatocellular Carcinoma Cell<br />

Growth and Metabolism: Potential for a New Combination<br />

Therapy<br />

Lilia Turcios 1 , Valery Vilchez 1 , David Butterfield 2 , Mihail I. Mitov 3 ,<br />

Francesc Marti 1 , Roberto Gedaly 1 ; 1 Surgery, Univ of Kentucky,<br />

Lexington, KY; 2 Chemistry, University of Kentucky, Lexington, KY;<br />

3 Markey Cancer Center - Core Support, University of Kentucky,<br />

Lexington, KY<br />

Background: Treatment of advanced hepatocellular carcinoma<br />

(HCC) remains a challenge. We aim to evaluate the impact<br />

of a β-catenin inhibitor, FH535, alone or in combination with<br />

a RAS/RAF/MAPK inhibitor, sorafenib, in liver cancer cells.<br />

Methods: Proliferation of liver cancer cell lines (Huh7, PLC,<br />

Hep3B) with or without addition of FH535 (10mM, 5mM,<br />

2.5mM) and sorafenib (2mM, 1 mM) alone or in combination<br />

was assayed by 3 H-thymidine incorporation at different time<br />

points (12h, 24h and 48h). The effect of drug treatment in mitochondrial<br />

function was determined by monitoring changes in<br />

mitochondrial transmembrane potential (ΔΨm) and mitochondrial<br />

respiratory activity. ΔΨm was assessed by tetramethylrhodamine<br />

ethyl ester (TMRE)-labeling and analyzed by flow<br />

cytometry. Mitochondrial respiration (OXPHOS) and glycolysis<br />

were measured by real-time analysis of oxygen consumption<br />

(OCR) and extra-cellular acidification (ECAR) rates, respectively,<br />

in a XF96 Flux Analyzer. Expression of β-catenin and<br />

apoptosis-related proteins were analyzed by Western Blot.<br />

Results: FH535 and sorafenib inhibited HCC cell proliferation<br />

in all three different cell lines tested (Huh7, PLC and Hep3B).<br />

We observed a significant dose-response inhibition with single<br />

treatments, although the effect of combined drugs was<br />

more potent. Our results also indicated the decreased levels of<br />

ΔΨm in cells exposed to the simultaneous exposure to FH535<br />

and sorafenib, which may reflect the efficient mitochondrial<br />

drug-targeting of the combined treatment. Mitochondrial dysfunction<br />

was confirmed by lower oxygen consumption rates<br />

and increased levels of apoptosis in cells treated with FH535<br />

plus sorafenib. Similar ECAR levels in treated and non-treated<br />

cells ruled out a broad cellular metabolic failure and rather<br />

confirmed the predominant mitochondrial effect of the drugs.<br />

Conclusions: The combination of FH535 and sorafenib inhibits<br />

HCC cell proliferation. Mechanistically, our results suggest that<br />

the combined therapeutic targeting of β-catenin and RAS/RAF/<br />

MAPK pathways may alter tumor cell growth by promoting<br />

mitochondrial dysfunction.<br />

Disclosures:<br />

The following authors have nothing to disclose: Lilia Turcios, Valery Vilchez,<br />

David Butterfield, Mihail I. Mitov, Francesc Marti, Roberto Gedaly<br />

674<br />

RA and butyrate induce liver and colon cancer cell<br />

apoptosis through miR-22 silencing of histone deacetylases<br />

Ying Hu, Hui-Xin Liu, Yu-Jui Yvonne Wan; Department of Medical<br />

Pathology and Laboratory Medicine, University of California,<br />

Davis Health Systems, Sacramento, CA<br />

Due to the signicant tumor suppressive role of microRNA-22<br />

(miR-22), it is important to establish the mechanism by which<br />

miR-22 expression is regulated. Our published data showed<br />

that bile acid-activated farnesoid x receptor (FXR) induces miR-<br />

22. In this report, we showed that all-trans retinoic acid (RA)<br />

and butyrate, normally present in digestive tract, also up-regulate<br />

miR-22. RA is a biologically active metabolite of vitamin<br />

A and a natural agonist for tumor suppressor retinoic acid<br />

receptor β (RARβ). Butyrate is a short chain fatty acid and<br />

histone deacetylase inhibitor (HDACi), produced by fiber-fed<br />

Gram-positive bacteria. Both agents can induce apoptosis in<br />

cancer cells. We tested the hypothesis that RA and butyrate<br />

act through miR-22 to inhibit HDACs and induce apoptosis in<br />

liver and colon cancer cells. Our data showed that RA/butyrate<br />

treatment synergistically promoted apoptosis and potently<br />

induced RARβ expression compared to single compound treatment<br />

in liver Huh7 and colon HCT116 cancer cells. miR-22<br />

induction by RA and butyrate was transcriptionally regulated<br />

by RARβ/RXRα and FXR/RXRα direct binding to a DR-5 motif<br />

located at -1568 to -1585 bp upstream of miR-22 based on<br />

transient transfection assays. Using 3’UTR luciferase assay,<br />

histone deacetylase 1 (HDAC1) was uncovered as a novel<br />

miR-22 target. Furthermore, miR-22 mimics also reduced the<br />

protein levels of HDAC4 and SIRT1, which are identified miR-<br />

22 targets in rat cardiomyocytes. Consistently, RA and butyrate<br />

reduced HDAC1, HDAC4, and SIRT1 protein levels in HCT116<br />

cells. In addition, the reduction in protein deacetylases was<br />

accompanied by increased histone acetylation in the RARβ<br />

regulatory region harboring a RA response element (DR-5).<br />

Histone modification of the RARβ gene could account for<br />

enhanced RARβ expression by RA/butyrate treatment. Moreover,<br />

miR-22 inhibitors abolished reduction of protein deacetylase<br />

levels and prevented apoptosis in HCT116 and Huh7 cells<br />

induced by RA/butyrate. Conclusion: RA and butyrate induce<br />

miR-22 to promote apoptosis of gastrointestinal cancer cells.<br />

miR-22 inhibition of HDACs is a novel pathway to explain the<br />

anti-carcinogenic effect of RARβ and FXR.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ying Hu, Hui-Xin Liu, Yu-Jui<br />

Yvonne Wan<br />

675<br />

IRAKM-Mincle axis contributes to alcohol-induced liver<br />

injury via inflammasome activation<br />

Hao Zhou 1 , Laura E. Nagy 2 , Xiaoxia Li 1 ; 1 Immunology, Lerner<br />

Research Institute, Cleveland Clinic Foundation, Cleveland, OH;<br />

2 Pathobiology, Lerner Research Institute, Cleveland Clinic Foundation,<br />

Cleveland, OH<br />

Alcohol-induced liver injury is driven by hepatocyte necrosis<br />

and increased endotoxin translocation into the hepatic portal<br />

system, which together trigger chronic inflammatory liver<br />

damage. In this study, we found that mice deficient in IRAKM,<br />

a proximal Toll-like receptor pathway molecule, are protected<br />

from alcohol-induced liver injury. In response to low doses<br />

of LPS, which better reflect physiologically relevant levels of<br />

endotoxemia, IRAKM mediates the up-regulation of Mincle, a<br />

receptor for danger signals released by damaged cells. Biochemical<br />

analysis revealed that low-dose LPS preferentially<br />

induces the formation of an IRAKM Myddosome, which in turn<br />

promotes MEKK3-dependent NFκB activation. Mincle-deficient<br />

mice are also protected from alcohol-induced liver injury. Ex<br />

vivo <strong>studies</strong> demonstrated that both IRAKM and Mincle are<br />

required for inflammasome activation by the endogenous Mincle<br />

ligand SAP130, which is a danger signal release by damaged<br />

hepatocytes. Taken together, this study reveals a novel<br />

IRAKM-Mincle axis that critically mediates the pathogenesis of<br />

alcohol-induced liver disease through inflammasome activation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hao Zhou, Laura E. Nagy, Xiaoxia<br />

Li


544A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

676<br />

Hepatic maturation of induced Pluripotent Stem Cells<br />

is regulated by paracrine signals from endothelial and<br />

mesenchymal stem cells in culture and during organoid<br />

formation<br />

Akihiro Asai 1 , Eitaro Aihara 2 , Tatsuki Mizuochi 1 , Kieran Phelan 1 ,<br />

Christopher Mayhew 1 , Pranavkumar Shivakumar 1 , Takanori<br />

Takebe 3 , James Wells 1 , Jorge A. Bezerra 1 ; 1 Pediatrics, Cincinnati<br />

Children’s Hospital Medical Center, Cincinnati, OH; 2 Department<br />

of Biology, University of Cincinnati, Cincinnati, OH; 3 Department<br />

of Regenerative Medicine, Yokohama City University, Yokohama,<br />

Japan<br />

Background: To recapitulate key stages of organ development,<br />

we engineered liver organoids using human induced pluripotent<br />

stem cells (iPSC), human mesenchymal stem cells (MSC), and<br />

human umbilical vein endothelial cells (HUVEC) as described<br />

previously. Because regulatory mechanisms of organoid maturation<br />

are not well defined, we investigated the differentiation<br />

patterns of liver organoids and tested if differentiation of iPSC<br />

requires direct contact with MSC and HUVEC. Methods: iPSC<br />

were differentiated to hepatic endoderm (iPS-HE) by Act,Wnt3a,DMSO,and<br />

Knock-Out Serum Replacement then co-cultured<br />

with MSC and HUVEC to generate 3D organoids (liver<br />

buds). Following implantation under kidney capsules of mice,<br />

plasma levels of human albumin (ALB) and alpha1-antitrypsin<br />

(A1AT) were monitored by ELISA. To investigate whether cellcell<br />

contact is a key mechanism by which MSC and HUVEC<br />

induce bud formation and maturation of iPS-hepatocytes,<br />

we cultured iPS-HE in the upper well of a transwell system,<br />

where MSC or HUVEC resided in the lower well individually<br />

or in co-culture. The expression of alphafetoprotein (AFP),<br />

HNF4a,ALB,A1AT,CPS1,BSEP,and CD31 was determined by<br />

immunostaining or by ELISA. Results: Temporal-spatial analysis<br />

by time-lapse microscopy revealed self assembly of iPS-HE,<br />

MSC, and HUVEC into a 3D tissue in 24 hr following a condensation<br />

and folding pattern to form a concave hemisphere.<br />

Hepatocytic differentiation in the buds was demonstrated by<br />

the AFP+/HNF4+ cells by whole-mount staining, in proximity to<br />

CD31+ endothelial cells. After implantation of the buds, human<br />

ALB (140-260ng/ml) and A1AT (62-174ng/ml) were detected<br />

in mouse plasma. Explanted organoids showed hepatocyte-like<br />

morphology organized in cellular clusters surrounded by vascular<br />

networks. Using the transwell culture to determine the role<br />

of paracrine effect, we found that ALB concentration in culture<br />

supernatants achieved 1587±233ng/ml/day in iPS-HE with<br />

MSC in the lower well (iPS/MSC) and 1881±162ng/ml/day<br />

in iPS-HE/HUVEC, which were higher than iPS-HE cultured with<br />

MSC+HUVEC in the lower (455±110ng/ml/day) or iPS-HE<br />

alone (81±21ng/ml/day, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 545A<br />

678<br />

Notch and Wnt control ductular reactions via the Igf1<br />

axis<br />

Sarah E. Minnis-Lyons 1 , Luke G. Boulter 2 , Tak Yung Man 1 , Michael<br />

J. Williams 1 , Rachel V. Guest 1 , Wei-Yu Lu 1 , Noemi Van Hul 4 , Isabelle<br />

A. Leclercq 4 , Owen J. Sansom 3 , Stuart J. Forbes 1 ; 1 MRC Centre<br />

for Regenerative Medicine, Edinburgh, United Kingdom; 2 MRC<br />

Human Genetics Unit, Edinburgh, United Kingdom; 3 Beatson Institute<br />

for Cancer Research, Glasgow, United Kingdom; 4 Universite<br />

Catholique de Louvain, Brussels, Belgium<br />

Ductular reactions (DR) are seen in chronic liver injury when<br />

hepatocyte regeneration is impaired and are thought to contain<br />

putative hepatic progenitor cells (HPCs). Their ability to contribute<br />

to large-scale parenchymal repair has been questioned.<br />

Notch and Wnt are key signals required for liver development<br />

and regeneration. Given their many actions include promoting<br />

cell expansion, we sought to identify if these signals control<br />

DRs after hepatocyte injury and importantly whether these<br />

pathways can be manipulated for future therapeutic targeting.<br />

Notch and Wnt pathways were analysed using a genetic in<br />

vivo model of hepatocellular injury and ductular activation that<br />

has allowed us to examine the temporal dynamics of the ductular<br />

response (AhCre MDM2fl/fl). We have also used K19<br />

and OPNCre lineage tracing tools, as well as HPC lines, to test<br />

small molecules, blocking antibodies and genetic loss of function<br />

in order to establish functionality and interactions of these<br />

signalling pathways. We have found distinct time-dependent<br />

Notch and Wnt signatures following hepatocyte injury. Small<br />

molecule inhibitors have revealed time-sensitive windows; with<br />

the early ductular proliferative response principally driven by<br />

Notch and the later response driven by Wnt. Anti-Notch1<br />

and Notch2 blocking antibodies and genetic loss of Notch3<br />

have confirmed DRs are driven by Notch1 and Notch3 but<br />

not Notch2. We have identified DRs as an additional source<br />

of the potent growth hormone Igf1 and this production is Wnt<br />

driven. Notch driven expression of Igf1-receptor within DRs<br />

highlights this axis as a node for cooperation between Notch<br />

and Wnt signals. Small molecule (AG1024) and genetic loss<br />

of function (Igf1 receptor fl/fl) confirms functionality of this axis<br />

and stresses a pivotal role for Igf1-receptor in the generation of<br />

the ductular reaction which can be enhanced by administration<br />

of recombinant Igf-1 (all P values =


546A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

However, the potential of hepatocytes to stably and functionally<br />

contribute to the biliary system is uncertain. Here, we use a<br />

mouse model of Alagille syndrome (ALGS) that lacks peripheral<br />

intrahepatic bile ducts (pIHBDs) to determine whether hepatocytes<br />

can transdifferentiate into normal cholangiocytes and<br />

assemble into functional bile ducts. In our ALGS mouse model<br />

genes encoding the NOTCH DNA binding partner RBP and the<br />

biliary transcription factor HNF6 are inactivated in embryonic<br />

liver progenitors by Cre-mediated recombination. Like humans<br />

with severe ALGS these mice lack pIHBDs at birth; remarkably,<br />

however, pIHBDs are formed in our ALGS mouse model with<br />

age. To determine whether hepatocytes are the source of the<br />

new pIHBDs in our Cre-based mouse model, we developed a<br />

Flp recombinase-mediated method of hepatocyte fate tracing.<br />

By following fate-traced hepatocytes through de novo pIHBD<br />

formation we determined that hepatocytes convert into cholangiocytes<br />

and undergo tubulogenesis. This results in a fully functional<br />

hepatocyte-derived biliary system capable of reverting<br />

cholestasis and liver injury. We found little clonal expansion of<br />

hepatocyte-derived cholangiocytes, which suggests that these<br />

cells do not transition through a transit amplifying cell stage,<br />

but rather many hepatocytes transdifferentiate and are incorporated<br />

into the new pIHBDs. Interestingly, we observed that<br />

biliary markers are activated in the hepatocytes near hypoxic<br />

regions, suggesting that local hypoxia may induce hepatocyte<br />

bile duct formation. Indeed, we show that activation of hypoxia<br />

signaling in hepatocytes in vivo can induce biliary differentiation.<br />

Our results demonstrate that hepatocytes can generate<br />

an intrahepatic biliary system effective in relieving liver injury<br />

in a mouse model of severe ALGS. These results define the<br />

plasticity of hepatocytes and add another dimension to the<br />

remodeling capacity of the adult liver. In addition, hepatocyte<br />

transdifferentiation independent of NOTCH signaling in our<br />

ALGS model suggests the existence of alternative drivers of<br />

biliary differentiation, like hypoxia signaling. Bile duct formation<br />

by hepatocytes may explain spontaneous improvement of<br />

cholestasis observed in some ALGS patients and could potentially<br />

be developed into an in vivo-reprogramming-based therapy<br />

for others.<br />

Disclosures:<br />

Johanna R. Schaub - Management Position: MDsave<br />

The following authors have nothing to disclose: Kari A. Huppert, Ashley E. Cast,<br />

Feng Chen, Stacey S. Huppert, Holger Willenbring<br />

681<br />

Identification of hepatic stem cell niche by label retaining<br />

cell assay with fetal and neonatal mice<br />

Reiichiro Kuwahara 1 , Neil D. Theise 2 , Takuji Torimura 1 ; 1 Department<br />

of Medicine, Division of Gastroenterology, Kurume University<br />

School of Medicine, Kurume, Japan; 2 Department of Pathology<br />

and Medicine, Beth Israel Medical Center of Albert Einstein College<br />

of Medicine, New York, NY<br />

Background and Aim: Previously we showed, in an acetaminophen<br />

injury model, that oval cells (OVc) appear to derive from<br />

the most proximal portions of the biliary tree (canals of Hering:<br />

CoH) in a time- and dose-dependent manner, and regenerating,<br />

differentiated peribiliary hepatocytes (PbH) also contributed<br />

to parenchymal repair (Kofman et al, Hepatology 2005;<br />

41: 1252). Subsequent label retaining cell (LRC) assays supported<br />

these findings (Kuwahara et al, Hepatology 2008; 47:<br />

1994). We have now investigated whether LRC assays could<br />

shed light on the possible role of CoH and PbH in normal liver<br />

development, in the absence of injury. Methods: Pregnant mice<br />

received water containing BrdU for scheduled term of 3 days<br />

(E10-12, E13-15, and E16-19). For the confirmation of BrdU<br />

labeling in the fetal liver, some mice were sacrificed immediately<br />

after each labeling term. The other mice were allowed to<br />

give birth to mice. Neonatal mice were raised to 8 weeks old<br />

to “chase” for label washout and then sacrificed. Other mice<br />

received intraperitoneal injection of BrdU for labeling at the<br />

age of 6, 13, 20, 27, 34, 41, 48 and 55 days. Then, some<br />

of them were sacrificed at 24 hours after each injection for the<br />

analyses of cell division. Other mice were raised to 8 weeks<br />

old to “chase” for label washout following the BrdU labeling<br />

and then sacrificed. PbH and biliary cell in CoH were analyzed<br />

by double immunostaining for biliary keratins (PanK)/<br />

BrdU (cell division), PanK/Ki-67 (proliferation). Results: Almost<br />

all cells in the fetal liver were labelled with maternally administered<br />

BrdU in drinking water. In all mice with BrdU labeling<br />

during pregnancy, neither label retaining PbH nor label retaining<br />

biliary cells in CoH were detected in the liver of 8 weeks.<br />

The Ki67 index of neonatal mice liver showed that frequencies<br />

of cell proliferation of hepatocytes and biliary cells in CoH<br />

were highest at the age of 7 days and then diminished steadily.<br />

At 8 weeks post “chase”, distributions of BrdU-retaining PbH<br />

and biliary cells in CoH to total BrdU positive cells (lobular and<br />

peribiliary hepatocytes + biliary cells in CoH) were highest in<br />

the liver with BrdU labeling at the age of 13 days. These distributions<br />

were significant higher than those of liver with labeling<br />

at the age of 6, 20 and 27 days, respectively. Conclusion:<br />

These data provides support that PbH and CoH cells serve resident<br />

stem cell functions including contributing to parenchymal<br />

mass during organogenesis and early post-natal development<br />

(< 3weeks), but not after.<br />

Disclosures:<br />

The following authors have nothing to disclose: Reiichiro Kuwahara, Neil D.<br />

Theise, Takuji Torimura<br />

682<br />

Disruption of TGF-β-regulated CTCF Suppression of<br />

Telomerase links a Human Stem Cell Disorder to Liver<br />

Tumorigenesis<br />

Jian Chen 1 , Jiun-Sheng Chen 1 , Young Jin Gi 1 , H. Franklin Herlong 1 ,<br />

Yun Seong Jeong 1 , Nipun Mistry 1 , Xiaoping Su 1 , Asif Rashid 1 ,<br />

Bibhuti Mishra 1 , Jon White 2 , Milind Javle 1 , Marta L. Davila 1 , John<br />

R. Stroehlein 1 , Rosanna Weksbergc 3 , Jerry W. Shay, 4 , Keigo<br />

Machida 5 , Hidekazu Tsukamoto 5 , Lopa Mishra 1 ; 1 The University<br />

of Texas MD Anderson Cancer Center, Houston, TX; 2 Institute of<br />

Clinical Research, Veterans Affairs Medical Center, Washington<br />

DC, DC; 3 Hospital for Sick Children, Toronto, ON, Canada; 4 The<br />

University of Texas Southwestern Medical Center, Dallas, TX; 5 University<br />

of Southern California, Los Angeles, CA<br />

Patients with a human stem cell disorder, the Beckwith-Wiedemann<br />

syndrome (BWS) are known to develop multiple liver<br />

tumor types (hepatoblastoma, hepatocellular cancer, and cholangiocarcinoma)<br />

within a single patient. These patients are at<br />

an 800 fold increased risk of tumorigenesis. However, a precise<br />

mechanism for the switch in tumorigenesis remains elusive.<br />

In a previous study we demonstrated that loss of Transforming<br />

Growth Factor-beta (TGF-β) signaling, was a causal factor in<br />

BWS. To determine the mechanism for the oncogenic switch,<br />

we performed a broad bioinformatics and functional analyses<br />

utilizing human BWS cells, tissues and our mutant models in the<br />

TGF-β pathway. Methods and Results: (1) Over 80% of TGF-β/<br />

Sptbn1 +/− /Smad3 +/− mutant mice spontaneously developed<br />

multiple tumors including the liver tumors that were phenotypically<br />

similar to those of patients with BWS. (2) Somatic mutations<br />

in SPTBN1 and SMAD3 are among the most frequent in<br />

human HCCs. (3) Both Sptbn1 +/− /Smad3 +/− mice and BWS<br />

cells express increased levels of stem cell-associated genes,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 547A<br />

including Nanog, Sox2 and Oct4. (4) Increased levels of several<br />

stem-ness genes and telomerase (TERT) (25-45 fold), c-Myc<br />

(20-100 fold), IGF2 (~20 fold), with decreased levels of molecules<br />

implicated in BWS- P57 and CTCF were observed in liver<br />

tumors that developed in Sptbn1 +/- /Smad3 +/- mice. (5) Disruption<br />

of the β2SP/SMAD3/CTCF complex raises TERT levels<br />

in BWS and in Sptbn1 +/− /Smad3 +/− mice. (6) We observed<br />

that disruption of the β2SP/SMAD3/CTCF complex increases<br />

stem-like properties (increased ALDH-positive population and<br />

Sphere formation) and enhances tumorigenesis in human HCC<br />

cell lines. (7) We identified a mechanism by which the long<br />

range chromatin modulator CTCF facilitates TGF-β-mediated<br />

repression of hTERT transcription via β2S/SMAD3/CTCF interactions.<br />

Conclusions: We present new insights into liver tumor<br />

formation through the human stem cell disorder BWS and our<br />

mutant models in the TGF-β pathway, demonstrating a potential<br />

shift arising from disruption of TGF-β/CTCF signaling leading<br />

to BWS-associated tumorigenesis, telomerase activation,<br />

and human stem cell associated cancer development. This syndrome<br />

can provide new strategies for preventing and targeting<br />

liver cancers.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jian Chen, Jiun-Sheng Chen,<br />

Young Jin Gi, H. Franklin Herlong, Yun Seong Jeong, Nipun Mistry, Xiaoping<br />

Su, Asif Rashid, Bibhuti Mishra, Jon White, Milind Javle, Marta L. Davila, John<br />

R. Stroehlein, Rosanna Weksbergc, Jerry W. Shay,, Keigo Machida, Hidekazu<br />

Tsukamoto, Lopa Mishra<br />

683<br />

Self-renewing diploid Axin2+ cells fuel homeostatic<br />

renewal of the liver<br />

Bruce M. Wang 1 , Roel Nusse 2,3 ; 1 Medicine, UCSF Liver Center,<br />

San Francisco, CA; 2 Developmental Biology, Stanford University,<br />

Stanford, CA; 3 Howard Hughes Medical Institute, Stanford, CA<br />

The cellular source of new hepatocytes in the adult liver and<br />

the molecular regulation of hepatocyte renewal are fundamental<br />

unanswered questions in liver biology. While it has been<br />

shown that new hepatocytes in the uninjured liver arise from<br />

pre-existing hepatocytes, hepatocytes are known to be heterogenous<br />

with striking differences in age and function across<br />

the liver lobule. It is unknown whether a specific subpopulation<br />

of hepatocytes serves homeostatic renewal in the liver. Using<br />

a novel mouse model for stably labeling Wnt-responsive cells<br />

in vivo, we have discovered a unique population of Wnt-controlled<br />

hepatocytes with stem cell characteristics surrounding<br />

the central vein. These pericentral cells express the early liver<br />

progenitor marker Tbx3, proliferate at a faster rate than other<br />

hepatocytes and are diploid, and thus differ from mature<br />

hepatocytes, which are mostly polyploid. Over time, these<br />

cells self-renew and give rise to descendants that differentiate<br />

into Tbx3-negative, polyploid hepatocytes and can replace all<br />

hepatocytes along the liver lobule during homeostatic renewal.<br />

Importantly, these Wnt-responsive cells are present in the normal<br />

liver, thereby distinguishing them from injury-induced multipotent<br />

progenitor cells (oval cells) associated with the portal<br />

region. Adjacent central vein endothelial cells provide essential<br />

Wnt signals that maintain the pericentral cell population,<br />

thereby constituting the hepatocyte stem cell niche. Thus, we<br />

describe for the first time a subset of hepatocytes that subserves<br />

homeostatic hepatocyte renewal, characterize its anatomical<br />

niche, and identify molecular signals that regulate its activity.<br />

Disclosures:<br />

The following authors have nothing to disclose: Bruce M. Wang, Roel Nusse<br />

684<br />

FAP disease modelling using patient-specific iPS cells<br />

derived from urine<br />

Christoph J. Niemietz, Vanessa Sauer, Jacquelline Stella, Gursimran<br />

Chandhok, Andree Zibert, Hartmut H. Schmidt; Klinik für<br />

Transplantationsmedizin, Universitätsklinikum Münster, Münster,<br />

Germany<br />

Transthyretin-related familial amyloid polyneuropathy (TTR-<br />

FAP) is a rare genetic neurodegenerative disease caused by<br />

mutations of TTR. TTR is primarily secreted by the liver and<br />

misfolding ultimately leads to aggregation and fibril deposition<br />

in peripheral tissues and organs. In order to prevent expression<br />

of TTR, gene silencing strategies are currently under clinical<br />

investigation. The use of primary cells derived from FAP<br />

patients for evaluation of antisense strategies is difficult. As<br />

an alternate, stem cell technology is used here to generate<br />

patient specific primary cells. Urine collections of FAP patients<br />

were processed for isolation of renal epithelial cells, followed<br />

by reprogramming into iPS cells (iPSCs) using non-integrating<br />

episomal vectors. After characterization of pluripotent cell lines,<br />

differentiation towards hepatocyte-like cells was accomplished<br />

after treatment with factors for 14 days. iPSC-Heps were characterized<br />

by analysis of typical hepatic markers via qRT-PCR,<br />

flow cytometry, and immunocytochemistry. ASOs and siRNAs<br />

were introduced into iPSC-Heps in order to evaluate gene<br />

silencing of TTR via qRT-PCR analysis, ELISA and Western-blot.<br />

Typically, 2-3 stable cell colonies emerged upon cultivation<br />

of urine-derived cells from FAP patients (n=12). iPSC-Heps<br />

showed similar morphology and gene expression in comparison<br />

to HepG2 control. TTR gene expression was high. ASOs or<br />

siRNA treatment resulted in downregulation of TTR. Our results<br />

indicate that iPSCs derived from urine cells of FAP patients can<br />

be routinely reprogrammed and differentiated into iPSC-Heps<br />

expressing TTR at comparable levels as control cells. TTR gene<br />

silencing analysis of iPSC-Heps allows the establishment of a<br />

patient-specific in vitro platform for evaluating drug efficiency.


548A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Thus, the technology is excellently suited for modelling of liver<br />

disease and may allow assessment of novel drugs for improved<br />

therapy of FAP.<br />

Disclosures:<br />

The following authors have nothing to disclose: Christoph J. Niemietz, Vanessa<br />

Sauer, Jacquelline Stella, Gursimran Chandhok, Andree Zibert, Hartmut H.<br />

Schmidt<br />

685<br />

CD34+ Liver Cancer Stem Cells were Formed by Fusion<br />

of Hepatobiliary Stem/Progenitor Cells with Hematopoietic<br />

Precursor-Derived Myeloid Intermediates<br />

Changjun Zeng 2 , Yanling Zhang 2 , Su Cheol Park 2 , Jong Ryeol<br />

Eun 2 , Ngoc Tue X. Nguyen 2 , Benjamin Tschudy-Seney 2 , Yong<br />

Jin Jung 2 , Neil D. Theise 3 , Mark Zern 2 , Yuyou Duan 1 ; 1 Dermatology,<br />

Internal Medicine, Institute for Regenerative Cures, UC Davis<br />

Medical Center, Sacramento, CA; 2 Internal Medicine, Institute for<br />

Regenerative Cures, University of California Davis Medical Center,<br />

Sacramento, CA; 3 Beth Israel Medical Center, Albert Einstein College<br />

of Medicine, New York, NY<br />

A large number of cancer stem cells (CSC) were identified and<br />

characterized; however, the origins and formation of these<br />

CSC remain elusive. Here, we examined the origination of<br />

newly-identified CD34+ liver CSC (LCSC). CD34+ LCSC was<br />

clonogenically cultured in vitro employing our newly-developed<br />

medium, and cloned CD34+ LCSC was used to inject<br />

into NOD/SCID mice to generate the xenografts of human<br />

liver carcinomas (HLC). Our results that CD34+ LCSC co-expressed<br />

liver stem cell and myelomonocytic cell markers, showing<br />

a mixed phenotype of hepatobiliary stem/progenitor cells<br />

(HSPC) and myelomonocytic cells, as determined by both qPCR<br />

and Immunohistochemistry. Moreover, human xenografts and<br />

parental cells which CD34+ LCSC was isolated, co-expressed<br />

liver cancer and meylomonocytic markers, also demonstrating<br />

mixed phenotypes. ELISA assay showed that the xenografts<br />

and parental cells secreted albumin demonstrating its hepatocyte<br />

origin, and these cells also expressed cytokines (IL-1b, IL-6,<br />

IL-12A, IL-18, TNF-α, and CSF1) and chemokines (IL-8, CCL2,<br />

and CCL5), it is well known that the macrophage is the primary<br />

cell which produces these factors. Expression of these cytokines<br />

and chemokines responded to the stimuli (INF-γ, IL-4, and LPS),<br />

response to stimulation by these three factors is a unique macrophage<br />

characteristic. Furthermore, human xenografts and<br />

the parental cells phagocytized E. Coli; thus, they appeared to<br />

express characteristics of the two parental cell types as hybrid<br />

cells do. CD34+ LCSC co-expressed CD45, demonstrating its<br />

origin appears to be from a hematopoietic precursor. The percentage<br />

of cells positive for OV6, CD34, and CD31 presenting<br />

the markers of HSPC, hematopoietic and myelomonocytic<br />

cells, were increased under treatment of CD34+ LCSC with<br />

drug cisplatin. Cytogenetic analysis showed that CD34+ LCSC<br />

contained a greater number of chromosomes (aneuploid).<br />

HBV DNA integrations and mutations in CD34+ LCSC and the<br />

parental cells, were identical to those in the literature published<br />

three decades ago or those in the Sanger COSMIC database.<br />

In conclusion, these results demonstrated that CD34+ LCSC<br />

were formed by fusion of HSPC with CD34+ hematopoietic<br />

precursor-derived myeloid intermediates, thus, we believe that<br />

this is the first report that human CSC appears to have been<br />

formed by the fusion, and it is also the first report that a HLC<br />

appears to be initiated and developed from the interaction of<br />

HSPCs with CD34+ hematopoietic precursors, thus revealing a<br />

diversity of origins for HLC. Therefore, our results represent a<br />

significant step towards better understanding of the formation<br />

of human CSC, and the diverse origins of human cancers.<br />

Disclosures:<br />

The following authors have nothing to disclose: Changjun Zeng, Yanling Zhang,<br />

Su Cheol Park, Jong Ryeol Eun, Ngoc Tue X. Nguyen, Benjamin Tschudy-Seney,<br />

Yong Jin Jung, Neil D. Theise, Mark Zern, Yuyou Duan<br />

686<br />

TGF-β inhibits lncRNA H19 expression via SOX2 in<br />

tumor-initiating hepatocytes<br />

Jinqiang Zhang, Chang Han, Weina Chen, Kyoungsub Song, Ying<br />

Wang, Lu Yao, Nathan Ungerleider, Hyunjoo Kwon, Tong Wu;<br />

Pathology and Laboratory Medicine, Tulane University, New Orleans,<br />

LA<br />

TGFβ pathway modulates various biological processes. The<br />

output of a TGFβ response is highly depending on tissue types<br />

and diseases. This is especially true in hepatocellular carcinoma<br />

(HCC) initiation and progression, where the tumor-initialing<br />

and malignant cells are highly influenced by the liver<br />

environments. To delineate the role of TGFβ in the process<br />

of hepatocarcinogenesis, we employed a novel tumor-initiating<br />

cell transplantation system that avoids the possible effects<br />

incurred from genetic manipulation of background liver. Specifically,<br />

TGFβ receptor II (Tgfbr2) flox/flox mice at 14 days of<br />

age were given one dose (25mg/g) of the hepatic carcinogen<br />

diethylnitrosamine via intraperitoneal injection. Three months<br />

later, the initiated hepatocytes from Tgfbr2 flox/flox mice<br />

were isolated and transplanted to the same strain (C57Bl/6)<br />

wild type recipient mice. Four weeks later, the recipient mice<br />

received one does of Cre recombinase adenovirus (Ad-Cre)<br />

through tail vein to delete Tgfbr2 in the transplanted tumor-initiating<br />

hepatocytes; separate group of the recipient mice were<br />

injected with the control adenovirus (Ad-GFP). The recipient<br />

mice were closely monitored for liver tumor burden. We<br />

observed that deletion of Tgfbr2 in tumor-initiating hepatocytes<br />

by tail vein injection of Ad-Cre led to formation of more and<br />

bigger liver tumors (compared to Ad-GFP group). This finding<br />

suggests that TGFβ inhibits the malignant potential of tumor-initiating<br />

hepatocytes. In parallel, we further analyzed the effect<br />

of TGFβ in tumor initiating hepatocytes in vitro. Specifically,<br />

tumor-initiating hepatocytes isolated from DEN-treated Tgfbr2<br />

flox/flox mice were infected with Ad-Cre or Ad-GFP prior to<br />

TGFβ1 treatment in vitro; RNAs were then isolated for transcriptome<br />

sequencing analysis. We identified a group of lncRNAs<br />

that were noticeably regulated by TGFβ, including the lncRNA<br />

H19. We found that deletion of Tgfbr2 by Ad-Cre in tumor initiating<br />

hepatocytes led to a 5-fold increase of H19 expression.<br />

We observed that deletion of H19 by siRNA decreased the<br />

tumor-initiating cell property, whereas forced overexpression<br />

of H19 enhanced it. Chromatin immunoprecipitation assay<br />

showed that SOX2 bound to the promoter region of H19 gene.<br />

While SOX2 overexpression in tumor-initiating hepatocytes<br />

enhanced H19 transcription, SOX2 knockdown reduced it.<br />

Furthermore, TGFβ treatment reduced the protein level of SOX2<br />

in tumor-initiating hepatocytes. Taken together, our findings<br />

disclose a novel mechanism that importantly regulates hepatocarcinogenesis<br />

—- TGFβ inhibits H19 expression via SOX2 in<br />

tumor-initiating hepatocytes and this effect leads to inhibition of<br />

HCC development.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jinqiang Zhang, Chang Han,<br />

Weina Chen, Kyoungsub Song, Ying Wang, Lu Yao, Nathan Ungerleider, Hyunjoo<br />

Kwon, Tong Wu


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 549A<br />

687<br />

Expression pattern of the stem/progenitor cell marker<br />

Neighbor of Punc E 11 in different mouse models of<br />

liver regeneration.<br />

Vera Hoffmann, Andrea Bowe, Harald M. Curth, Tobias Goeser,<br />

Dirk Nierhoff; University Hospital of Cologne, Department of Gastroenterology<br />

and Hepatology, Cologne, Germany<br />

Background: Regeneration after acute liver injury naturally<br />

arises from the compartment of parenchymal liver cells. However,<br />

if proliferation capacity of hepatocytes is inadequate,<br />

liver progenitor cells are activated and can give rise to mature<br />

hepatocytes and cholangiocytes. In this project, we have<br />

focused on the expression pattern of the novel stem/progenitor<br />

cell marker Neighbor of Punc E 11 (Nope) in liver regeneration<br />

after acute and chronic liver injury. Methods: C57Bl6 mice on<br />

normal chow, on a 3,5-diethoxycarbonyl-1,4-dihydro-collidin<br />

(DDC) or a choline-deficient, ethionine-supplemented (CDE)<br />

diet were investigated. DDC and CDE diet were optionally<br />

followed by a regenerative period on normal chow. For additional<br />

acute liver injury, subgroups of normal mice as well as<br />

mice on a DDC diet were subjected to a partial hepatectomy<br />

(PH) with or without retrorsine to block hepatocyte proliferation.<br />

After defined time periods between 3 to 6 weeks, mice were<br />

sacrificed and quantitative RT-PCR and immunohistochemical<br />

costainings of the liver tissue for Nope with CK19, A6, EpCAM<br />

(oval cell markers), E-cadherin or HNF4α (hepatocyte markers)<br />

were performed. Results: Partial hepatectomy with or without<br />

retrorsine did not induce any expression of Nope above background<br />

level in the adult liver. After DDC diet the expression<br />

level of Nope was increased in the RT-PCR. In the CDE model,<br />

the expression level of Nope even reached the same level as<br />

in the fetal liver. We were able to detect ductular proliferations<br />

with coexpression of CK19, A6 and EpCAM with Nope.<br />

In the DDC model, only A6 also stained positive in a minor<br />

cell fraction within periductular Nope positive hepatocytic cell<br />

populations. After PH in the DDC model, followed by normal<br />

chow, the Nope expression level decreased, but small hepatocytic<br />

cells in proximity to ductular structures stained positiv for<br />

Nope. In the CDE model, we were able to detect clearly confined<br />

Nope positive hepatocytic cell clusters mainly in proximity<br />

to Nope positive ductular structures. Clearly distinguishable<br />

from these Nope expressing cell populations, expression of<br />

Nope was infrequently induced in HNF4α positive parenchymal<br />

hepatocytes. Discussion: The oncofetal marker Nope is<br />

expressed in CK19, A6 and EpCAM positive ductular cells as<br />

well as in CK19 negative periductular small hepatocytic cells<br />

presumably representing early hepatocytes after liver injury.<br />

These Nope positive early hepatocytes can form confined<br />

regenerative clusters with a minor fraction staining positive for<br />

A6. In conclusion, Nope is a marker for progenitor cells and<br />

periductular early hepatocytes in different mouse models of<br />

liver regeneration.<br />

Disclosures:<br />

Tobias Goeser - Advisory Committees or Review Panels: Gilead, Johnson, BMS,<br />

BMS; Grant/Research Support: Roche; Speaking and Teaching: Essex, BMS,<br />

Gilead, Novartis, Falk, Roche, Essex, BMS, Gilead, Novartis, Falk<br />

Dirk Nierhoff - Consulting: AbbVie; Speaking and Teaching: Gilead, AbbVie,<br />

BMS, Janssen, MSD<br />

The following authors have nothing to disclose: Vera Hoffmann, Andrea Bowe,<br />

Harald M. Curth<br />

688<br />

Ex vivo-expansion of human CD34 + cells from patients<br />

with liver cirrhosis enhances therapeutic efficacy of cell<br />

transplantation for rat cirrhotic liver<br />

Toru Nakamura 1,2 , Hironori Koga 1,2 , Hideki Iwamoto 1,2 , Yu<br />

Ikezono 1,2 , Fumitaka Wada 1,2 , Mitsuhiko Abe 1,2 , Takahiko<br />

Sakaue 1,2 , Takato Ueno 3,2 , Takuji Torimura 1 ; 1 Division of Gastroenterology,<br />

Department of Medicine, Kurume University School of<br />

Medicine, Kurume, Japan; 2 Liver Cancer Division, Research Center<br />

for Innovative Cancer Therapy, Kurume University, Kurume, Japan;<br />

3 Asakura Medical Association Hospital, Asakura, Japan<br />

Background: We demonstrated that the transplantation of<br />

human CD34 + cells into an immunodeficient rat liver fibrosis<br />

model reduced liver fibrosis by suppressing activated hepatic<br />

stellate cells and increasing MMPs activity, and led to hepatic<br />

regeneration. Recently, we reported that autologous granulocyte-colony<br />

stimulating factor (G-CSF)-mobilized CD34 + cell<br />

transplantation for patients with decompensated liver cirrhosis<br />

(LC) had therapeutic potential, but the colony-forming ability<br />

of CD34 + cells from patients with decompensated LC was<br />

reduced. Thus, recovery of CD34 + cell function is indispensable<br />

for cell transplantation therapy of patients with LC. The aim of<br />

this study was to investigate the efficacy of cell transplantation<br />

therapy with ex vivo-expanded human CD34 + cells for carbon<br />

tetrachloride (CCl 4<br />

)-induced liver fibrosis model. Methods:<br />

Human G-CSF-mobilized CD34 + cells of patients with LC were<br />

isolated by magnetic cell sorting system. Recipient nude rats<br />

were injected intraperitoneally with CCl 4<br />

twice weekly for 3<br />

weeks before initial treatment. Then, saline, 5×10 4 , or 1×10 6<br />

non-expanded and expanded CD34 + cells/kg body weight<br />

were transplanted via spleen, respectively. The administration<br />

of CCl 4<br />

was continued for three more weeks until the rats were<br />

sacrificed. Examination items were as follows: 1) FACS and<br />

Real-Time PCR analysis of non-expanded and expanded CD34 +<br />

cells, 2) morphometry of fibrotic areas by Azan-Mallory staining,<br />

3) immunohistochemistry using anti-CD31, smooth muscle<br />

myosin heavy chain-1 (SM1), αSMA, Ki67, and PCNA antibodies,<br />

and 4) the RT 2 Profiler TM PCR Array analysis. Results:<br />

For 7 days, CD34 + cells were effectively expanded. Increased<br />

expression of VE-cadherin, KDR and Tie-2 was determined by<br />

FACS analysis. The expression of pro-angiogenic growth factors<br />

in expanded CD34 + cells increased compared with non-expanded<br />

CD34 + cells. The transplanted cells differentiated into<br />

CD31 + and SM1 + cells. Expanded CD34 + cell transplantation<br />

had dose-dependently reduced liver fibrosis. Assessments<br />

of hepatocytes and sinusoidal endothelial cells proliferative<br />

activity indicated the superior potency of expanded CD34 +<br />

cells over non-expanded CD34 + cells. The PCR Array analysis<br />

against the adhesion molecules was showed that the expression<br />

of integrin αvβ3 was the most up-regulated gene compared<br />

with before culture. Three weeks of treatment with Cilengitide,<br />

which specially inhibits integrin αvβ3 signaling, resulted in the<br />

inhibition of CD34 + cells migration and worsened liver fibrosis<br />

in a dose-dependent manner. Conclusion: These findings suggest<br />

that expanded CD34 + cell transplantation promote better<br />

therapeutic effects for liver cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Toru Nakamura, Hironori Koga,<br />

Hideki Iwamoto, Yu Ikezono, Fumitaka Wada, Mitsuhiko Abe, Takahiko Sakaue,<br />

Takato Ueno, Takuji Torimura


550A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

689<br />

Attenuation of liver fibrosis development by angiotensin<br />

2-receptor blocker 1- treatment augments hepatocyte<br />

differentiation of hepatic progenitor cells<br />

Mitsuteru Kitade, Norihisa Nishimura, Hitoshi Yoshiji; Nara Medical<br />

University, Kashihara, Japan<br />

Recent research has elucidated mechanisms and potentials of<br />

hepatic progenitor cells (HPC) on regeneration of diseased<br />

liver. It is commonly known that hepatic stellate cells (HSC)<br />

play important roles as niche component cells of HPC, and<br />

that HPC expands along with HSC activation and liver fibrosis<br />

development. However, less is known about whether liver fibrosis<br />

development affects HPC-mediated regeneration. We have<br />

reported that angiotensin 2 strongly augments proliferation of<br />

activated-HSC, and angiotensin 2-receptor blocker 1 (ARB)<br />

exerts anti-fibrotic effect on diseased liver by inhibition of HSC<br />

proliferation. Our current study was performed to elucidate<br />

interactions between HSC and HPC-mediated liver regeneration<br />

using losartan, an ARB reagent on a DDC-induced mouse<br />

liver model. DDC-treatment augmented liver injury with fibrosis<br />

development and HPC expansion. ARB-treatment attenuated<br />

both a-SMA-positive activated HSC and laminin-positive extracellular<br />

matrix with increased hepatocyte (HC) differentiation<br />

of HPC, which may result in gain of liver/body weight ratio.<br />

Our in vitro experiments showed that ARB treatment did not<br />

alter capacity for HPC differentiation. In contrast, conditional<br />

medium of human HSC line (LX-2) augmented biliary epithelial<br />

cell (BEC) differentiation of HPC but attenuated efficiency for<br />

hepatocyte differentiation. Conditional medium of LX-2 incubated<br />

with angiotensin 2 enhanced this polarity of HPC differentiation<br />

towards BEC, which was blocked by ARB. These<br />

results indicated that HSC augments HPC differentiation fates<br />

towards BEC in paracrine manner. We also confirmed that<br />

both LX-2 and primary activated HSC strongly express Jagged1,<br />

a ligand for NOTCH, which plays central role in HPC<br />

differentiation fate towards BEC, which may controls HPC differentiation<br />

fate towards BEC. In conclusion, anti-fibrosis treatment<br />

may also be beneficial on efficient HPC-mediated liver<br />

regeneration via preferential differentiation of HPC towards HC<br />

by blocking NOTCH pathway in HPC niche.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mitsuteru Kitade, Norihisa<br />

Nishimura, Hitoshi Yoshiji<br />

690<br />

Bone Marrow Mesenchymal Stem Cells (BMSCS) From<br />

Decompensated Cirrhotics Show Premature Senescence<br />

And Limited Tissue Repair And Regenerative Potential<br />

Dhananjay Kumar 1 , Sheetalnath B. Rooge 1 , Smriti Shubham 1 ,<br />

Lovkesh Anand 2 , Sujata Mohanty 3 , Chhagan Bihari 4 , Anupam<br />

Kumar 1 , Shiv K. Sarin 2 ; 1 Department of Research, Institute of Liver<br />

and Biliary Sciences (ILBS), New Delhi, India; 2 Department of<br />

Hepatology, Institute of Liver and Biliary Sciences, New Delhi,<br />

India; 3 Stem Cell Facility, All India Institute of Medical Sciences,<br />

New Delhi, India; 4 Department of Pathology, Institute of Liver and<br />

Biliary Sciences, New Delhi, India<br />

Background & aim: Use of autologous BMSCs in chronic liver<br />

disease have shown varied clinical response ranging from negligible<br />

to mild improvement in CTP scores. Whether BMSCs<br />

from CLD patients are functionally efficient for repair & regeneration<br />

of cirrhotic liver is largely unknown. We investigated<br />

the functional potential of BMSCs of decompensated liver disease<br />

(DLD) patients. Patients & Methods: MSCs were isolated<br />

from DLD(N=10) patients & matched healthy BM donors(N=8)<br />

& characterized by surface marker expression & in-vitro differentiation<br />

to adipocytes & osteocytes. Levels of trophic &<br />

paracrine factors (RTPCR & cytokine array), angiogenic properties<br />

& immunomodulatory functions were studied. Population<br />

doubling time, CFU & SA-βGal staining were used to study<br />

cellular aging & senescence. Results: All MSCs fulfilled the minimal<br />

criteria for mesenchymal stromal cells; plastic adherence,<br />

spindle-shaped morphology, inducible osteo & adepogenesis<br />

& specific surface expression patterns. In RT-PCR analysis,<br />

DLD-MSCs(dMSCs) showed more than 2 fold decrease in<br />

expression of SDF1, HGF, FGF2, ANG1, JAG1 in comparison<br />

to healthy-MSCs(hMSCs). dMSCs also showed reduced capacity<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 551A<br />

bution of GFP was inversely correlated with alpha-fetoprotein<br />

(AFP) expression, suggesting local insulin signaling mediated<br />

by IRS2 may play a role in tumor heterogeneity and cellular<br />

differentiation. Exogenous expression of IRS2 promoted tumorigenesis<br />

in HepG2 soft agar and clonogenic assays and also<br />

corresponded with decreased AFP immunostaining. In HepaRG<br />

pIRS2-GFP positive cells were observed in discrete sites within<br />

the cultures surrounding “islands” of hepatocyte differentiation.<br />

IRS2 enriched cells had progenitor-like properties and we show<br />

that insulin/IRS2 signaling at a cellular level determines the<br />

differentiation, proliferative expansion and survival of hepatocyte-like<br />

tumor cells within the cultures, whilst expression of<br />

IRS2 is dynamically regulated during the timecourse of hepatocyte<br />

differentiation such that levels diminished as cells matured.<br />

Taken together, our results underscore the heterogeneity of<br />

insulin sensitivity at the cellular level within “homogeneous”<br />

cancer cell lines. We show that IRS2 expression is spatially<br />

patterned both in monolayers in vitro and in tumors in vivo and<br />

serves as a proxy for insulin sensitivity. Further investigation of<br />

how this novel cellular niche reacts to aberrant insulin signaling<br />

in the context of metabolic disease may provide future insights<br />

into the links that underpin the association between type II diabetes<br />

and HCC risk.<br />

Disclosures:<br />

The following authors have nothing to disclose: Fátima Manzano Núñez, Carlos<br />

Acosta Umanzor, Aránzazu Leal Tassias, Deborah J. Burks, Luke A. Noon<br />

692<br />

Sonic Hedgehog-Containing Exosomes Mediate Biliary<br />

Differentiation and Fibronectin Deposition via Epigenetic<br />

Regulation<br />

Nidhi Jalan-Sakrikar, Thiago de Assuncao, Jie Lu, Gwen Lomberk,<br />

Martin E. Fernandez-Zapico, Raul A. Urrutia, Robert C. Huebert;<br />

Gastroenterology and Hepatology, Mayo Clinic, Rochester, MN<br />

Background: Developmental morphogens play an important<br />

role in coordinating the ductular reaction and portal fibrosis<br />

that occur in the setting of cholangiopathies. However, little is<br />

known about how injured cholangiocytes signal to adjacent<br />

progenitor cells to promote repair after biliary injury. Recent<br />

<strong>studies</strong> show that sonic hedgehog (Shh) signaling is activated<br />

during liver regeneration and repair. In addition, cholangiocytes<br />

release exosomes that contain biologically active Shh and<br />

the secretion of these organelles increases during biliary fibrosis.<br />

Our lab has recently reported a differentiation protocol that<br />

generates induced pluripotent stem cell (iPSC)-derived cholangiocytes<br />

(iDC), a model useful for studying the stem cell to cholangiocyte<br />

transition. The purpose of this study was to dissect<br />

the mechanisms whereby injured cholangiocytes communicate<br />

with the neighboring progenitor cells via Shh-containing exosomes<br />

to initiate both cholangiocyte differentiation and extracellular<br />

matrix (ECM) remodeling. Methods and Results: Next<br />

generation RNA sequencing at each phase of the iPSC to iDC<br />

transition revealed significant upregulation of the Smoothened<br />

(SMO) receptor as well as Gli-1 and Gli-2 transcription factors<br />

during biliary specification and these changes correlated<br />

with a 60-fold increase in fibronectin (FN) levels. Exosomes<br />

isolated from LPS-injured cholangiocytes enhanced progenitor<br />

cell acquisition of cholangiocyte markers (CK7: 2.7±0.3-fold;<br />

CK19: 2.3±0.2-fold) as well as the expression and release<br />

of FN (8.3±2.4 fold). Cholangiocyte differentiation and FN<br />

release were both mediated by Shh as shown using pharmacologic<br />

(cyclopamine) or genetic (SMO shRNA) inhibition of<br />

Shh signaling. Concurrent alterations in epigenetic regulators<br />

(SetD7, KDM5D, EZH2) suggested that a common epigenetic<br />

regulatory program, driven by Shh, may mediate both cholangiocyte<br />

differentiation and ECM deposition. Cholangiocyte<br />

differentiation was associated with temporal alterations in histone<br />

methylation patterns that were Shh-responsive, including<br />

early increases in H3K4me3 (4.6±1.6-fold) and late decreases<br />

in H3K27me3 (-3.8±1.2-fold). In vivo, mice fed the choline-deficient,<br />

ethanolamine supplemented diet had an expansion of<br />

LGR5+ progenitor cells that was associated with Shh activation,<br />

increased FN deposition, and peri-portal fibrosis. Conclusions:<br />

We conclude that Shh-containing exosomes play an<br />

integral role in cholangiocyte differentiation from progenitor<br />

cells following biliary injury and that an epigenetically-driven<br />

cascade of gene regulation may simultaneously promote FN<br />

deposition and biliary fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nidhi Jalan-Sakrikar, Thiago de<br />

Assuncao, Jie Lu, Gwen Lomberk, Martin E. Fernandez-Zapico, Raul A. Urrutia,<br />

Robert C. Huebert<br />

693<br />

Interleukin-17 mediates liver progenitor cell transformation<br />

into cancer stem cells in hepatocellular carcinoma.<br />

Imène Gasmi 1,2 , Adrien Guillot 2,3 , Nabila Hamdaoui 1,2 , Arthur<br />

Brouillet 1,2 , Julien Calderaro 1,2 , Benoit Rousseau 1,2 , Jean-Michel<br />

Pawlotsky 1,2 , Fouad Lafdil 1,2 ; 1 INSERM U955 Henri Mondor Hospital,<br />

INSERM / University of Paris Est, Créteil, France; 2 INSERM<br />

/ University of Paris Est, Creteil, France; 3 NIH NIAAA, Bethesda,<br />

MD<br />

Introduction. Hepatocellular carcinoma (HCC) is the third leading<br />

cause of cancer-related death. HCC arises in the setting<br />

of cirrhotic livers in 80 to 90% of cases and progresses in an<br />

inflammatory context. Cancer stem cell biology has recently<br />

sparked the interest of scientists because of their capacity to<br />

initiate and to enhance tumor progression. However the underlying<br />

mechanisms by which expansion of CSCs is initiated are<br />

still unclear. After severe and chronic liver injury, liver progenitor<br />

cell compartment is activated under sustained inflammatory<br />

response. LPCs participate to the regenerative process and are<br />

also observed in HCC. Interestingly, among the large spectrum<br />

of released cytokines in HCC, recent <strong>studies</strong> identified<br />

IL-17-producing cells as a factor associated with a poor prognosis<br />

of the disease. In this study, we propose to determine<br />

whether IL-17 could be involved in tumorigenesis, in particular,<br />

by promoting the transformation of resident liver progenitor<br />

cells (LPCs) into CSCs. Materiels and methods. Serial sections<br />

from 70 HCC-patients were immuno-stained to identify LPCs<br />

(with anti-CK19) and IL-17-producing cells (with anti-IL-17). In<br />

vitro, a murine LPC line (BMOL) was used for long-term culture<br />

with or without IL-17 for 10, 20, 30 or 40 days. The expression<br />

of cancer stem cell markers were analyzed by quantitative<br />

RT-PCR and flow cytometry. Cell cycle controlling factors<br />

were assessed by western blot. Acquired self-renewal property<br />

of LPCs after long term culture with IL-17 was assessed by<br />

spheroid formation capacity analysis. Results. Identification<br />

and semi-quantitative analysis of CK19+ and IL-17+ cells in<br />

HCC patients showed a positive correlation between the number<br />

of infiltrated IL-17+ cells and the number of LPCs. In vitro,<br />

LPC stimulation with IL-17 led to increased expression of CSC<br />

marker mRNA expression including Klf4, ALDH1A1 or CD133,<br />

EpCAM and Glypican-3. CD133 protein and ALDH1A1 activity<br />

analyzed by flow cytometry, were enhanced in LPC cultured<br />

for 10, 20 or 30 days when compared to none treated LPCs.<br />

Cell cycle analysis showed that IL-17 induces the expression<br />

of the proto-oncogene c-Raf, and cyclin D1. In addition, LPCs<br />

cultured for 10 days in presence of IL-17 were able thereafter<br />

to form spheroids when transferred in low-attachment culture


552A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

dishes, in contrast to none pre-treated LPCs. Conclusion. Taken<br />

together, these results strongly suggest that IL-17 could contribute<br />

to HCC development, especially by leading LPC to acquire<br />

cancer stem cell-like properties. Therefore, IL-17 neutralization<br />

or IL-17 signaling pathway disruption in LPCs may contribute to<br />

CSC niche eradication in HCC treatment.<br />

Disclosures:<br />

Jean-Michel Pawlotsky - Consulting: Abbvie, Achillion, Bristol-Myers Squibb, Gilead,<br />

Janssen, Merck; Speaking and Teaching: Bristol-Myers Squibb, Gilead,<br />

Merck, Janssen<br />

The following authors have nothing to disclose: Imène Gasmi, Adrien Guillot,<br />

Nabila Hamdaoui, Arthur Brouillet, Julien Calderaro, Benoit Rousseau, Fouad<br />

Lafdil<br />

694<br />

Growth Factor Mobilized CD34+ Bone Marrow Stem<br />

Cells (BMSCs) Rescue the Loss of Regenerative Microenvironment<br />

in Decompensated liver Cirrhosis<br />

Sheetalnath B. Rooge 1 , Lovkesh Anand 2 , Dhananjay Kumar 1 ,<br />

Smriti Shubham 1 , Rakhi Maiwall 2 , Chhagan Bihari 3 , Anupam<br />

Kumar 1 , Shiv K. Sarin 2 ; 1 Department of Research, Institute of<br />

Liver and Biliary Sciences (ILBS), New Delhi, India; 2 Department<br />

of Hepatology, Institute of Liver and Biliary Sciences (ILBS), New<br />

Delhi, India; 3 Department of Pathology, Institute of Liver and Biliary<br />

Sciences (ILBS), New Delhi, India<br />

Background: Therapeutic effect of growth factor mobilized<br />

CD34+ BMSCs in management of chronic Liver Disease (CLD)<br />

has been shown by several investigators including our group<br />

(Gastroenterology 2012, 2015). However the underlying<br />

mechanisms operating in the hepatic tissue are not clearly<br />

understood. Aim: To study the cellular & molecular mechanisms<br />

of growth factor activities on liver regeneration in CLD<br />

patients. Patients & Method: Fourty-one Patients with cirrhosis<br />

were administered growth factors [G-CSF at 5mcg/kg at days<br />

1,2,3,4,5 & then every 3rd day till day 60, Erythropoietin<br />

at 500 I.U/Kg s/c twice a week for 2 months & tranjugular<br />

liver biopsy & hepatic vein samples were obtained before<br />

& after the therapy. Regenerative response was studied by<br />

immunohistochemistry (IHC) using CK19 for hepatic progenitor<br />

cells, Ki67 for hepatocyte replication. Associated cellular<br />

microenvironment was analyzed by IHC by using cell type<br />

specific markers i.e. myofibroblasts (α-SMA), BMSCs (CD34),<br />

tissue macrophage (CD68) & M2 macrophages (CD163).<br />

Further, to study the change in secretary microenvironment a<br />

panel of cytokines, chemokines & growth factors were analyzed<br />

in hepatic vein plasma using Millipore Milliplex Map<br />

Kit & compared between responders (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 553A<br />

Fig 1<br />

results demonstrated that TSG-6 promoted the autophagical<br />

function in hepatocytes, contributing to the liver regeneration.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sihyung Wang, Jeongeun Hyun,<br />

Jieun Kim, Youngmi Jung<br />

Disclosures:<br />

The following authors have nothing to disclose: Ritu Khosla, Archana Rastogi,<br />

Shyam Singh, Madavan Vasudevan, Viniyendra Pamecha, Gayatri Ramakrishna,<br />

Shiv K. Sarin, Nirupma Trehanpati<br />

696<br />

TSG-6 enhances autophagy in chronically damaged<br />

liver<br />

Sihyung Wang, Jeongeun Hyun, Jieun Kim, Youngmi Jung; Pusan<br />

National University, Pusan, Korea (the Republic of)<br />

Tumor necrosis factor-inducible gene 6 protein (TSG-6), one<br />

of cytokines secreted from mesenchymal stem/stromal cells, is<br />

known to act as an anti-inflammatory factor and reduce several<br />

pathological conditions. Recently, TSG-6 has been shown to<br />

protect liver from injury; however, the mechanism underlying the<br />

effect is poorly understood. Autophagy is the catabolic process<br />

that targets cell constituents to the lysosomes for degradation<br />

and known to be dysregulated in several diseases, including<br />

liver diseases. Emerging evidence suggests that the autopagic<br />

functions protect hepatocytes from damages. Hence, we<br />

hypothesize that TSG-6 promotes the autophagical clearance<br />

systems in hepatocytes and protects liver from chronic injury.<br />

To prove this hypothesis, mice were fed with Methionine choline-deficient<br />

ethionine (MCDE) for 4 weeks, followed by being<br />

injected intraperitonially with TSG-6 (50ng) (TSG-6 group) or<br />

saline (MCDE+vehicle group) with MCDE supplemented diets<br />

for additional 2 weeks. The histomorphological liver injury and<br />

increased level of liver enzymes were shown in MCDE-treated<br />

mice with or without saline, whereas those observations were<br />

markedly ameliorated in TSG-6-treated mice with MCDE diet<br />

(AST; control: 31.32±1.74, MCDE: 198.42±7.09, MCDE+vehicle:<br />

184.35±26.01, MCDE+TSG-6: 140.91±31.33/ ALT;<br />

control: 33.95±3.09, MCDE: 151.34±40, MCDE+vehicle:<br />

156.14±32.34, MCDE+TSG-6:111.91±10.39). RNA analysis<br />

showed the up-regulation of autophagy-relate genes, atg3 and<br />

atg7 (2.38±0.37 fold increase for atg3, 1.8±0.17 fold increase<br />

for atg7; p


554A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

698<br />

Decellularized Spleen Matrix for Regeneration of Functional<br />

Hepatic-like Tissue by Reseeding Bone Marrow<br />

Mesenchymal Stem Cells<br />

Junxi Xiang 1,2 , Xinglong Zheng 1,2 , Lifei Yang 2 , Rui Gao 2 , Yi Lv 1,2 ;<br />

1 Department of Hepatobiliary Surgery, First Affiliated Hospital<br />

of Xi’an Jiaotong University, Xi’an, China; 2 Research Institute of<br />

Advanced Surgical Technology and Engineering, Xi’an Jiaotong<br />

University, Xi’an, China<br />

Purpose This present study intends to regenerate hepatic-like<br />

tissue by reseeding and inducing differentiation of bone marrow<br />

mesenchymal stem cells (BMSCs) in decellularized spleen<br />

matrix (DSM), in order to expand the donor organ pool for liver<br />

transplantation. Methods The spleen, which provide access to<br />

a wider range of alternative sources, was chosen for preparing<br />

decellularized scaffold. We developed a physical-chemical<br />

method, namely, repetitive freezing/thawing cycles, deionized<br />

water, trypsin and Triton X-100 perfusion, to produce rats’<br />

DSM scaffold. Then, 2×10 7 BMSCs were repopulated into<br />

DSM for further dynamic culture and hepatic differentiation by<br />

using a dened two-step inducing protocol. Two-dimensional<br />

tissue culture asks were used as control. Results After decellularization,<br />

the spleen matrix presented a biomimetic three-dimensional<br />

porous architecture with intact vascular networks.<br />

The native extracellular matrix proteins including collagen I,<br />

collagen IV, bronectin and laminin were preserved, indicating<br />

a similar three-dimensional microenvironment mimicking the<br />

decellularized liver scaffold. When compared with two-dimensional<br />

culture, the produced DSM bio-scaffold promoted the cell<br />

engraftment and differentiation of BMSCs into hepatocyte-like<br />

cells using a dened protocol during a 21-day differentiation<br />

period, evidenced by morphological change, expression of<br />

hepatic-associated genes and proteins, glycogen storage, indocyanine<br />

green uptake, albumin secretion and urea production.<br />

Conclusions The physical-chemical strategy was suitable for<br />

producing decellularized spleen matrix, and the DSM scaffold<br />

might have considerable potential in cell-based therapy and<br />

fabricating functional hepatic-like tissue particularly because<br />

the DSM can support hepatic differentiation of BMSCs and<br />

effectively expand the donor pool.<br />

Disclosures:<br />

The following authors have nothing to disclose: Junxi Xiang, Xinglong Zheng, Lifei<br />

Yang, Rui Gao, Yi Lv<br />

699<br />

Wnt/beta-catenin signaling activates dUTP pyrophosphatase,<br />

a regulator of cellular dUTP levels, to prevent<br />

uracil misincorporation into DNA in human liver cancer<br />

stem cells<br />

Yoshiro Asahina, Taro Yamashita, Hajime Takatori, Naoki Oishi,<br />

Kouki Nio, Takehiro Hayashi, Yoshimoto Nomura, Tomoyuki<br />

Hayashi, Mariko Yoshida, Tomomi Hashiba, Tsuyoshi Suda,<br />

Masao Honda, Shuichi Kaneko; Disease Control and Homeostasis<br />

Internal Medicine, Kanazawa University Hospital, Kanazawa,<br />

Japan<br />

BackgroundRecent evidence suggests that hepatocellular carcinoma<br />

(HCC) is characterized by a hierarchical organization of<br />

a subset of cells termed cancer stem cells (CSCs), which may<br />

play a central role in resistance against cytotoxic reagents.<br />

However, it is still unclear, mechanistically, how CSCs prevent<br />

DNA damage induced by the reagents. In this study, we evaluated<br />

the role of dUTP pyrophosphatase (dUTPase), known to<br />

catalyze the hydrolysis of dUTP to dUMP, in the maintenance of<br />

the dUMP pool and prevention of uracil misincorporation into<br />

DNA in HCC CSCs. MethodsHuh7 and Huh1 cells were cultured<br />

routinely in DMEM supplemented with 10% fetal bovine<br />

serum. EpCAM-positive CSCs were purified by flow cytometry.<br />

The expression levels of EpCAM and dUTPase were evaluated<br />

by immunohistochemistry in 107 primary HCC samples that<br />

were surgically resected. The promoter activity of DUT (encoding<br />

dUTPase) was evaluated by a luciferase assay using Huh7<br />

cells transfected with pGL3-DUT-full and pRL-SV40 plasmids.<br />

The GSK3-beta inhibitor BIO, control MeBIO, and small interfering<br />

RNAs (siRNAs) targeting scramble or CTNNB1 sequences<br />

were used to regulate Wnt signaling in HCC cells. Results-<br />

Sorted EpCAM-positive CSCs showed more nuclear accumulation<br />

of dUTPase than EpCAM-negative Huh1 and Huh7 cells.<br />

Immunohistochemical analysis indicated a strong positive correlation<br />

between EpCAM and dUTPase expression in primary<br />

HCCs. High-dUTPase HCCs showed lower overall survival than<br />

that of low-dUTPase HCCs, and this difference was statistically<br />

significant (P = 0.0012). BIO treatment increased, whereas<br />

si-CTNNB1 treatment decreased the expression of EpCAM and<br />

dUTPase compared with that of the control Huh7 cells. We<br />

found a TCF4 binding element in the DUT promoter region,<br />

and BIO treatment enhanced, whereas si-CTNNB1 treatment<br />

attenuated DUT luciferase promoter activity compared with that<br />

of the control cells. In contrast, although 5-fluorouracil treatment<br />

enriched EpCAM-positive CSCs with nuclear accumulation of<br />

dUTPase, it had no effect on DUT promoter activity, suggesting<br />

no direct effect of 5-fluorouracil on DUT promoter activity. ConclusionTaken<br />

together, our data suggest that Wnt/beta catenin<br />

signaling activates dUTPase and may prevent DNA damage in<br />

liver CSCs. dUTPase may be a good target to eradicate liver<br />

CSCs resistant to cytotoxic reagents.<br />

Disclosures:<br />

Mariko Yoshida - Grant/Research Support: Bayer<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Yoshiro Asahina, Taro Yamashita,<br />

Hajime Takatori, Naoki Oishi, Kouki Nio, Takehiro Hayashi, Yoshimoto Nomura,<br />

Tomoyuki Hayashi, Tomomi Hashiba, Tsuyoshi Suda, Masao Honda<br />

700<br />

Predictors of Response to Grazoprevir/Elbasvir Among<br />

HCV Genotype 1 (GT1)–Infected Patients: Integrated<br />

Analysis of Phase 2-3 Trials<br />

Stefan Zeuzem 1 , Jürgen K. Rockstroh 2 , Paul Y. Kwo 3 , David Roth 4 ,<br />

Eric Lawitz 5 , Mark S. Sulkowski 6 , Xavier Forns 7 , Janice Wahl 8 ,<br />

Michael Robertson 8 , Bach-Yen T. Nguyen 8 , Eliav Barr 8 , Anita Y.<br />

Howe 8 , Michael D. Miller 8 , Peggy Hwang 8 , Erluo Chen 8 , Kenneth<br />

J. Koury 8 ; 1 Department of Internal Medicine I, J.W. Goethe University<br />

Hospital, Frankfurt, Germany; 2 University of Bonn, Bonn,<br />

Germany; 3 Indiana University School of Medicine, Indianapolis,<br />

IN; 4 University of Miami Miller School of Medicine, Miami, FL;<br />

5 Texas Liver Institute, University of Texas Health Science Center,<br />

San Antonio, TX; 6 Johns Hopkins University School of Medicine,<br />

Baltimore, MD; 7 Hospital Clinic, Barcelona, Spain; 8 Merck & Co.,<br />

Inc., Kenilworth, NJ<br />

Purpose: In phase 2-3 trials, 95% of GT1-infected patients<br />

(±cirrhosis, HIV coinfection, prior treatment, end-stage renal<br />

disease) who received grazoprevir 100 mg + elbasvir 50 mg<br />

(GZR/EBR) ± ribavirin (RBV) achieved a sustained virologic<br />

response 12 weeks after end of therapy (SVR12); 3% experienced<br />

virologic failure. This analysis assessed potential predictors<br />

of response to therapy. Methods: Analyses were performed<br />

on 2 pooled datasets of 1408 GT1-infected patients enrolled


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 555A<br />

in phase 2/3 trials of GZR/EBR ± RBV: (1) treatment-naive<br />

patients (TN; N=801), and (2) patients who previously failed<br />

peginterferon + RBV ± first-generation protease inhibitor (TE;<br />

N=607). Demographic factors and presence of resistance-associated<br />

variants (RAVs) that were fit (ie, >25% of the overall<br />

baseline viral load) were considered. Univariate logistic regression<br />

models were fitted one variable at a time in assessing<br />

the potential association with SVR12. Multivariable logistic<br />

regression (MVLR) models with forward selection were then<br />

applied to identify significant (ie, P < 0.1) independent predictors<br />

of SVR12. Results: In the MVLR, among TN patients, age,<br />

sex, cirrhosis, HIV coinfection, baseline NS3 RAVs, and use<br />

of RBV did not impact SVR12 rates. Among TN GT1a-infected<br />

patients, baseline HCV RNA level and baseline NS5A RAVs<br />

conferring >5-fold shift in the potency of EBR in vitro (termed<br />

NS5A >5× RAVs) were identified as significant predictors<br />

of SVR12. Baseline HCV RNA had a significant impact only<br />

when NS5A >5× RAVs were present, representing 5.3% of<br />

the GT1a population. No significant predictors were identified<br />

among GT1b-infected patients. Among TE patients, no significant<br />

predictors were identified for GT1b-infected patients or<br />

in GT1a-infected patients with prior relapse. Among GT1a-infected<br />

patients with prior on-treatment failure, female patients<br />

and noncirrhotics had higher SVR12 rates, and the addition of<br />

RBV and/or longer treatment durations had a positive impact<br />

on SVR12. The most prevalent (>1%) NS5A >5× RAVs in both<br />

TN and TE patients were L31M and Y93H. Baseline NS5A >5×<br />

RAVs, which were detected in 8% of TE GT1a-infected patients<br />

had a significant negative impact on SVR12; this impact was<br />

confined to the 12-wk treatment duration, as no patient treated<br />

for 16 or 18 weeks with RBV experienced virologic failure.<br />

Conclusion: GZR/EBR ± RBV is highly effective among GT1-infected<br />

patients. Among GT1a-infected patients, presence of fit<br />

NS5A >5× RAVs at baseline impacts efficacy. These RAVs are<br />

uncommon (


556A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

John M. Vierling - Advisory Committees or Review Panels: Abbvie, Bristol-Meyers-Squibb,<br />

Gilead, Hyperion, Intercept, Janssen, Novartis, Merck, Sundise,<br />

HepQuant, Salix, Immuron, Exalenz, Chronic Liver Disease Foundation; Board<br />

Membership: Clinical Research Centers of America, LLC; Grant/Research Support:<br />

Abbvie, Bristol-Meyers-Squibb, Eisai, Gilead, Hyperion, Intercept, Janssen,<br />

Novartis, Merck, Sundise, Ocera, Mochida, Immuron, Exalenz, Conatus; Speaking<br />

and Teaching: GALA, Chronic Liver Disease Foundation, ViralEd, Chronic<br />

Liver Disease Foundation, Clinical Care Options<br />

Anita Y. Howe - Employment: Merck Research Laboratory<br />

Peggy Hwang - Employment: Merck, Merck<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Joan R. Butterton - Employment: Merck Sharp & Dohme Corp.; Stock Shareholder:<br />

Merck Sharp & Dohme Corp.<br />

Janice Wahl - Employment: Merck & Co,<br />

Eliav Barr - Employment: Merck<br />

Barbara A. Haber - Employment: Merck<br />

702<br />

Short-duration therapy with daclatasvir/asunaprevir/<br />

beclabuvir fixed-dose combination plus sofosbuvir in<br />

patients with chronic hepatitis C genotype 1 (FOURward<br />

Study)<br />

Mark S. Sulkowski 1 , Steven L. Flamm 2 , Zeid Kayali 3 , Eric Lawitz 4 ,<br />

Paul Y. Kwo 5 , Fiona McPhee 6 , Anne Torbeyns 7 , Eric A. Hughes 8 ,<br />

Eugene S. Swenson 6 , Philip Yin 6 , Misti Linaberry 8 ; 1 Johns Hopkins<br />

University, Baltimore, MD; 2 Northwestern University, Chicago, IL;<br />

3 Inland Empire Liver Foundation, Rialto, CA; 4 The Texas Liver Institute,<br />

University of Texas Health Center, San Antonio, TX; 5 Indiana<br />

University, Indianapolis, IN; 6 Bristol-Myers Squibb, Wallingford,<br />

CT; 7 Bristol-Myers Squibb, Braine l’Alleud, Belgium; 8 Bristol-Myers<br />

Squibb, Princeton, NJ<br />

Background: DCV-TRIO is a fixed-dose, twice-daily combination<br />

of daclatasvir 30mg (pangenotypic NS5A inhibitor),<br />

asunaprevir 200mg (NS3 inhibitor), and beclabuvir 75mg<br />

(non-nucleoside NS5B inhibitor). DCV-TRIO has achieved high<br />

rates of sustained virologic response at posttreatment Week 12<br />

(SVR12) after 12 weeks of treatment in cirrhotic and non-cirrhotic<br />

populations with HCV genotype (GT)1 infection. The<br />

FOURward study investigated the efficacy and safety of DCV-<br />

TRIO in combination with sofosbuvir (SOF; nucleotide NS5B<br />

inhibitor) for shortened treatment durations of 4 or 6 weeks.<br />

Methods: In this open-label, phase 2 study, non-cirrhotic treatment-naive<br />

patients aged ≥18 years with HCV GT1a or 1b<br />

infection (GT1b capped at 40%) were randomly assigned (1:1)<br />

to receive DCV-TRIO BID and SOF 400mg QD for 4 or 6 weeks.<br />

The primary endpoint was SVR12. Patients with virologic failure<br />

were offered 12 weeks of retreatment with DCV-TRIO+RBV<br />

(with optional PegIFNα) or SOF+RBV+PegIFNα-2a, dependent<br />

upon resistance testing by population-based sequencing at failure.<br />

Results: Of the 28 treated patients, 61% were female,<br />

89% white, 79% HCV GT1a, with a median age of 58.5 yrs.<br />

Median baseline viral load was high (approx. 9×10 6 IU/mL),<br />

61% were non-CC IL28B GT. All patients completed treatment<br />

and entered follow-up. HCV RNA was 2M IU/mL (7/21, 33%) vs


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 557A<br />

+ RBV (PR) ± first-generation PI. TE pts were further classified<br />

as those who had relapsed after completing prior treatment<br />

and those who experienced virologic failure (VF) on prior<br />

treatment (eg, null or partial responders, or virologic breakthrough).<br />

Results: With a 12-week regimen of GZR/EBR (no<br />

RBV), SVR was achieved in 94% of TN pts (5% VF), 96% of<br />

prior relapsers (0% VF) and 89% of prior on-treatment failures<br />

(10% VF). With a 16- or 18-week (+ RBV) regimen, 95% of<br />

prior on-treatment failures achieved SVR12 (0% VF). In each<br />

group, efficacy was similar in noncirrhotic and cirrhotic pts<br />

(Table 1). Baseline NS5A RAVs were uncommon (10%); however,<br />

efficacy was lower in pts with baseline NS5A RAVs conferring<br />

>5-fold shift in the potency of EBR in vitro, particularly<br />

in TN pts with high viral load and in prior on-treatment failures<br />

treated for 12 weeks. Conclusion: A 12-week regimen of GZR/<br />

EBR (no RBV) is highly effective among GT1a-infected TN pts<br />

and prior relapsers. A 16-week regimen of GZR/EBR (+ RBV) is<br />

highly effective among TE GT1a-infected pts who experienced<br />

virologic failure on prior treatment. Efficacy is similar in noncirrhotic<br />

and cirrhotic pts.<br />

Table 1<br />

a Lost to follow-up, withdrew consent, discontinued due to adverse<br />

event<br />

b Excludes nonvirologic failures<br />

Disclosures:<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Jürgen K. Rockstroh - Advisory Committees or Review Panels: Abbvie, BI, BMS,<br />

Merck, Roche, Tibotec, Abbvie, Bionor, Tobira, ViiV, Gilead, Janssen; Consulting:<br />

Novartis; Grant/Research Support: Merck; Speaking and Teaching: Abbott,<br />

BI, BMS, Merck, Roche, Tibotec, Gilead, Janssen, ViiV<br />

Paul Y. Kwo - Advisory Committees or Review Panels: Abbvie, Abbott, Novartis,<br />

Merck, Gilead, BMS, Janssen; Consulting: BMS; Grant/Research Support:<br />

Roche, Abbvie, Merck, BMS, Abbott, Idenix, Vital Therapeutics, Gilead, Vertex,<br />

Merck, Idenix; Speaking and Teaching: Merck, Merck<br />

David Roth - Advisory Committees or Review Panels: Merck, Sharp and Dome;<br />

Consulting: Merck, Sharp and Dome<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Xavier Forns - Consulting: Jansen, Abbvie; Grant/Research Support: Jansen,<br />

Gilead<br />

Janice Wahl - Employment: Merck & Co,<br />

Bach-Yen T. Nguyen - Employment: Merck<br />

Eliav Barr - Employment: Merck<br />

Anita Y. Howe - Employment: Merck Research Laboratory<br />

Michael D. Miller - Employment: Merck & Co., Inc.; Stock Shareholder: Merck<br />

& Co., Inc.<br />

Peggy Hwang - Employment: Merck, Merck<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

704<br />

Prevention of allograft HCV reinfection with peri-transplant<br />

MBL-HCV1 monoclonal antibody in combination<br />

with oral direct-acting antiviral<br />

Parvez S. Mantry 1 , Heidi L. Smith 2 , Raymond T. Chung 3 , William<br />

C. Chapman 4 , Michael P. Curry 5 , Thomas D. Schiano 6 , Yang<br />

Wang 2 , Deborah C. Molrine 2 ; 1 The Liver Institute, Methodist Dallas<br />

Medical Center, Dallas, TX; 2 MassBiologics of the University<br />

of Massachusetts Medical School, Boston, MA; 3 Department of<br />

Gastroenterology, Massachusetts General Hospital, Boston, MA;<br />

4 Department of Surgery, Washington University, St. Louis, MA;<br />

5 Division of Gastroenterology, Beth Israel Deaconess Medical Center,<br />

Boston, MA; 6 Recanati-Miller Transplantation Institute, Mount<br />

Sinai Medical Center, New York, NY<br />

Background Patients with hepatitis C viremia at the time of<br />

liver transplantation experience rapid infection of the donor<br />

allograft, increased risk of graft failure, and accelerated rates<br />

of fibrosis. To evaluate a peri-transplant treatment strategy to<br />

prevent allograft hepatitis C virus (HCV) infection, MBL-HCV1,<br />

a neutralizing human monoclonal antibody (mAb) targeting a<br />

conserved linear epitope of the HCV envelope, was combined<br />

with licensed direct-acting antivirals (DAA) in an open-label<br />

exploratory efficacy trial. Methods All subjects had HCV RNA<br />

titers > 10 4 IU/mL at the time of transplantation. MBL-HCV1<br />

was dosed intravenously at 50 mg/kg, beginning just prior to<br />

the anhepatic phase of the transplantation procedure. Subjects<br />

received three mAb infusions on the day of transplant and<br />

daily infusions on days 1-7 post-transplantation. Oral DAA therapy<br />

was initiated between days 3 and 7 with documentation<br />

of adequate graft function and administered in combination<br />

with weekly mAb infusions through day 28. During the second<br />

post-transplant month, MBL-HCV1 was administered bi-weekly<br />

while continuing oral DAA therapy; study treatment could be<br />

extended until day 84 (12 weeks) post-transplant in subjects<br />

whose HCV RNA remained undetectable. In Part 1 of the study,<br />

eight subjects received MBL-HCV1 in combination with telaprevir<br />

for up to 12 weeks. In Part 2 of the study, two subjects<br />

have received MBL-HCV1 in combination with sofosbuvir for<br />

12 weeks. Results In Part 1 of the study, prolonged aviremia<br />

was observed in 5 of 8 subjects and one subject (12.5%) had<br />

a sustained virologic response (SVR24). There were 11 serious<br />

adverse events assessed as unrelated to study treatment. In Part<br />

2 of the study, both subjects (100%) achieved SVR12 and the<br />

first subject has completed the 24 week visit with an SVR. One<br />

subject had a history of prior virologic breakthrough on interferon<br />

therapy, while the other previously experienced virologic<br />

relapse following a 12 week pre-transplant course of sofosbuvir/simeprevir.<br />

With MBL-HCV1 plus sofosbuvir treatment,<br />

both subjects were aviremic by day 21 post-transplantation.<br />

Two serious adverse events were reported > 2 months after<br />

completion of study treatment and assessed as unrelated to<br />

study treatment. Conclusions Peri-transplant therapy with MBL-<br />

HCV1 in combination with an oral DAA is capable of preventing<br />

post-transplant HCV recurrence and the combination<br />

of MBL-HCV1 with sofosbuvir appears promising. Given the<br />

observed viral kinetics, this treatment approach has the potential<br />

to achieve viral eradication in the immediate post-transplant<br />

period, thereby mitigating the risks of severe allograft HCV<br />

infection.


558A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Parvez S. Mantry - Consulting: Salix, Gilead, Janssen, Abbvie, BMS; Grant/<br />

Research Support: Salix, Merck, Gilead, Boehringer-Ingelheim, Mass Biologics,<br />

Vital Therapies, Santaris, mass biologics, Bristol-Myers Squibb, Abbive, Bayer-Onyx,<br />

Shinogi, Tacere, Intercept; Speaking and Teaching: Gilead, Janssen,<br />

Salix<br />

Heidi L. Smith - Employment: MassBiologics of UMass Medical School<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

Michael P. Curry - Advisory Committees or Review Panels: Bristol Meyers Squib,<br />

Abbvie; Grant/Research Support: Gilead Sciences, Mass Biologics, Merck,<br />

Salix, Conatus; Stock Shareholder: Achilion<br />

Deborah C. Molrine - Employment: MassBiologics of Univ of Massachusetts<br />

Medical School<br />

The following authors have nothing to disclose: William C. Chapman, Thomas<br />

D. Schiano, Yang Wang<br />

705<br />

Analysis of HCV Genotype 1 Variants Detected During<br />

Monotherapy and Combination Therapy with Next Generation<br />

HCV Direct-Acting Antiviral Agents ABT-493 and<br />

ABT-530<br />

Teresa Ng, Tami Pilot-Matias, Rakesh Tripathi, Gretja Schnell,<br />

Thomas Reisch, Jill Beyer, Tanya Dekhtyar, Armen Asatryan, Federico<br />

J. Mensa, Andrew L. Campbell, Jens Kort, Christine Collins;<br />

AbbVie, North Chicago, IL<br />

Background: ABT-530 (NS5A inhibitor) and ABT-493 (NS3/4A<br />

protease inhibitor identified by AbbVie and Enanta) are next<br />

generation HCV direct-acting antiviral agents (DAAs). Each<br />

has demonstrated potent anti-HCV activity in genotype (GT)<br />

1-infected patients in 3-day monotherapy (Study M13-595) as<br />

well as in combination therapy for 12 weeks (Phase 2 study<br />

M14-867). In this report we present the characterization of HCV<br />

GT1 variants detected in samples from these 2 <strong>studies</strong>. Methods:<br />

Population sequencing was performed on the NS3/4A<br />

and NS5A genes from all available baseline (BL) samples from<br />

both <strong>studies</strong>, last available samples with HCV RNA ≥ 1000 IU/<br />

mL during monotherapy in study M13-595, and first available<br />

sample after virologic failure with HCV RNA ≥1000 IU/mL in<br />

the combination study M14-867. Sequences were examined<br />

for the presence of resistance-associated variants (RAVs) at<br />

positions where variants have been shown to confer resistance<br />

to HCV protease or NS5A inhibitors. NS3 or NS5A RAVs were<br />

introduced into appropriate subgenomic HCV GT1 replicons<br />

and their susceptibility to ABT-493 or ABT-530 was evaluated<br />

in a transient transfection assay. Results: In ABT-493 monotherapy<br />

study (n=49), none of the BL samples had NS3 RAVs. A<br />

single NS3 RAV (GT1a A156T) emerged in 1 subject during<br />

therapy. In ABT-530 monotherapy study (n=40), NS5A RAVs<br />

(M28V, Q30R, L31M, H/P58 variants, and Y93 variants)<br />

were present in 8 subjects at BL. Most of these were present as<br />

NS5A single RAVs that do not confer resistance to ABT-530.<br />

During monotherapy, additional RAVs emerged in 3 of these<br />

8 subjects, producing double RAVs that conferred a higher<br />

level of resistance (27- to >300-fold) to ABT-530 than either<br />

RAV alone. In the M14-867 combination study of ABT-493 +<br />

ABT-530 (n=79), 2 subjects had NS3 RAVs (R155K) and 13<br />

subjects had NS5A RAVs (M28V, Q30 variants, L31M, H/P58<br />

variants, and Y93H) at BL. To date, 78 subjects have achieved<br />

SVR 4<br />

(99% SVR 4<br />

) and 1 subject relapsed at post treatment<br />

week 4. In the relapse subject, no BL RAVs were detected,<br />

but a double NS5A RAV (Q30K+H58D) emerged at failure.<br />

Conclusions: Variants resistant to ABT-493 were not detected<br />

in subjects at BL, and a single NS3 RAV (A156T) emerged<br />

in 1 subject during ABT-493 monotherapy. During ABT-530<br />

monotherapy, additional NS5A variants emerged in 3 of the 8<br />

subjects who already had variants present at BL. These double<br />

RAVs conferred medium to high resistance to ABT-530. In the<br />

combination study, a treatment-emergent double NS5A RAV<br />

was detected at failure in the subject who relapsed. All other<br />

subjects from this study have achieved SVR 4<br />

, regardless of the<br />

presence of BL NS3 and/or NS5A variants.<br />

Disclosures:<br />

Teresa Ng - Employment: AbbVie; Patent Held/Filed: AbbVie; Stock Shareholder:<br />

AbbVie<br />

Tami Pilot-Matias - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Rakesh Tripathi - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

Gretja Schnell - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

Thomas Reisch - Employment: Abbvie; Stock Shareholder: Abbvie<br />

Jill Beyer - Employment: Abbvie; Stock Shareholder: Abbvie<br />

Tanya Dekhtyar - Employment: Abbvie; Stock Shareholder: Abbvie<br />

Armen Asatryan - Employment: AbbVie<br />

Federico J. Mensa - Employment: Abbvie Inc.; Stock Shareholder: Abbvie Inc.<br />

Andrew L. Campbell - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

Christine Collins - Employment: AbbVie, Inc.<br />

706<br />

Improvement in liver disease parameters following<br />

treatment with daclatasvir + sofosbuvir and ribavirin<br />

in patients with chronic HCV infection and advanced<br />

cirrhosis<br />

Robert J. Fontana 1 , Fred Poordad 2 , Eugene R. Schiff 3 , John M.<br />

Vierling 4 , Charles S. Landis 5 , Rafia Bhore 6 , Philip Yin 7 , Stephanie<br />

Noviello 6 , Eugene S. Swenson 7 ; 1 University of Michigan Medical<br />

Center, Ann Arbor, MI; 2 University of Texas Health Science Center,<br />

Texas Liver Institute, San Antonio, TX; 3 University of Miami<br />

Milller School of Medicine, Schiff Center for Liver Diseases, Miami,<br />

FL; 4 Baylor College of Medicine, Houston, TX; 5 University of Washington<br />

School of Medicine, Seattle, WA; 6 Bristol-Myers Squibb,<br />

Princeton, NJ; 7 Bristol-Myers Squibb, Wallingford, CT<br />

Background: The pangenotypic combination of daclatasvir<br />

(DCV) and sofosbuvir (SOF), with or without ribavirin (RBV) has<br />

achieved SVR12 rates of 82-98% in multiple high-need patient<br />

populations with chronic HCV infection. In the phase 3 ALLY-1<br />

study, DCV+SOF+RBV achieved SVR12 in 83% of patients<br />

with advanced cirrhosis and in 94% of those with post-liver<br />

transplant recurrence. We evaluated changes over time in liver<br />

disease parameters among patients from the advanced cirrhosis<br />

cohort of ALLY-1. Methods: This open-label study enrolled<br />

treatment-naive or -experienced adults with HCV infection of<br />

any genotype (GT). Patients in the advanced cirrhosis cohort<br />

(N=60) received 12 weeks of treatment with DCV 60 mg<br />

+ SOF 400 mg once-daily and RBV (initially 600 mg/day,<br />

adjusted for hemoglobin and creatinine clearance). The primary<br />

endpoint was HCV RNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 559A<br />

clusions: DCV+SOF+RBV treatment in patients with advanced<br />

cirrhosis achieved high SVR12 rates and improved clinical and<br />

biochemical indicators of liver disease.<br />

Disclosures:<br />

Robert J. Fontana - Consulting: GlaxoSmithKline, CLDF; Grant/Research Support:<br />

Gilead, vertex, BMS, Jansen, Gilead<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Eugene R. Schiff - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead, Merck, Janssen, Salix Pharmaceutical, Pfizer, Arrowhead; Consulting:<br />

Acorda; Grant/Research Support: Bristol Myers Squibb, Abbott / AbbVee, Gilead,<br />

Merck, Conatus, Medmira, Roche, Janssen, Orasure Technologies, Discovery<br />

Life Sciences, Siemens, Beckman Coulter, Siemens<br />

John M. Vierling - Advisory Committees or Review Panels: Abbvie, Bristol-Meyers-Squibb,<br />

Gilead, Hyperion, Intercept, Janssen, Novartis, Merck, Sundise,<br />

HepQuant, Salix, Immuron, Exalenz, Chronic Liver Disease Foundation; Board<br />

Membership: Clinical Research Centers of America, LLC; Grant/Research Support:<br />

Abbvie, Bristol-Meyers-Squibb, Eisai, Gilead, Hyperion, Intercept, Janssen,<br />

Novartis, Merck, Sundise, Ocera, Mochida, Immuron, Exalenz, Conatus; Speaking<br />

and Teaching: GALA, Chronic Liver Disease Foundation, ViralEd, Chronic<br />

Liver Disease Foundation, Clinical Care Options<br />

Charles S. Landis - Grant/Research Support: Gilead, Abbvie, BMS<br />

Philip Yin - Stock Shareholder: Bristol-Myers Squibb<br />

Stephanie Noviello - Consulting: Merck/Schering-Plough; Employment: Bristol-Myers<br />

Squibb, Merck/Schering-Plough; Stock Shareholder: Merck/Schering-Plough,<br />

J&J<br />

Eugene S. Swenson - Employment: Bristol-Myers Squibb<br />

The following authors have nothing to disclose: Rafia Bhore<br />

up to follow-up week 24 (FU24) will be reported at the presentation.<br />

Methods: The Phase 2 portion is a randomized, double-blind<br />

trial (unblinding occurred when all patients achieved<br />

FU4) to evaluate the tolerability, efficacy and pharmacokinetics<br />

(PK) of GZR/EBR. Non-cirrhotic treatment-naive or -experienced<br />

GT1 CHC patients were randomized to receive either 50 or<br />

100 mg of GZR co-administered with 50 mg EBR; all study<br />

medications were given QD for 12 weeks. Treatment response<br />

was assessed using COBAS TaqMan v2.0 (lower limit of quantitation<br />

< 1.2 Log IU/mL). Intensive PK sampling over a 24-hour<br />

dosing period for non-compartmental analyses was performed<br />

in a subset of patients at treatment week 4 (TW4). Results:<br />

Sixty-three patients were enrolled (42% male; mean age 61;<br />

100% GT1b); one subject never took study medication due<br />

to an abnormal ECG. Mean baseline HCV-RNA was 6.2 Log<br />

IU/mL. The most common adverse events (AEs) were headache,<br />

nausea and fatigue. There was no dose relationship for<br />

AEs. No patient discontinued due to an AE or study medication<br />

intolerance. All patients who received study medication<br />

(n=62) suppressed HCV-RNA to TND (Target Not Detected)<br />

during treatment. No patient experienced virologic breakthrough/rebound<br />

during treatment and all patients achieved<br />

SVR4; however, 1 patient in the GZR 100 mg arm relapsed<br />

at FU12. PK parameters (e.g., C 2hr<br />

, C 24hr<br />

and AUC 0-24<br />

) were<br />

calculated and compared with non-Japanese data from another<br />

GZR/EBR study that administered 100 mg GZR + 50 mg EBR.<br />

Conclusions: GZR/EBR therapy was well-tolerated in Japanese<br />

patients with GT1 CHC infection, across GZR doses. In<br />

addition, SVR12 was achieved in 61/62 patients (98.4%).<br />

Exposures of GZR and EBR were slightly higher in Japanese<br />

non-cirrhotic patients compared to non-Japanese non-cirrhotic<br />

patients. GZR 100 mg + EBR 50 mg QD was selected as the<br />

dose to move forward into the Phase 3 portion of the study<br />

based on comparable efficacy and tolerability.<br />

Patients with undetectable HCV-RNA<br />

707<br />

Efficacy, Safety And Pharmacokinetics Of Grazoprevir<br />

(MK-5172) And Elbasvir (MK-8742) In Hepatitis C Genotype<br />

1 Infected Non-Cirrhotic Japanese Patients (Phase<br />

2 Portion In Phase 2/3 Combined Study)<br />

Norifumi Kawada 1 , Fumitaka Suzuki 2 , Yoshiyasu Karino 3 , Kazuaki<br />

Chayama 4 , Yoshito Itoh 5 , Takeshi Okanoue 6 , Makoto Nakamuta 7 ,<br />

Naoyoshi Yatsuzuka 8 , Go Fujimoto 8 , Lisa Chiacchierini Lupinacci 9 ,<br />

Michael Robertson 9 , Luzelena Caro 9 , Daniel A. Tatosian 9 , Anita Y.<br />

Howe 9 , Hiromitsu Kumada 2 ; 1 Department of Hepatology, Osaka<br />

City University Medical School, Osaka, Japan; 2 Department of<br />

Hepatology, Toranomon Hospital, Tokyo, Japan; 3 Department of<br />

Gastroenterology, Sapporo Kosei General Hospital, Hokkaido,<br />

Japan; 4 Department of Gastroenterology and Metabolism, Hiroshima<br />

University, Hiroshima, Japan; 5 Department of Molecular<br />

Gastroenterology and Hepatology, Kyoto Prefectural University<br />

of Medicine, Kyoto, Japan; 6 Department of Gastroenterology and<br />

Hepatology, Saiseikai Suita Hospital, Osaka, Japan; 7 Department<br />

of Gastroenterology, Kyushu Medical Center, National Hospital<br />

Organization, Fukuoka, Japan; 8 MSD K.K., Tokyo, Japan; 9 Merck<br />

& Co., Inc., Kenilworth, NJ<br />

Background: Grazoprevir (MK-5172, GZR; NS3/4A protease<br />

inhibitor) and elbasvir (MK-8742, EBR; NS5A inhibitor) are<br />

being developed as components of an all-oral direct-acting<br />

antiviral regimen for the treatment of chronic hepatitis C (CHC).<br />

The aim of the Phase 2 portion of this study was to compare<br />

regimens containing 50 or 100 mg GZR + 50 mg EBR with the<br />

objective of assessing the tolerability and efficacy in genotype<br />

1 (GT1)-infected non-cirrhotic CHC Japanese patients. Results<br />

Disclosures:<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

Fumitaka Suzuki - Speaking and Teaching: BMS<br />

Yoshiyasu Karino - Speaking and Teaching: BMS KK<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

Naoyoshi Yatsuzuka - Employment: MSD K.K.<br />

Go Fujimoto - Employment: MSD K.K.; Stock Shareholder: Merck<br />

Lisa Chiacchierini Lupinacci - Employment: Merck<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Luzelena Caro - Employment: Merck & Co., Inc.<br />

Daniel A. Tatosian - Employment: Merck & Co; Stock Shareholder: Merck & Co<br />

Anita Y. Howe - Employment: Merck Research Laboratory<br />

Hiromitsu Kumada - Patent Held/Filed: SRL; Speaking and Teaching: Bristol-Myers<br />

Squibb,Pharma International, MSD, Janssen Pharma., Glaxosmithkline<br />

The following authors have nothing to disclose: Takeshi Okanoue, Makoto<br />

Nakamuta


560A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

708<br />

Efficacy and Safety of Co-Formulated Ombitasvir/Paritaprevir/Ritonavir<br />

with Ribavirin in Adults with Chronic<br />

HCV Genotype 4 Infection in Egypt (AGATE-II)<br />

Gamal E. Esmat 2 , Wahid H. Doss 3 , Roula B. Qaqish 1 , Imam<br />

Waked 4 , Gamal E. Shiha 5 , Ayman Yosry 3 , Mohamad Hassany 3 ,<br />

Jennifer King 1 , Coleen Hall 1 , Rebecca Redman 1 , Niloufar Mobashery<br />

1 ; 1 AbbVie, Inc, North Chicago, IL; 2 Cairo University, Cairo,<br />

Egypt; 3 National Hepatology and Tropical Medicine Research<br />

Institute, Cairo, Egypt; 4 National Liver Institute, Menoufiya, Egypt;<br />

5 Egyptian Liver Research Institute and Hospital, Dakahliah, Egypt<br />

Purpose: Chronic hepatitis C virus (HCV) infection is the main<br />

cause of liver cirrhosis and liver cancer in Egypt and one of the<br />

five leading causes of death. The prevalence of HCV infection<br />

in Egypt is the highest in the world (10-14%) with over 90%<br />

infected with HCV genotype (GT) 4. In the Phase 2b PEARL-I<br />

study, the efficacy and safety of the direct acting antiviral<br />

agents (DAA) ombitasvir (OBV), a NS5A inhibitor, and paritaprevir,<br />

a NS3/4A protease inhibitor identified by AbbVie and<br />

Enanta, co-dosed with ritonavir (PTV/r) with or without ribavirin<br />

(RBV) were assessed in 135 patients with HCV GT4 infection<br />

without cirrhosis. SVR12 was 100% in both treatment naïve<br />

(TN) and prior interferon (IFN) and RBV treatment experienced<br />

(TE) patients receiving OBV plus PTV/r with RBV for 12 weeks.<br />

AGATE II is the first phase 3 trial to evaluate OBV/PTV/r with<br />

RBV in Egypt for GT4 infected subjects with and without compensated<br />

cirrhosis. Methods: This ongoing Phase 3, multicenter,<br />

open label trial enrolled 160 subjects across 5 sites in Egypt.<br />

Non-cirrhotic patients (n=100) received co-formulated OBV/<br />

PTV/r once-daily (25 mg/150 mg/100 mg) with weight based<br />

RBV for 12 weeks (Arm A). Cirrhotic subjects (n=60) were randomized<br />

1:1 to the same regimen for either 12 or 24 weeks<br />

(Arms B and C; n=30/arm). The primary efficacy endpoint is<br />

SVR12. Safety is being evaluated by adverse event (AE) monitoring,<br />

laboratory testing, and other standard assessments.<br />

Subjects will be followed for 48 weeks post treatment. Results:<br />

A total of 160 noncirrhotic (Arm A, n= 101) and cirrhotic<br />

(Arm B, n=30, C n=29) subjects were enrolled. Approximately<br />

half were TE (61% prior nulls, 24% prior relapsers and 15%<br />

partial responders). Overall, 76% are male. The average age<br />

is 54 years and mean BMI 29.5 kg/m 2 , with 18% reporting a<br />

history of diabetes and 67% with HOMA-IR scores >3. Overall<br />

DAA-related treatment emergent adverse events (AEs) occurring<br />

in >10% of subjects were fatigue (12%) and headache (15%).<br />

There was 1 subject with a serious AE (SAE) of deep venous<br />

thrombosis deemed reasonabiy possibly related to study drug.<br />

There were no AEs leading to discontinuation of study drug,<br />

1 subject withdrew consent (Arm A) and two subjects met the<br />

on-treatment criteria for virologic failure (Arm B, n=1 and C,<br />

n=1). Conclusion: The study regimen was generally well tolerated.<br />

SVR12 for non-cirrhotic and cirrhotic subjects in the 12<br />

week Arms A and B, respectively, will be presented as well<br />

as the current number of on-treatment virologic failures and<br />

post-treatment relapses for cirrhotic subjects in arm C.<br />

Disclosures:<br />

Gamal E. Esmat - Advisory Committees or Review Panels: MSD &BMS companies,<br />

MSD &BMS companies, Abbvie; Grant/Research Support: Gilead Sc;<br />

Speaking and Teaching: Roche & GSK companies, Roche & GSK companies<br />

Roula B. Qaqish - Employment: AbbVie<br />

Imam Waked - Advisory Committees or Review Panels: Janssen; Speaking and<br />

Teaching: Hoffman L Roche, Merck, BMS, Gilead, AbbVie<br />

Mohamad Hassany - Grant/Research Support: Gilead Sc<br />

Jennifer King - Employment: AbbVie<br />

Coleen Hall - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Rebecca Redman - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Niloufar Mobashery - Employment: Abbvie; Stock Shareholder: abbvie<br />

The following authors have nothing to disclose: Wahid H. Doss, Gamal E. Shiha,<br />

Ayman Yosry<br />

709<br />

Baseline HCV NS5A resistance-associated variants do<br />

not impact SVR12 rates in non-cirrhotic and post-liver<br />

transplant patients with genotype 1 infection treated<br />

with daclatasvir and sofosbuvir with or without ribavirin<br />

for 12 weeks: An integrated analysis<br />

Fiona McPhee 1 , Dennis Hernandez 1 , Nannan Zhou 1 , Joseph<br />

Ueland 1 , Vincent Vellucci 1 , Sandra Hartman-Neumann 1 , Xiaoyan<br />

Yang 1 , Zhou Han 1 , Fei Yu 1 , Rong Yang 2 , Stephanie Noviello 2 ;<br />

1 Bristol-Myers Squibb, Wallingford, NJ; 2 Bristol-Myers Squibb,<br />

Princeton, NJ<br />

Background: High sustained virologic response 12 weeks posttreatment<br />

(SVR12) has been achieved in patients infected with<br />

HCV genotype (GT)1 treated with daclatasvir (DCV) and sofosbuvir<br />

(SOF)±ribavirin (RBV) for 12 weeks. We evaluated the<br />

impact of baseline NS5A variants on SVR12 rates by pooling<br />

individual patient data from phase 2 and 3 <strong>studies</strong>. Methods:<br />

Data were derived from GT1, non-cirrhotic treatment-naive<br />

or -experienced patients with HCV infection and post-liver<br />

transplant recurrence (ALLY-1 [N=41]), HIV/HCV coinfection<br />

(ALLY-2 [N=105]), and chronic HCV infection (AI444-040<br />

[N=82]). Plasma samples from 228 (182 GT1a, 46 GT1b)<br />

patients receiving DCV 60mg (except for HIV/HCV coinfected<br />

patients on boosted protease inhibitors (30mg), or efavirenz or<br />

neviripine (90mg)) once daily (QD) and SOF 400mg QD±RBV<br />

were collected at baseline. HCV RNA was isolated and the<br />

NS5A region amplified and sequenced. The impact of DCV-resistant<br />

variants (at NS5A amino acid positions 28, 30, 31,<br />

or 93 for GT1a and at 31 or 93 for GT1b) on SVR12 rates<br />

were assessed in all patients with baseline sequence (n=220);<br />

8 patients (6 GT1a, 2 GT1b) were excluded due to poor compliance<br />

(N=1), incarceration (N=1), loss to follow-up (N=1) or<br />

no available baseline NS5A sequence (N=5). Results: Of 220<br />

GT1 patients with baseline NS5A sequence, 99.1% (218/220<br />

[39/41 posttransplant and 179/179 non-cirrhotic patients])<br />

achieved SVR12. Baseline NS5A variants at amino acid positions<br />

associated with DCV resistance were detected in 11.8%<br />

(26/220 [GT1a: 10.8% (19/176); GT1b: 15.9% (7/44)]) of<br />

patients. SVR12 was achieved in 100.0% (26/26) of patients<br />

with these NS5A variants compared with 99.0% (192/194)<br />

without. Observed baseline NS5A variants included M28T/V,<br />

Q30H/L/R, L31M, and Y93C/H/S in GT1a and L31M and<br />

Y93H in GT1b. One posttransplant patient had NS5A variants<br />

M28T+Q30R at baseline; a loss in DCV activity of 44000-<br />

fold was observed in-vitro with M28T+Q30R vs. a GT1a reference<br />

sequence without these substitutions. Two posttransplant<br />

patients (1 GT1a, 1 GT1b) without baseline NS5A variants<br />

relapsed; NS5A-Q30R+L31V emerged in the GT1a patient<br />

and a deletion at NS5A-P32 was observed in the GT1b<br />

patient. These substitutions conferred >333,333-fold DCV<br />

resistance in-vitro. The 5 patients without available baseline<br />

NS5A sequence achieved SVR12. Conclusions: In this analysis<br />

of patients with HCV GT1 infection (non-cirrhotic patients<br />

with monoinfection or HIV/HCV coinfection and posttransplant<br />

HCV recurrence), 11.8% had baseline DCV-resistant variants.<br />

SVR12 was achieved in 100% of GT1 patients with these baseline<br />

DCV-resistant variants even when high-level DCV-resistant<br />

variants were detected.<br />

Disclosures:<br />

Fiona McPhee - Employment: Bristol-Myers Squibb<br />

Dennis Hernandez - Employment: Bristol-Myers Squibb<br />

Nannan Zhou - Employment: Bristol Myers Squibb Company


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 561A<br />

Joseph Ueland - Employment: Bristol-Myers Squibb<br />

Fei Yu - Employment: Bristol-Myers Squibb<br />

Rong Yang - Employment: BMS; Stock Shareholder: BMS<br />

Stephanie Noviello - Consulting: Merck/Schering-Plough; Employment: Bristol-Myers<br />

Squibb, Merck/Schering-Plough; Stock Shareholder: Merck/Schering-Plough,<br />

J&J<br />

The following authors have nothing to disclose: Vincent Vellucci, Sandra Hartman-Neumann,<br />

Xiaoyan Yang, Zhou Han<br />

710<br />

Drug-drug interactions between next generation direct<br />

acting antivirals ABT-493 and ABT-530 with cyclosporine<br />

or tacrolimus in healthy subjects<br />

Matthew P. Kosloski, Sandeep Dutta, Weihan Zhao, Jingtao Wu,<br />

David Pugatch, Armen Asatryan, Jens Kort, Wei Liu; AbbVie,<br />

North Chicago, IL<br />

Purpose: A next generation direct acting antiviral (DAA) combination<br />

of ABT-493 (NS3/4A protease inhibitor discovered<br />

by AbbVie and Enanta) + ABT-530 (NS5A inhibitor) is being<br />

developed for the treatment of chronic hepatitis C (HCV) genotype<br />

1-6 infection. ABT-493 + ABT-530 combination demonstrated<br />

high sustained virologic response rate in Phase 2<br />

<strong>studies</strong>. Two Phase 1, open-label <strong>studies</strong> were conducted to<br />

assess the pharmacokinetics, safety and tolerability when ABT-<br />

493 + ABT-530 is coadministered with immunosuppressants<br />

cyclosporine or tacrolimus. Methods: Healthy adults subjects<br />

received single doses of cyclosporine 100 mg (n=12) or tacrolimus<br />

1 mg (n=12) alone or in combination with ABT-493<br />

300 mg QD + ABT-530 120 mg QD. Intensive blood sampling<br />

for determination of cyclosporine, tacrolimus, ABT-493<br />

and ABT-530 concentrations was performed and pharmacokinetic<br />

parameters (maximum observed concentration [C max<br />

],<br />

area under the concentration-time curve [AUC t<br />

or AUC inf<br />

]<br />

and trough concentration [C 24<br />

]) were estimated. Safety and<br />

tolerability were assessed throughout the study. Results: Cyclosporine<br />

C max<br />

, AUC t<br />

, and AUC inf<br />

in blood were minimally<br />

affected (≤14% change) when coadministered with steady-state<br />

ABT-493 + ABT-530. Steady-state C max<br />

, AUC 24<br />

, and C 24<br />

in<br />

plasma were slightly increased for ABT-493 (30%, 37%, and<br />

34%, respectively) and for ABT-530 (11%, 22%, and 26%,<br />

respectively) when coadministered with cyclosporine. Tacrolimus<br />

C max<br />

, AUC t<br />

, and AUC inf<br />

in blood were slightly increased<br />

(50%, 53%, and 45%, respectively) when coadministered with<br />

steady-state ABT-493 + ABT-530. Steady-state C max<br />

, AUC 24<br />

,<br />

and C 24<br />

in plasma were minimally affected for ABT-493 (≤11%<br />

change) and for ABT-530 (≤2% change) when coadministered<br />

with tacrolimus. No serious adverse events were observed<br />

in either study. There were no patterns to the adverse events<br />

reported, and no new safety issues were identified. Conclusions:<br />

No dose adjustment is required for ABT-493, ABT-530,<br />

and cyclosporine when coadministered. No dose adjustment is<br />

required for ABT-493 and ABT-530 when coadministered with<br />

tacrolimus. It is recommended that subjects receiving tacrolimus<br />

should continue to use their current dose when initiating treatment<br />

with DAAs, and reduce the dose of tacrolimus if necessary<br />

based on therapeutic monitoring.<br />

Disclosures:<br />

Sandeep Dutta - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Weihan Zhao - Employment: AbbVie, Inc; Stock Shareholder: AbbVie, Inc, Exxon<br />

Mobile, American Airlines<br />

Jingtao Wu - Employment: Abbvie<br />

Armen Asatryan - Employment: AbbVie<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

Wei Liu - Employment: AbbVie<br />

The following authors have nothing to disclose: Matthew P. Kosloski, David<br />

Pugatch<br />

711<br />

Comparative Efficacy and Tolerability of Daclatasvir<br />

+ Sofosbuvir versus Sofosbuvir + Ribavirin in Patients<br />

with Chronic Hepatitis C Coinfected with HIV: A Matching-Adjusted<br />

Indirect Comparison<br />

Elyse Swallow 1 , Jinlin Song 1 , Yong Yuan 2 , Anupama Kalsekar 2 ,<br />

Caroline Kelley 1 , Miranda Peeples 1 , Fan Mu 1 , Stephanie Noviello<br />

2 , James Signorovitch 1 ; 1 Health Economics and Outcomes<br />

Research, Analysis Group, Inc., Boston, MA; 2 Bristol-Myers<br />

Squibb, Princeton, NJ<br />

Background and aims: To compare the efficacy and tolerability<br />

of daclatasvir and sofosbuvir (DCV+SOF) versus sofosbuvir<br />

and ribavirin (SOF+R) in patients coinfected with human immunodeficiency<br />

virus (HIV) and hepatitis C virus (HCV). Methods:<br />

A systematic literature review was conducted to identify Phase<br />

3 clinical trials suitable for comparison with the trial of DCV+-<br />

SOF in HIV/HCV-coinfected patients (ALLY-2). Two eligible trials<br />

of SOF+R (PHOTON-1 and PHOTON-2) were identified.<br />

Individual patient data from ALLY-2 were available; published<br />

summary data were extracted and pooled for the PHOTON<br />

trials. When multiple durations of SOF+R were assessed for<br />

the same genotype, only the arm with the FDA/EMA approved<br />

duration was included in the present analysis. Patients enrolled<br />

in ALLY-2 were subject to the inclusion and exclusion criteria<br />

reported in the PHOTON trials. To adjust for cross-trial differences,<br />

ALLY-2 patients were weighted to match all available<br />

summary baseline characteristics reported in both the ALLY-2<br />

and PHOTON trials; specifically, age, BMI, gender, race, ethnicity,<br />

treatment naïve, viral genotypes, HCV RNA level, IL28B<br />

genotype, cirrhosis status, CD4 T-cell count and combination<br />

anti-retroviral therapy (cART) regimens were included in the<br />

analysis. Sustained virologic response at week 12 post-treatment<br />

(SVR12), discontinuation due to adverse events (AEs),<br />

and rates of specific AEs reported for both SOF+R trials were<br />

compared between the treatments. Results: 91 of 153 patients<br />

from ALLY-2 who were treated with 12 weeks of DCV+SOF met<br />

the inclusion criteria of the PHOTON trials and were included<br />

in the analysis. 42 of the 497 patients from the PHOTON trials<br />

who were treatment-naïve, genotype 3, and treated with 12<br />

rather than 24 weeks of SOF+R were excluded. The SVR12<br />

rate was significantly higher among patients treated with DCV<br />

+ SOF than those treated with SOF + R both before (96.7% vs.<br />

84.6%; p=0.002) and after (100.0% vs. 84.6%; p


562A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

712<br />

Safety and Tolerability of Grazoprevir/Elbasvir in<br />

Patients With Chronic Hepatitis C (HCV) Infection: Integrated<br />

Analysis of Phase 2-3 Trials<br />

Geoffrey M. Dusheiko 1 , Michael P. Manns 2 , John M. Vierling 3 , K.<br />

Rajender Reddy 4 , Mark S. Sulkowski 5 , Paul Y. Kwo 6 , Eric Lawitz 7 ,<br />

Deborah D. Brown 8 , Stephanie Klopfer 8 , Michael Robertson 8 , Janice<br />

Wahl 8 , Eliav Barr 8 , Edgar Charles 8 ; 1 Hepatology, Royal Free<br />

and University College, London, United Kingdom; 2 Hannover Medical<br />

School, Hannover, Germany; 3 Baylor College of Medicine,<br />

Houston, TX; 4 University of Pennsylvania, Philadelphia, PA; 5 Johns<br />

Hopkins University School of Medicine, Baltimore, MD; 6 ndiana<br />

University School of Medicine, Indianapolis, IN; 7 Texas Liver Institute,<br />

University of Texas Health Science Center, San Antonio, TX;<br />

8 Merck & Co., Inc., Kenilworth, NJ<br />

Purpose: In phase 2-3 <strong>studies</strong>, treatment with grazoprevir<br />

(GZR) 100 mg/elbasvir (EBR) 50 mg ± ribavirin (RBV) resulted<br />

in high rates of sustained virologic response (SVR) in HCV-infected<br />

patients, including those with compensated cirrhosis.<br />

The purpose of this analysis was to define the overall safety<br />

profile of GZR/EBR given for 8, 12, 16, or 18 weeks in these<br />

<strong>studies</strong>. Methods: Clinical adverse events and laboratory<br />

abnormalities reported on therapy, or within 14 days of endof-treatment<br />

were compared in 1795 patients (of whom 1033<br />

received GZR/EBR without RBV, 657 received GZR/EBR with<br />

RBV, and 105 received placebo). Results: A diverse population<br />

was enrolled: 61% were male, 13% Black/African-American,<br />

mean age 52.6 years (11% ≥65 years); 27% had compensated<br />

cirrhosis, and 18% were HIV/HCV coinfected. The overall<br />

safety profiles of GZR/EBR (without RBV) and placebo were<br />

comparable (Table 1). The addition of RBV was associated with<br />

more AEs. Among patients who received GZR/EBR, there were<br />

3 deaths (ventricular arrhythmia, strangulated hernia, motor<br />

vehicle accident); all 3 were unrelated to study medication. No<br />

patients who received placebo died. The frequency of adverse<br />

events was not associated with sex, age, presence of cirrhosis,<br />

or HIV/HCV coinfection. Late serum ALT elevations (defined<br />

as >5× ULN, in patients who had normal ALT values between<br />

treatment week [TW] 2-4) were noted in 0.8% of patients,<br />

generally at/after TW8. These were typically asymptomatic,<br />

resolving with continued therapy, scheduled end of therapy,<br />

or (in 3/1690, 0.18%) a protocol-mandated stop of therapy,<br />

and they were not associated with hyperbilirubinemia. Conclusion:<br />

GZR/EBR ± RBV was generally well tolerated in a large,<br />

diverse patient population with low SAE rate. Fewer discontinuations<br />

due to AEs and overall improved tolerability with lower<br />

reductions in hemoglobin and lower elevations in total bilirubin<br />

were demonstrated in the RBV-free cohort. ALT elevations<br />

occurring late in the course of therapy were infrequent and not<br />

clinically significant.<br />

Table 1<br />

Disclosures:<br />

Geoffrey M. Dusheiko - Advisory Committees or Review Panels: AbbVie, BMS,<br />

BMS, Merck; Board Membership: Gilead Sciences, Gilead Sciences<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

John M. Vierling - Advisory Committees or Review Panels: Abbvie, Bristol-Meyers-Squibb,<br />

Gilead, Hyperion, Intercept, Janssen, Novartis, Merck, Sundise,<br />

HepQuant, Salix, Immuron, Exalenz, Chronic Liver Disease Foundation; Board<br />

Membership: Clinical Research Centers of America, LLC; Grant/Research Support:<br />

Abbvie, Bristol-Meyers-Squibb, Eisai, Gilead, Hyperion, Intercept, Janssen,<br />

Novartis, Merck, Sundise, Ocera, Mochida, Immuron, Exalenz, Conatus; Speaking<br />

and Teaching: GALA, Chronic Liver Disease Foundation, ViralEd, Chronic<br />

Liver Disease Foundation, Clinical Care Options<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Paul Y. Kwo - Advisory Committees or Review Panels: Abbvie, Abbott, Novartis,<br />

Merck, Gilead, BMS, Janssen; Consulting: BMS; Grant/Research Support:<br />

Roche, Abbvie, Merck, BMS, Abbott, Idenix, Vital Therapeutics, Gilead, Vertex,<br />

Merck, Idenix; Speaking and Teaching: Merck, Merck<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Deborah D. Brown - Employment: Merck & Co., Inc.; Stock Shareholder: Merck<br />

& Co., Inc<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Janice Wahl - Employment: Merck & Co,<br />

Eliav Barr - Employment: Merck<br />

Edgar Charles - Employment: Merck & Co.<br />

The following authors have nothing to disclose: Stephanie Klopfer


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 563A<br />

713<br />

Resistance Analysis of Treatment-Naïve and DAA-Experienced<br />

Genotype 1 Patients with and without Cirrhosis<br />

Who Received Short-Duration Treatment with Sofosbuvir/GS-5816+<br />

GS-9857<br />

Edward J. Gane 2 , Evguenia S. Svarovskaia 1 , Robert H. Hyland 1 ,<br />

Luisa M. Stamm 1 , Anu Osinusi 1 , Diana M. Brainard 1 , Krishna<br />

Chodavarapu 1 , Michael D. Miller 1 , Hongmei Mo 1 , Christian<br />

Schwabe 3 ; 1 Gilead Sciences Inc, Foster City, CA; 2 New Zealand<br />

Liver Transplant Unit, Auckland City Hospital, Auckland, New Zealand;<br />

3 Auckland Clinical Studies Ltd, Auckland, New Zealand<br />

Introduction: Pretreatment resistance associated variants (RAVs)<br />

have been associated with relapse following short treatment<br />

durations with regimens containing three or more directly<br />

acting antivirals (DAAs). GS-9857, a potent pan-genotypic<br />

HCV NS3 protease inhibitor (PI) administered with SOF/<br />

GS-5816, a fixed dose combination of pan-genotypic NS5B<br />

and NS5A inhibitors, for 6 weeks was highly effective in treatment-naïve<br />

genotype (GT) 1 HCV patients without cirrhosis.<br />

Each compound in this three-drug combination demonstrated<br />

a high barrier to resistance, in vitro. The aim of this study<br />

was to characterize the effect of pretreatment RAVs on the<br />

treatment outcome and the viral resistance in patients who<br />

relapsed following treatment with SOF/GS-5816+GS-9857<br />

for 4 or 6 weeks. Methods: Treatment-naïve GT1 HCV-infected<br />

patients with or without cirrhosis were treated for 4 or<br />

6 weeks with SOF/GS-5816+GS-9857 (n=30). Treatment-experienced<br />

patients with or without cirrhosis who failed a<br />

prior multi-DAA regimen were treated for 6 weeks with SOF/<br />

GS-5816+GS-9857 (n=45). NS3, NS5A and NS5B were<br />

amplified and deep sequenced (1% detection assay cut-off).<br />

All patients were sequenced pretreatment and at the time<br />

of virologic failure (all relapse). Results: Pretreatment deep<br />

sequencing analysis of NS3, NS5A and NS5B was successful<br />

for all 75 patients enrolled. Pretreatment NS3 and/or NS5A<br />

RAVs were detected in 51% (23/45) of treatment naïve and<br />

46% (14/30) in DAA-experienced patients. NS5B RAVs were<br />

detected in 1/75 (1%) of patients. Among the different treatment<br />

groups, SVR12 rates in patients with and without pretreatment<br />

RAVs were similar (Table). In the 4 week treatment arm,<br />

76% (11/15) of patients relapsed; no RAVs emerged in any<br />

of these 11 patients. In combined 6 week treatment arms, 22%<br />

(13/60) of patients relapsed; of these 13 patients, one treatment<br />

naïve patient with cirrhosis had low levels of a NS3 RAV<br />

emerge (V55A, 1.9%). No other NS3, NS5A, or NS5B RAVs<br />

emerged. Conclusions: In contrast to DAA regimens with SOF<br />

and other NS5A and NS3 inhibitors, pretreatment RAVs did<br />

not affect response to short durations of treatment with SOF/<br />

GS-5816+GS-9857.Virologic failure was not associated with<br />

emergence of resistance. These clinical results suggest a high<br />

barrier to resistance with co-administration with SOF, GS-5816<br />

and GS-9857 as was seen in vitro with the individual agents.<br />

Disclosures:<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Evguenia S. Svarovskaia - Employment: Gilead Sciences Inc; Stock Shareholder:<br />

Gilead Sciences Inc<br />

Robert H. Hyland - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Luisa M. Stamm - Employment: Gilead Sciences<br />

Anu Osinusi - Employment: gilead sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Michael D. Miller - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

The following authors have nothing to disclose: Krishna Chodavarapu, Christian<br />

Schwabe<br />

714<br />

Efficacy and Safety of Ombitasvir/Paritaprevir/Ritonavir<br />

Co-Administered with Ribavirin in Adults with<br />

Genotype 4 Chronic Hepatitis C Infection and Cirrhosis<br />

(AGATE-I)<br />

Tarik Asselah 2 , Tarek I. Hassanein 3 , Roula B. Qaqish 1 , Jordan J.<br />

Feld 4 , Christophe Hezode 5 , Stefan Zeuzem 6 , Peter Ferenci 7 , Tami<br />

Pilot-Matias 1 , Yao Yu 1 , Rebecca Redman 1 , Niloufar Mobashery 1 ;<br />

1 AbbVie, Inc, North Chicago, IL; 2 Centre de Recherche sur l’Inflammation,<br />

Inserm UMR 1149, Université Paris Diderot, AP-HP<br />

Hôpital Beaujon, Clichy, France; 3 Southern California Liver Centers<br />

and Southern California Research Center, Coronado, CA;<br />

4 Toronto Centre for Liver Disease, University of Toronto, Toronto,<br />

ON, Canada; 5 Henri Mondor University Hospital, AP-HP, Université<br />

Paris-Est, Creteil, France; 6 J.W. Goethe University, Frankfurt,<br />

Germany; 7 Medical University Vienna, Vienna, Austria<br />

Purpose: HCV genotype 4 (GT4) represents approximately<br />

20% of global HCV infection. Although GT4 infection is<br />

more common in the Middle East and sub-Saharan Africa,<br />

with globalization, GT4 is now seen increasingly in Europe<br />

and many other countries. In the Phase 2b PEARL-I study, the<br />

efficacy and safety of the two direct acting antiviral agents<br />

(2DAA) ombitasvir (OBV), a NS5A inhibitor and paritaprevir,<br />

a NS3/4A protease inhibitor identified by AbbVie and<br />

Enanta, co-dosed with ritonavir (PTV/r) with or without ribavirin<br />

(RBV) were assessed in 135 subjects with HCV GT4 infection<br />

without cirrhosis. SVR12 was 100% in both treatment naïve<br />

(TN) and prior interferon (IFN) and RBV treatment experienced<br />

(TE) subjects receiving 2DAA+RBV for 12 weeks. This study<br />

extends those observations by evaluating OBV/PTV/r with<br />

RBV in HCV GT4-infected subjects with compensated cirrhosis.<br />

Methods: This ongoing Phase 3, randomized, open-label, multinational<br />

study (NCT 02265237) enrolled HCV GT4-infected<br />

TN subjects or IFN/RBV or pegIFN/RBV TE subjects with compensated<br />

cirrhosis. Subjects were randomized 1:1 to receive<br />

co-formulated OBV/PTV/r once daily (25 mg/150 mg/100<br />

mg) co-administered with weight based RBV for 12 (Arm A) or<br />

16 weeks (Arm B) with an approximately equal number of TN<br />

and TE subjects in each arm. A 24 week treatment arm (C) and<br />

an exploratory assessment in subjects who have experienced<br />

virologic failure with either sofosbuvir/pegIFN/RBV or sofosbuvir/RBV<br />

will follow. The primary objectives are to assess safety<br />

and SVR12 rates of these 2 DAA regimens as compared to a<br />

historical SVR12 rate for HCV GT4-infected subjects treated<br />

with pegIFN/RBV. Results: At the time of the abstract, 55 and<br />

56 cirrhotic subjects were randomized into Arms A and B,<br />

respectively. Of the 111 subjects, 48% were TN and 52%<br />

were TE with IFN/RBV or pegIFN/RBV (30% prior nulls, 12%<br />

prior relapsers and 10% partial responders). At baseline, 91%<br />

of subjects had a Child-Pugh score of 5, 6% of 6 and 3% of<br />

7. Overall, 72% are male, 78% White and 17% Black or<br />

African American. The mean age is 57 years and mean BMI


564A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

28 kg/m 2 , with 29% reporting a history of diabetes. Overall<br />

DAA-related treatment emergent adverse events (AEs) occurring<br />

in ≥10% of subjects were fatigue and headache (~15% each).<br />

5 subjects reported a total of 12 treatment-emergent serious<br />

AEs; 1 deemed related to study drug (manic crisis). No AE led<br />

to discontinuation of study drug; 1 subject withdrew consent<br />

and 1 subject met the on-treatment criteria for virologic failure.<br />

Conclusion: OBV/PTV/r with RBV for 12 and 16 weeks was<br />

generally well tolerated. SVR4 and SVR12 for subjects in Arms<br />

A and B will be presented.<br />

Disclosures:<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

Tarek I. Hassanein - Advisory Committees or Review Panels: AbbVie, Bristol-Myers<br />

Squibb; Grant/Research Support: AbbVie Pharmaceuticals, Obalon, Bristol-Myers<br />

Squibb, Eiasi Pharmaceuticals, Gilead Sciences, Janssen R&D, Idenix<br />

Pharmaceuticals, Ikaria Therapeutics, Merck Sharp & Dohme, NGM BioPharmaceuticals,<br />

Ocera Therapeutics, Salix Pharmaceuticals, Sundise, TaiGen Biotechnology,<br />

Takeda Pharmaceuticals, Vital Therapies, Tobria; Speaking and<br />

Teaching: Baxter, Bristol-Myers Squibb, Gilead, Salix, AbbVie<br />

Roula B. Qaqish - Employment: AbbVie<br />

Jordan J. Feld - Advisory Committees or Review Panels: Merck, Janssen, Gilead,<br />

AbbVie, Theravance, Bristol Meiers Squibb; Grant/Research Support: AbbVie,<br />

Boehringer Ingelheim, Janssen, Gilead, Merck<br />

Christophe Hezode - Speaking and Teaching: Roche, BMS, MSD, Janssen, abbvie,<br />

Gilead<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Peter Ferenci - Advisory Committees or Review Panels: Idenix, Gilead, MSD,<br />

Janssen, Salix, AbbVie, BMS; Patent Held/Filed: Madaus Rottapharm; Speaking<br />

and Teaching: Gilead, Roche<br />

Tami Pilot-Matias - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Yao Yu - Employment: Abbvie<br />

Rebecca Redman - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Niloufar Mobashery - Employment: Abbvie; Stock Shareholder: abbvie<br />

12 weeks after end of therapy (SVR12). Results: A total of 886<br />

TN pts (164 black, 722 nonblack) and 650 TE pts (88 black,<br />

562 nonblack) with GT1, 4, or 6 infection were included in the<br />

analysis. Demographics and efficacy are displayed in Table 1.<br />

A total of 144/154 (94%) black TN pts who received 12-week<br />

regimens without RBV achieved SVR12. Among black TE pts<br />

who received 12-week/no RBV, 12-week/RBV, 16–18-week/<br />

no RBV, or 16–18-week/RBV regimens, 31/33 (94%), 27/27<br />

(100%), 11/11 (100%), and 17/17 (100%), respectively,<br />

achieved SVR12. Conclusion: GZR/EBR was highly effective<br />

among black HCV-infected TN and TE pts with comparable<br />

responses observed in the nonblack population.<br />

Table 1<br />

715<br />

Efficacy of Grazoprevir (GZR) and Elbasvir (EBR) in<br />

Black HCV-Infected Patients: Results of a Pooled Analysis<br />

of Phase 2/3 Studies<br />

Philippe J. Zamor 1 , Jonathan McCone 2 , John M. Vierling 9 , Reem<br />

H. Ghalib 3 , Velimir A. Luketic 4 , Brian Pearlman 5 , Natarajan Ravendhran<br />

6 , Luis A. Balart 7 , Michael Robertson 8 , Peggy Hwang 8 ,<br />

Bach-Yen T. Nguyen 8 , Janice Wahl 8 , Eliav Barr 8 , Rohit Talwani 8 ;<br />

1 Hepatology, Carolinas Medical Center, Charlotte, NC; 2 Mount<br />

Vernon Endoscopy and Liver Center, Alexandria, VA; 3 Texas<br />

Clinical Research Institute, Dallas, TX; 4 Virginia Commonwealth<br />

University School of Medicine, Richmond, VA; 5 Atlanta Medical<br />

Center, Atlanta, GA; 6 Digestive Disease Associates, Baltimore,<br />

MD; 7 Tulane University School of Medicine, New Orleans, LA;<br />

8 Merck & Co., Inc., Kenilworth, NJ; 9 Baylor College of Medicine,<br />

Houston, TX<br />

Background: Interferon-based HCV treatments are generally<br />

less effective in black compared with nonblack patients (pts).<br />

In phase 2/3 trials, the fixed-dose combination of grazoprevir<br />

100 mg (NS3/4A protease inhibitor) with elbasvir 50 mg<br />

(NS5A inhibitor) (GZR/EBR), with/without (±) ribavirin (RBV)<br />

had a favorable efficacy/safety profile in a diverse population<br />

of GT1/4/6-infected pts. We evaluated the efficacy of GZR/<br />

EBR ± RBV among self-identified black pts enrolled in this program.<br />

Methods: Black and nonblack HCV-infected pts were<br />

drawn from 2 pooled datasets of phase 2/3 <strong>studies</strong> of GZR/<br />

EBR: (1) treatment-naive (TN) pts and (2) treatment-experienced<br />

(TE) pts (who previously failed peginterferon/RBV ± first-generation<br />

protease inhibitor). Pts received GZR/EBR ± RBV for 12<br />

weeks (TN, TE) or 16/18 weeks (TE). Efficacy was measured<br />

as the proportion of pts who achieved a sustained virologic<br />

response, defined as HCV RNA below limits of quantification,<br />

Disclosures:<br />

Philippe J. Zamor - Grant/Research Support: AbbVie, Bristol Myers Squibb,<br />

Gilead, Merck & Co. ; Speaking and Teaching: AbbVie, Bristol Myers Squibb,<br />

Janssen<br />

Jonathan McCone - Speaking and Teaching: Schering-Plough, Roche, Schering-Plough,<br />

Roche, Schering-Plough, Roche, Schering-Plough, Roche<br />

John M. Vierling - Advisory Committees or Review Panels: Abbvie, Bristol-Meyers-Squibb,<br />

Gilead, Hyperion, Intercept, Janssen, Novartis, Merck, Sundise,<br />

HepQuant, Salix, Immuron, Exalenz, Chronic Liver Disease Foundation; Board<br />

Membership: Clinical Research Centers of America, LLC; Grant/Research Support:<br />

Abbvie, Bristol-Meyers-Squibb, Eisai, Gilead, Hyperion, Intercept, Janssen,<br />

Novartis, Merck, Sundise, Ocera, Mochida, Immuron, Exalenz, Conatus; Speaking<br />

and Teaching: GALA, Chronic Liver Disease Foundation, ViralEd, Chronic<br />

Liver Disease Foundation, Clinical Care Options<br />

Reem H. Ghalib - Grant/Research Support: Bristol Myers Squibb Pharmaceuticals,<br />

Vertex Pharmaceuticals, Janssen, Merck, Genentech, Idenix, Zymogenetics,<br />

Pharmasset, Anadys, Duke Clinical Research Institute, Achillion, Boehringer Ingelheim,<br />

Gilead Pharmaceuticals, Virochem Pharmaceuticals, Abbott, Medtronic<br />

Inc, Novartitis, Roche, Schering Plough, Salix, tibotec, Inhibitex, Takeda, Abbvie<br />

Brian Pearlman - Grant/Research Support: Merck, BI, BMS, J&J, Abbvie, Gilead;<br />

Speaking and Teaching: Gilead, Abbvie<br />

Natarajan Ravendhran - Grant/Research Support: Schering-Plough, Roche, Gilead,<br />

Bristol-Myers Squibb, Schering-Plough, Roche, Gilead, Bristol-Myers Squibb,<br />

Schering-Plough, Roche, Gilead, Bristol-Myers Squibb, Schering-Plough, Roche,<br />

Gilead, Bristol-Myers Squibb; Speaking and Teaching: onyx and otsuka, onyx<br />

and otsuka, onyx and otsuka, onyx and otsuka<br />

Luis A. Balart - Advisory Committees or Review Panels: abbvie, Genentech;<br />

Grant/Research Support: Merck, Genentech, Bayer, conatus, Ocera, Hyperion,<br />

Gilead Sciences, Bristol Myers Squibb, abbvie, Vertex, Merck, Genentech,<br />

Bayer, Conatus, Ocera, Hyperion, Gilead Sciences, Bristol Myers Squibb,<br />

Mochida, Eisai, Vertex, takeda, GI Dynamics, tobira; Speaking and Teaching:<br />

Merck, Merck, Merck, Merck, Abbvie, janssen<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Peggy Hwang - Employment: Merck, Merck<br />

Bach-Yen T. Nguyen - Employment: Merck<br />

Janice Wahl - Employment: Merck & Co,<br />

Eliav Barr - Employment: Merck<br />

Rohit Talwani - Employment: Merck


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 565A<br />

The following authors have nothing to disclose: Velimir A. Luketic<br />

716<br />

An integrated safety analysis of daclatasvir + sofosbuvir,<br />

with or without ribavirin, in patients with chronic<br />

HCV infection<br />

Charles S. Landis 1 , David R. Nelson 2 , Mark S. Sulkowski 3 , Peter<br />

J. Ruane 4 , Anthony M. Mills 5 , Fred Poordad 6 , David L. Wyles 7 ,<br />

Rafia Bhore 8 , Peter Ackerman 9 , Khurram Rana 9 , Eugene S. Swenson<br />

9 , Stephanie Noviello 8 ; 1 University of Washington School of<br />

Medicine, Seattle, WA; 2 University Of Florida, Gainesville, FL;<br />

3 Johns Hopkins University, Baltimore, MD; 4 Ruane Medical and<br />

Liver Health Institute, Los Angeles, CA; 5 Anthony Mills MD, Inc,<br />

Los Angeles, CA; 6 Texas Liver Institute, University of Texas Health<br />

Sciences, San Antonio, TX; 7 University of California, San Diego,<br />

CA; 8 Bristol-Myers Squibb, Princeton, NJ; 9 Bristol-Myers Squibb,<br />

Wallingford, CT<br />

Background: The pangenotypic combination of daclatasvir<br />

(DCV) and sofosbuvir (SOF), with or without ribavirin (RBV),<br />

has achieved high SVRs (82%–98%) in multiple patient populations<br />

with HCV infection. We evaluated the safety of DCV+-<br />

SOF±RBV by pooling individual patient data from phase 2 and<br />

3 <strong>studies</strong>. Methods: Data were derived from treatment-naive<br />

or -experienced patients with HCV infection and advanced<br />

cirrhosis or post-transplant recurrence (ALLY-1), HIV/HCV coinfection<br />

(ALLY-2), HCV genotype (GT) 3 infection (ALLY-3), and<br />

chronic HCV infection (AI444-040). Overall, 679 patients<br />

received DCV+SOF±RBV for 8, 12, or 24 weeks. Pooled<br />

data were analyzed for on-treatment adverse events (AEs),<br />

serious AEs, AE-related discontinuations, and grade 3/4 AEs<br />

and lab abnormalities. Results: Patients were 67% male, 80%<br />

white/16% black, 62% naive; median age was 55 years and<br />

18% had cirrhosis, of whom half had hepatic decompensation<br />

(Child-Pugh class B and C). 49% of patients had HCV<br />

GT1a infection, 13% GT1b, 7% GT2, 30% GT3, 1% GT4,<br />

0.1% GT6. Overall, 98% of patients completed treatment.<br />

The most common AEs (any grade) were fatigue, headache,<br />

and nausea. Serious AEs, grade 3/4 AEs, and AE-related discontinuations<br />

were infrequent but more common in patients<br />

receiving RBV (Table); few were treatment-related. All 4 deaths<br />

were posttreatment and unrelated. Safety with DCV+SOF was<br />

comparable in patients with or without compensated cirrhosis.<br />

DCV+SOF+RBV patients with cirrhosis, most with hepatic<br />

decompensation, had slightly higher rates of serious AEs, anemia,<br />

and liver-related lab abnormalities than those without<br />

cirrhosis. There were few safety signals in this diverse population<br />

receiving a broad range of concomitant medications,<br />

including antiretroviral and immunosuppressive agents. Conclusion:<br />

DCV+SOF±RBV is associated with low rates of safety<br />

events. The presence of GT3 infection, HIV/HCV coinfection,<br />

advanced cirrhosis, or post-transplant HCV recurrence has minimal<br />

impact on safety, suggesting that DCV+SOF±RBV is a safe<br />

and well-tolerated treatment for a broad range of patients with<br />

chronic HCV infection.<br />

a<br />

ALLY-2, ALLY-3, AI444040<br />

b<br />

ALLY-1, AI444040<br />

c<br />

Includes ALLY-1 post-transplant patients with cirrhosis status<br />

missing<br />

d<br />

Excludes 5 patients with study drug overdoses reported as SAEs<br />

Disclosures:<br />

Charles S. Landis - Grant/Research Support: Gilead, Abbvie, BMS<br />

David R. Nelson - Advisory Committees or Review Panels: Merck; Grant/Research<br />

Support: Abbot, BMS, Beohringer Ingelheim, Gilead, Genentech, Merck, Bayer,<br />

Idenix, Vertex, Jansen<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Peter J. Ruane - Advisory Committees or Review Panels: BMS; Consulting: Gilead,<br />

Abbvie, Janssen; Grant/Research Support: BMS, Gilead, Merck, Abbvie, Idenix,<br />

Idenix, Janssen, Viiv; Speaking and Teaching: Gilead, Merck, Abbvie, Abbvie,<br />

Janssen; Stock Shareholder: Gilead, Gilead<br />

Anthony M. Mills - Advisory Committees or Review Panels: Gilead, ViiV, Merck;<br />

Grant/Research Support: Gilead, ViiV, Merck, BMS; Speaking and Teaching:<br />

Gilead<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

David L. Wyles - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Janssen, Merck; Grant/Research Support: Gilead, Merck, Bristol Myers Squibb,<br />

AbbVie, Tacere<br />

Peter Ackerman - Employment: Bristol-Myers, Squibb<br />

Khurram Rana - Employment: Bristol-Myers Squibb<br />

Eugene S. Swenson - Employment: Bristol-Myers Squibb<br />

Stephanie Noviello - Consulting: Merck/Schering-Plough; Employment: Bristol-Myers<br />

Squibb, Merck/Schering-Plough; Stock Shareholder: Merck/Schering-Plough,<br />

J&J<br />

The following authors have nothing to disclose: Rafia Bhore<br />

717<br />

C-EDGE TN: Impact Of 12-Week Oral Regimen Of Grazoprevir<br />

(GZR, MK-5172)/Elbasvir (EBR, MK-8742) On<br />

Patient-Reported Outcomes (PROs) In Treatment-Naïve<br />

Patients With Chronic Hepatitis C Virus (HCV) Genotype<br />

(GT) 1, 4, Or 6 Infection<br />

Jean Marie Arduino, Yang Wang, Deborah D. Brown, Shazia<br />

Khawaja, Elisa Martinez, Joan R. Butterton, Michael Robertson,<br />

Chizoba Nwankwo, T. Christopher Mast; Merck & Co., Inc.,<br />

Kenilworth, NJ<br />

Background:Chronic HCV infection negatively impacts patients’<br />

health and well-being.A Phase 3, double-blind, placebo-control,<br />

randomized trial of an oral fixed-dosed combination of GZR<br />

100 mg/EBR 50 mg once daily for 12 weeks was conducted<br />

among HCV GT1-,4-,or 6-infected patients.GZR/EBR was highly<br />

effective (SVR12:94.6%(95% CI: 91.5,96.8%)) and generally<br />

well-tolerated,with a similar safety profile to placebo.Our aim<br />

was to evaluate whether HCV treatment with GZR/EBR altered<br />

PROs.Methods:Subjects completed 5 PROs at baseline, treatment<br />

week 4 (TW4),TW12, follow-up week 4(FW4):SF-36v2®,<br />

EQ-VAS,FACIT-Fatigue Scale,Chronic Liver Disease Question-


566A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

naire-HCV(CLDQ-HCV) and Work Productivity and Activity<br />

Impairment.421 patients were randomized and received ≥1<br />

dose of study drug(GZR/EBR:n=316,Placebo:n=105).Mean<br />

change from baseline in PRO scores (95% CIs) by treatment<br />

group and treatment group differences in mean change scores<br />

(95%CIs) were estimated.Results:At TW12, in general, GZR/<br />

EBR had more favorable changes than placebo [Figure].GZR/<br />

EBR group had mean improvement (95% CIs exclude 0) in<br />

general and HCV-specific HRQOL. Placebo group had mean<br />

improvement in HCV-specific HRQOL, but a mean decline in<br />

Social Functioning (SF).Treatment group differences favored<br />

GZR/EBR over placebo in MCS (2.33 (95%CI:0.28,4.37))<br />

and SF (7.25 (95%CI:2.47,12.04)). At FW4, for GZR/EBR<br />

(data not shown),mean improvements in general and HCV-specific<br />

HRQOL,fatigue levels,and activity impairment were<br />

observed.Treatment group differences favored GZR/EBR over<br />

placebo in Role Limitations-Physical(4.97(95%CI:0.17,9.77)),-<br />

General Health(4.94(95%CI:1.30,8.59)),S-<br />

F(6.97(95%CI:2.38,11.57)) and overall health(EQ-VAS<br />

5.42(95% CI:1.87,8.97)).Conclusion:Treatment with GZR/<br />

EBR had a positive impact on PROs compared to placebo.In<br />

addition, changes in PROs were substantially more favorable<br />

than the large declines in PROs historically associated with<br />

interferon and ribavirin-containing regimens.<br />

Disclosures:<br />

Jean Marie Arduino - Employment: Merck & Co., Inc<br />

Yang Wang - Employment: Merck<br />

Deborah D. Brown - Employment: Merck & Co., Inc.; Stock Shareholder: Merck<br />

& Co., Inc<br />

Shazia Khawaja - Employment: Merck<br />

Joan R. Butterton - Employment: Merck Sharp & Dohme Corp.; Stock Shareholder:<br />

Merck Sharp & Dohme Corp.<br />

Michael Robertson - Employment: Merck; Stock Shareholder: Merck<br />

Chizoba Nwankwo - Employment: Merck<br />

T. Christopher Mast - Employment: Merck Research Laboratories<br />

The following authors have nothing to disclose: Elisa Martinez<br />

718<br />

Characterization of HCV Resistance from a 3-Day<br />

Monotherapy Study of GS-9857, a Novel Pangenotypic<br />

NS3/4A Protease Inhibitor<br />

Eric Lawitz 2 , Hadas Dvory-Sobol 1 , Jenny C. Yang 1 , Luisa M.<br />

Stamm 1 , James Taylor 1 , Diana M. Brainard 1 , Michael D. Miller 1 ,<br />

Hongmei Mo 1 , Maribel Rodriguez-Torres 3 ; 1 Gilead Sciences, Foster<br />

City, CA; 2 The Texas Liver Institute, University of Texas Health<br />

Science Center, San Antonio, TX; 3 Fundacion de Investigacion,<br />

Rio Pedras, PR<br />

Background: GS-9857 is a pangenotypic hepatitis C virus<br />

(HCV) NS3/4A protease inhibitor (PI) with potent antiviral<br />

activity against HCV genotypes (GT) 1-6 and improved<br />

coverage of GT1 NS3 resistance associated variants (RAVs)<br />

associated with other HCV PIs, in preclinical replicon <strong>studies</strong>.<br />

In a 3-day monotherapy study in patients infected with HCV<br />

GT 1, 2, 3 or 4, GS-9857 was well tolerated and resulted<br />

in maximal mean viral load reduction >3 log 10<br />

IU/mL at the<br />

100mg dose across all genotypes. This analysis characterizes<br />

the HCV NS3 sequence data from this phase 1b study. Methods:<br />

The full-length NS3 gene was amplified and successfully<br />

deep sequenced using MiSeq for 58 patients at baseline and<br />

53 patients post-baseline using a 1% cutoff for reporting. NS3<br />

RAVs were defined as amino acid substitutions historically<br />

associated with resistance to HCV PIs, including positions 36,<br />

41, 43, 54, 55, 80, 122, 155, 156, 168 and 170 of the NS3<br />

protease gene of HCV GT1. Results: Baseline HCV NS3 RAVs<br />

were present in the HCV of 37% (9/24) of patients with GT1a,<br />

17% (1/6) of patients with GT1b and 16% (3/19) of patients<br />

with GT3; there were no baseline NS3 RAVs present in patients<br />

with GT2 (n=5) or GT4 (n=4). Mean maximal viral load reductions<br />

over 3 days of GS-9857 administration were similar in<br />

patients with and without baseline NS3 RAVs. The table summarizes<br />

the number of patients with treatment-emergent (TE)<br />

NS3 RAVs by genotype. Overall, by deep sequencing analysis,<br />

the majority of patients (74%, 39/53) did not have TE NS3<br />

RAVs, and there were no TE NS3 RAVs in patients with GT2<br />

or 4. A156T or A156V were the most prevalent substitutions<br />

in patients with GT1a or 1b infection and were not observed<br />

in patients with GT3 infection. One GT3 patient had low levels<br />

of A156P (2.3%). Conclusions: The presence of NS3 RAVs at<br />

baseline had no impact on treatment response to 3 days of<br />

monotherapy with GS-9857 in this phase 1b study. The lack of<br />

selection of NS3 RAVs in the majority of patients demonstrates<br />

an improved resistance profile of GS-9857.<br />

Number of Patients with Treatment Emergent (TE) NS3 RAVs Following<br />

3 Days Monotherapy with GS-9857 by Deep Sequencing<br />

Analysis<br />

Disclosures:<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 567A<br />

Hadas Dvory-Sobol - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Jenny C. Yang - Employment: Gilead Sciences, Inc<br />

Luisa M. Stamm - Employment: Gilead Sciences<br />

James Taylor - Employment: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Michael D. Miller - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Maribel Rodriguez-Torres - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen R&D Ireland; Consulting: Glaxo Smith Kline, Janssen R&D Ireland,<br />

Theravance; Grant/Research Support: Merck, Bristol-Myers Squibb, Merck,<br />

Pfizer, Gilead, Johnson & Johnson, Beckman Coulter, Theravance<br />

719<br />

Pharmacokinetics of Coadministration of Pan-Genotypic,<br />

Direct Acting Antiviral Agents, ABT-493 and ABT-<br />

530, with or without Ribavirin for 12 weeks in HCV<br />

Infected Subjects without Cirrhosis<br />

Chih-Wei Lin, Wei Liu, Armen Asatryan, Stanley Wang, Federico<br />

J. Mensa, Jens Kort, Sandeep Dutta; Abbvie, North Chicago, IL<br />

Purpose: A direct acting antiviral agent (DAA) combination of<br />

ABT-493 (NS3/4A protease inhibitor discovered by AbbVie<br />

and Enanta) and ABT-530 (NS5A inhibitor) is being developed<br />

for the treatment of chronic hepatitis C (HCV) genotype<br />

(GT) 1-6 infection. In two Phase 2 <strong>studies</strong> (SURVEYOR-1 and<br />

-2), the ABT-493 and ABT-530 combination has demonstrated<br />

high sustained virologic response rates in GT1-, GT2- and<br />

GT3-infected subjects without cirrhosis following 12-weeks of<br />

treatment. Pharmacokinetics of ABT-493 and ABT-530 with or<br />

without ribavirin (RBV) were evaluated in these <strong>studies</strong>. Methods:<br />

Two open-label, multicenter <strong>studies</strong>, SURVEYOR-1 and<br />

-2, were conducted evaluate the efficacy, safety, and pharmacokinetics<br />

of co-administration of ABT-493 (200 or 300<br />

mg QD) and ABT-530 (40 or 120 mg QD) with or without<br />

RBV in GT1-, GT2- or GT3-infected subjects. Blood samples for<br />

pharmacokinetic analysis were collected throughout the study<br />

treatment period. ABT-493 and ABT-530 pharmacokinetics following<br />

a single dose (Day 1) and at steady state (Week 4)<br />

were assessed by non-compartmental methods. Results: A total<br />

of 274 subjects received ABT-493 and ABT-530 with or without<br />

RBV. Both ABT-493 and ABT-530 showed rapid absorption<br />

with Tmax ranging from 2 to 4 hours. Steady state ABT-493<br />

exposure (area under the curve from 0 to 4 hour) following<br />

300 mg was 2570 ng.h/mL, approximately 3.7-fold of the<br />

exposure following 200 mg administration. Coadministration<br />

of either 40 mg or 120 mg ABT-530 each with 200 mg ABT-<br />

493 resulted in ABT-530 exposure of 157 or 372 ng.h/mL,<br />

respectively. ABT-493 300 mg increased 120 mg ABT-530<br />

exposure by an additional 20% to 30%. Minimal accumulation<br />

in ABT-493 or ABT-530 exposure was observed at Week 4<br />

compared to Day 1. ABT-530 had minimal impact on ABT-493<br />

exposures, however, ABT-493 200 mg and 300 mg increased<br />

120 mg ABT-530 exposures to 3- to 4-fold of ABT-530 exposures<br />

when administered alone. HCV genotype and RBV coadministration<br />

had no impact on ABT-493 or ABT-530 exposures.<br />

Conclusions: ABT-493 exhibited non-linear pharmacokinetics<br />

with more than dose-proportional increase in exposures with<br />

increasing doses, while ABT-530 exposures increased in an<br />

approximately dose-proportional manner when coadministered<br />

with ABT-493. ABT-530 had minimal impact on ABT-<br />

493, while ABT-493 increased ABT-530 exposures, with the<br />

increase in ABT-530 exposure being dependent on ABT-493<br />

dose. ABT-493 or ABT-530 had minimal accumulation in exposures<br />

following multiple dosing in HCV-infected subjects.<br />

Disclosures:<br />

Chih-Wei Lin - Employment: Abbvie<br />

Wei Liu - Employment: AbbVie<br />

Armen Asatryan - Employment: AbbVie<br />

Stanley Wang - Employment: AbbVie<br />

Federico J. Mensa - Employment: Abbvie Inc.; Stock Shareholder: Abbvie Inc.<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

Sandeep Dutta - Employment: AbbVie; Stock Shareholder: AbbVie<br />

720<br />

Daclatasvir exposure does not explain lower sustained<br />

virologic response rates in cirrhotic patients with HCV<br />

genotype 3 following 12 weeks of daclatasvir plus<br />

sofosbuvir treatment<br />

Timothy Eley, Reena Wang, Shu-Pang Huang, Yash Gandhi,<br />

Brenda Cirincione, Frank LaCreta, Tushar Garimella; Bristol-Myers<br />

Squibb, Princeton, NJ<br />

Background: The ALLY-3 study evaluated daclatasvir (DCV; 60<br />

mg daily) plus sofosbuvir (SOF; 400 mg daily) for 12 weeks in<br />

treatment-naive and -experienced patients with HCV genotype<br />

3 (GT 3) infection. Sustained virologic response at posttreatment<br />

week 12 (SVR12) was lower in patients with compensated<br />

cirrhosis than in those without (63% [20/32] vs 96%<br />

[105/109]), mostly due to posttreatment relapse. Pharmacokinetic<br />

(PK) exposure to DCV in ALLY-3 was evaluated in patients<br />

with and without cirrhosis to explore this observation. Methods:<br />

Systemic exposure to total (protein-bound + unbound) DCV on<br />

dosing Day 1 was evaluated in a prespecified 24-hour intensive<br />

PK substudy of cirrhotic GT 3 patients in ALLY-3 (N=10)<br />

using noncompartmental methods. DCV exposure parameters<br />

were compared to those of non-cirrhotic ALLY-3 patients derived<br />

from population PK (PopPK) modeling, and against historical<br />

single-dose DCV data in HCV-uninfected subjects with hepatic<br />

impairment. Model-predicted steady-state average DCV concentrations<br />

across the dosing interval (C avgss<br />

) for cirrhotic and<br />

non-cirrhotic ALLY-3 patients with and without SVR12 were<br />

compared visually in an exploratory exposure-response assessment.<br />

Results: Taking into account the historical DCV accumulation<br />

ratio in HCV-infected patients (1.2–1.4), median observed<br />

Day 1 AUC 0-24h<br />

in cirrhotic GT 3 patients in ALLY-3 (6210<br />

[range 3141–10895] ng.h/mL) was comparable to PopPK<br />

model estimates of median steady-state AUC 0-24h<br />

in both cirrhotic<br />

patients (7980 [3744–17856] ng.h/mL) and non-cirrhotic<br />

patients (9192 [3696–19776] ng.h/mL). Similarly, after<br />

considering the accumulation, the median observed Day 1<br />

AUC 0-24h<br />

in cirrhotic patients was comparable to historical single-dose<br />

median AUC inf<br />

for DCV 30 mg in HCV-uninfected<br />

subjects with Child-Pugh class A, B or C hepatic impairment<br />

normalized to 60 mg (9638–11398 ng.h/mL). Calculated<br />

median C avgss<br />

exposures were similar in cirrhotic and non-cirrhotic<br />

ALLY-3 patients in the PopPK model (non-cirrhotic: 358<br />

ng/mL without SVR12 [n=4], 408 ng/mL with SVR12 [n=105];<br />

cirrhotic 283 ng/mL without SVR12 [n=12], 382 ng/mL with<br />

SVR12 [n=20]). Despite a weak trend towards lower C avgss<br />

in<br />

patients without SVR12 there was substantial overlap between<br />

those who did and those who did not achieve SVR12 in both<br />

groups, though numbers were small. Conclusions: The data<br />

indicate that higher relapse rates among GT 3-infected cirrhotic<br />

patients in ALLY-3 cannot be explained by DCV exposure<br />

alone. These data suggest that cirrhotic patients with HCV GT 3<br />

may require longer (>12 weeks) treatment duration with DCV+-<br />

SOF or the addition of ribavirin.<br />

Disclosures:<br />

Timothy Eley - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb


568A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Shu-Pang Huang - Employment: Bristol-Myers Squibb Company; Stock Shareholder:<br />

Bristol-Myers Squibb Company<br />

Yash Gandhi - Employment: Bristol-Myers Squibb<br />

Brenda Cirincione - Employment: BMS<br />

Frank LaCreta - Employment: Bristol-Myers Squibb; Management Position: Bristol-Myers<br />

Squibb; Stock Shareholder: Bristol-Myers Squibb<br />

Tushar Garimella - Employment: Bristol Myers-Squibb; Stock Shareholder: Abbvie,<br />

Bristol Myers-Squibb<br />

The following authors have nothing to disclose: Reena Wang<br />

721<br />

Pharmacokinetics of Narlaprevir an NS3/4A Protease<br />

Inhibitor with or without Ritonavir Following Single<br />

Dose Use in Patients with Compensated Cirrhosis and<br />

Healthy Volunteers<br />

Dmitry Koloda 1 , Natalia Tikhonova 1 , Irina Malaya 2 , Miroslav<br />

Ryska 3 , Mikhail Samsonov 1 , Vasily Isakov 4 ; 1 Medical Department,<br />

R-Pharm, Moscow, Russian Federation; 2 Ascent Clinical Research<br />

Solutions, Moscow, Russian Federation; 3 Quinta Analytica,<br />

Prague, Czech Republic; 4 Department of gastroenterology & hepatology,<br />

Institute of Nutrition, Moscow, Russian Federation<br />

Background and Aims: Narlaprevir (NVR), a potent ritonavir-boosted<br />

inhibitor of HCV NS3 protease, is currently being<br />

developed for therapy of chronic HCV genotype 1 infection<br />

in Russia and CIS countries. The purpose of this study was to<br />

evaluate the pharmacokinetics (PK) after a single oral dose of<br />

NVR alone and in combination with ritonavir (RTV) in healthy<br />

controls and in patients with compensated cirrhosis. Methods:<br />

This was a two-part, open-label, parallel group, single-dose<br />

phase I study in patients with compensated cirrhosis Child-Pugh<br />

class A (CPA) (n=16), and matched healthy subjects (n=16),<br />

all Caucasians. In Part 1 of the study 8 healthy adult subjects<br />

and 8 patients with CPA received single dose of NVR 200<br />

mg under fed condition. In Part 2 of the study 8 healthy subjects<br />

and 8 CPA patients received NVR 100 mg (dose was<br />

decreased based on the Part 1 results) in combination with RTV<br />

100 mg. Noncompartmental model was used for PK analysis of<br />

NVR in plasma and urine. Simulation of PK parameters for NVR<br />

at steady state was performed. Geometric least square means<br />

(GMEAN) and 90% confidence intervals were calculated.<br />

Results: In Part I GMEAN C max,<br />

and AUC (0-inf)<br />

were 1.6 and 2.7<br />

times higher in CPA patients vs. healthy subjects, respectively<br />

(Table 1). Simulated steady-state (tau=24 h) GMEAN C max,<br />

and<br />

AUC tau,<br />

in patients with CPA were 167% and 237% vs. healthy<br />

subjects respectively (Table 2). Increased NVR exposure in<br />

patients with CPA was well tolerated and not associated with<br />

increased adverse events rate. There were no significant difference<br />

in plasma NVR exposure in CPA patients and healthy subjects<br />

after single dose NVR 100 mg in combination with 100<br />

mg RTV and after steady-state simulation (tau=24 h): GMEAN<br />

C max,<br />

and AUC tau,<br />

were 107% and 105% in patients with CPA<br />

vs. healthy subjects, respectively (Table 2). Conclusions: NVR<br />

exposure after a single dose 200 mg was higher in patients<br />

with CPA than in healthy subjects. No significant effect on<br />

NVR exposure in CPA subjects compared to healthy subjects<br />

was found when NRV 100 mg was co-administered with RTV<br />

100 mg.<br />

Disclosures:<br />

Dmitry Koloda - Employment: R-Pharm<br />

Natalia Tikhonova - Employment: R-Pharm<br />

Mikhail Samsonov - Employment: RPharm<br />

Vasily Isakov - Advisory Committees or Review Panels: Abbvie, Bristol-Myers<br />

Squibb, Gilead, Janssen, Merck, Vertex, R-Pharm; Consulting: Bristol-Myers<br />

Squibb, Merck; Speaking and Teaching: Bristol-Myers Squibb, Janssen, Merck<br />

The following authors have nothing to disclose: Irina Malaya, Miroslav Ryska<br />

722<br />

Ombitasvir/Paritaprevir/r and Dasabuvir with Ribavirin<br />

for HCV Genotype 1 Patients with Decompensated Cirrhosis<br />

Parvez S. Mantry 1 , John Hanson 2 , Roger Trinh 3 , Alnoor Ramji 4 ,<br />

Linda Fredrick 3 , Manal Abunimeh 3 , Leticia Canizaro 3 , Li Liu 3 ,<br />

Nancy Shulman 3 , Stuart C. Gordon 5 ; 1 The Liver Institute at Methodist<br />

Dallas, Dallas, TX; 2 Charlotte Gastroenterology and Hepatology,<br />

PLLC, Charlotte, NC; 3 AbbVie, Inc., North Chicago, IL;<br />

4 University of British Columbia, Vancouver, BC, Canada; 5 Henry<br />

Ford Health System, Detroit, MI<br />

Objective: Patients with HCV and decompensated cirrhosis are<br />

at greatest risk for life-threatening complications leading to<br />

liver transplantation. The 3 direct-acting antiviral (3D) regimen<br />

of ombitasvir (OBV) co-formulated with paritaprevir/ritonavir<br />

(PTV/r, identified by AbbVie and Enanta) and dasabuvir (DSV)<br />

with ribavirin (RBV) has high efficacy in HCV genotype (GT)<br />

1-infected patients with compensated cirrhosis. We assessed<br />

the safety, pharmacokinetics, and efficacy of the 3D + RBV<br />

regimen in an initial cohort of HCV GT1-infected patients with<br />

decompensated cirrhosis. Methods: Patients with decompensated<br />

cirrhosis (Child-Pugh score 7 – 9) and the following laboratory<br />

parameters at the time of screening were enrolled:<br />

platelet count ≥25 x 10 9 /L, serum albumin ≥2.8 g/dL, MELD<br />

score ≤18. Those with GT1b or GT1a infection received standard<br />

dosing of OBV/PTV/r (25/150/100 mg/day) and DSV<br />

(250 mg twice daily) plus weight-based RBV (1000 – 1200<br />

mg/day) for 12 or 24 weeks, respectively. A data review<br />

was planned once the initial cohort reached week 12 of treatment,<br />

prior to enrollment of a larger cohort. Results: Baseline<br />

characteristics of the 11 patients enrolled are detailed in the<br />

Table. Suppression of HCV RNA below the level of quantitation<br />

occurred in 7/11 (64%) patients by treatment week 2 and in<br />

all patients by week 6. At treatment week 12, all 11 patients<br />

remained HCV RNA suppressed below the level of detection.<br />

Sustained virologic response at post-treatment week 12 will


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 569A<br />

be presented. No patients prematurely discontinued from the<br />

study, experienced a treatment-emergent adverse event (AE)<br />

leading to interruption of study drug (Table), or experienced<br />

post-baseline ALT elevations. Grade 3 bilirubin elevations,<br />

predominantly indirect, have occurred in 7/11 patients. The<br />

median starting RBV dose was 1200 mg. Hemoglobin declines<br />

below 10 g/dL occurred in 4 patients, 3 of whom modified<br />

RBV dose; none interrupted study drug. Cross-study comparison<br />

revealed that the range of PTV/r, OBV, DSV and DSV metabolite<br />

exposures observed were comparable or slightly higher<br />

than those observed in patients with Child-Pugh A cirrhosis.<br />

Conclusions: In this initial cohort to assess safety of the 3D +<br />

RBV regimen in HCV GT1-infected patients with decompensated<br />

cirrhosis, treatment was generally well tolerated. Further<br />

assessment of this regimen in patients with decompensated cirrhosis<br />

is underway with additional GT1-infected patients who<br />

will be treated with 3D + RBV, and GT4-infected patients who<br />

will be treated with OBV/PTV/r + RBV.<br />

Disclosures:<br />

Parvez S. Mantry - Consulting: Salix, Gilead, Janssen, Abbvie, BMS; Grant/<br />

Research Support: Salix, Merck, Gilead, Boehringer-Ingelheim, Mass Biologics,<br />

Vital Therapies, Santaris, mass biologics, Bristol-Myers Squibb, Abbive, Bayer-Onyx,<br />

Shinogi, Tacere, Intercept; Speaking and Teaching: Gilead, Janssen,<br />

Salix<br />

John Hanson - Advisory Committees or Review Panels: Abbvie, Gilead; Grant/<br />

Research Support: Abbvie, Takeda, Salix, Gilead, Celltrion, Genentech; Speaking<br />

and Teaching: Janssen, Abbvie, Prometheus, Takeda<br />

Alnoor Ramji - Advisory Committees or Review Panels: Roche, Merck, Gilead,<br />

Janssen, Boehringer Ingelheim; Grant/Research Support: BMS, Roche, Merck,<br />

Gilead, Vertex, Novartis, Abbvie, Boehringer Ingelheim<br />

Linda Fredrick - Employment: AbbVie; Management Position: AbbVie; Stock<br />

Shareholder: AbbVie<br />

Manal Abunimeh - Employment: AbbVie<br />

Nancy Shulman - Employment: Abbvie<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

The following authors have nothing to disclose: Roger Trinh, Leticia Canizaro,<br />

Li Liu<br />

723<br />

Absence of significant drug-drug interactions between<br />

next generation direct acting antivirals ABT-493 and<br />

ABT-530 and methadone or buprenorphine/naloxone<br />

in subjects on opioid maintenance therapy<br />

Matthew P. Kosloski, Sandeep Dutta, Weihan Zhao, Armen Asatryan,<br />

Jens Kort, Wei Liu; AbbVie, North Chicago, IL<br />

Purpose: A next generation direct acting antiviral (DAA) combination<br />

of ABT-493 (NS3/4A protease inhibitor discovered by<br />

AbbVie and Enanta) + ABT-530 (NS5A inhibitor) is being developed<br />

for the treatment of chronic hepatitis C (HCV) genotype<br />

1-6 infection. ABT-493 + ABT-530 combination demonstrated<br />

high sustained virologic response rate in Phase 2 <strong>studies</strong>. A<br />

Phase 1, open-label study was conducted to assess the pharmacokinetics,<br />

safety and tolerability of ABT-493 + ABT-530 and<br />

methadone or buprenorphine/naloxone. Methods: Otherwise<br />

healthy adults subjects on individualized regimens of methadone<br />

(n=12) or buprenorphine/naloxone (n=12) for opioid<br />

addiction received ABT-493 300 mg QD + ABT-530 120 mg<br />

QD for 7 days. Intensive blood sampling for determination<br />

of methadone, buprenorphine, norbuprenorphine, naloxone,<br />

ABT-493 and ABT-530 concentrations was performed and<br />

pharmacokinetic parameters (maximum observed concentration<br />

[C max<br />

], area under the concentration-time curve [AUC 24<br />

or<br />

AUC t<br />

] and trough concentration [C 24<br />

]) were estimated. Safety<br />

and tolerability were assessed throughout the study. Potential<br />

opioid withdrawal or overdose symptoms (pharmacodynamics)<br />

were assessed with validated instruments including the short<br />

opiate withdrawal scale, desire for drugs questionnaire, and<br />

pupillometry measurements throughout the study. Results: For<br />

subjects on methadone maintenance therapy, dose-normalized<br />

C max<br />

, AUC 24<br />

, and C 24<br />

for R- and S-methadone were unaffected<br />

by coadministration with ABT-493 and ABT-530 at steady-state<br />

(≤5 % change). For subjects on buprenorphine/naloxone maintenance<br />

therapy, dose-normalized C max<br />

, AUC 24<br />

, and C 24<br />

were<br />

slightly increased for buprenorphine (8%, 17%, and 24%,<br />

respectively) and norbuprenorphine (25%, 30%, and 21%,<br />

respectively) when coadministered with ABT-493 and ABT-530<br />

at steady-state; naloxone dose-normalized C max<br />

and AUC t<br />

were<br />

minimally affected (≤12% change). Pharmacodynamics of the<br />

methadone or buprenorphine/naloxone regimens were not<br />

significantly impacted by coadministration with ABT-493 or<br />

ABT-530 for either regimen. ABT-493 and ABT-530 exposures<br />

following coadministration with methadone or buprenorphine/<br />

naloxone were similar to the observed values in previous <strong>studies</strong>.<br />

Subjects experienced adverse events of mild intensity, with<br />

the most common (reported in >5 subjects) being abdominal<br />

pain, constipation, and headache; all subjects completed the<br />

study. There were no clinically relevant abnormal laboratory<br />

abnormalities, ECG or vital sign findings. Conclusions: No<br />

dose adjustments are required for coadministration of ABT-493<br />

and ABT-530 with methadone or buprenorphine/naloxone.<br />

No pharmacodynamic interaction is expected.<br />

Disclosures:<br />

Sandeep Dutta - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Weihan Zhao - Employment: AbbVie, Inc; Stock Shareholder: AbbVie, Inc, Exxon<br />

Mobile, American Airlines<br />

Armen Asatryan - Employment: AbbVie<br />

Jens Kort - Employment: AbbVie Inc.; Stock Shareholder: AbbVie Inc.<br />

Wei Liu - Employment: AbbVie<br />

The following authors have nothing to disclose: Matthew P. Kosloski


570A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

724<br />

A Pilot Evaluation of Twice Daily Fixed Dose Combination<br />

Asunaprevir + Daclatasvir + Beclabuvir ± Weight<br />

Based Ribavirin in Treatment-Naïve, Non-cirrhotic<br />

Patients with Chronic Genotype 1a Hepatitis- C for<br />

Eight, Six or Four Weeks (RHACE-1)<br />

Gregory J. Botwin 2 , Aliya Asghar 2 , Deval Modi 2 , Dean Bosman 2 ,<br />

Timothy R. Morgan 1,3 ; 1 Gastroenterology (11), VA Long Beach<br />

Healthcare System, Long Beach, CA; 2 Research Service, VA Long<br />

Beach Healthcare System, Long Beach, CA; 3 Division of Gastroenterology,<br />

University of California - Irvine, Irvine, CA<br />

Background: The all oral, fixed-dose combination of daclatasvir,<br />

asunaprevir, and beclabuvir (collectively FDC) without ribavirin<br />

(RBV) achieved sustained virologic response (SVR 12<br />

) rates of<br />

>92% in genotype 1, treatment naïve, non-cirrhotic patients<br />

treated for 12 weeks in a Phase II trial. Purpose: To assess<br />

the feasibility of an 8, 6 and/or 4 week treatment duration<br />

with FDC ± weight based RBV in treatment-naïve, non-cirrhotic,<br />

HCV genotype 1a subjects. Methods: We designed a single<br />

site, open-label, multi-arm, adaptive clinical trial and planned<br />

to enroll 12, non-cirrhotic, genotype (GT)-1a, treatment-naïve<br />

subjects per arm. To minimize the number of subjects exposed<br />

to a potentially ineffective treatment regimen, only one arm was<br />

to be enrolled at a time. Subjects enrolled in the first arm, Arm<br />

A, were to be treated for 8 weeks with FDC + RBV. If a low<br />

relapse rate was observed, then sequentially shorter treatment<br />

durations of 6 and 4 weeks ± RBV would be tested. The results<br />

presented here are for Arms A and B. Results: Arm A (n=12)<br />

had an average age of 60 years, BMI of 28, and baseline<br />

HCV RNA of 6.5 log10 IU/mL. All 12 subjects were male, 7<br />

were non-CC IL28B, and 3 were African-American (AA). All 12<br />

subjects had HCV RNA less than the limit of detection (LOD) at<br />

the end of treatment (Week 8). Five subjects experienced post<br />

treatment viral relapse and the remaining 7 (58%) achieved<br />

SVR 12<br />

. Due to the non-optimal relapse rate in Arm A, Arm B<br />

was adapted to 12 weeks of treatment with RBV and no other<br />

arms were enrolled. Arm B (n=12) had an average age of 60<br />

years, BMI of 27, and baseline HCV RNA of 6.6 log10 IU/<br />

mL. All 12 subjects were male, 10 were non-CC IL28B, and 4<br />

were AA. To date all subjects in Arm B have reached the end<br />

of treatment (Week 12) and all have HCV RNA < LOD. Half<br />

of the subjects in each arm experienced a >2.5g/dL drop in<br />

hemoglobin on-treatment. There were no treatment-related serious<br />

adverse events in either group. Conclusion: Eight weeks,<br />

or less, of treatment with FDC plus RBV is likely to result in a<br />

non-optimal SVR 12<br />

amongst non-cirrhotic, GT-1a subjects. Less<br />

than 12 weeks of treatment with FDC may still be feasible in<br />

easier to treat populations. SVR 12<br />

results will be available for<br />

subjects enrolled in Arm B.<br />

Viral Response<br />

† One subject was excluded (n=11)<br />

‡ Missing data on one subject, counted as failure.<br />

Note: Number and percent of subjects with HCV RNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 571A<br />

Dennis Wolford - Employment: Merck Sharpe & Dohme Corp., a subsidiary of<br />

Merck and Co., Inc.; Stock Shareholder: Merck Sharpe & Dohme Corp., a subsidiary<br />

of Merck and Co., Inc.<br />

William L. Marshall - Employment: Merck<br />

Marian Iwamoto - Employment: Merck Sharp & Dohme Corp.<br />

Joan R. Butterton - Employment: Merck Sharp & Dohme Corp.; Stock Shareholder:<br />

Merck Sharp & Dohme Corp.<br />

The following authors have nothing to disclose: Katherine M. Dunnington, Nadia<br />

Cardillo Marricco, Michael Gartner, Dennis Swearingen, John Brejda, Angela<br />

O. Choi<br />

726<br />

Integrated Safety Analysis of Daclatasvir Plus Sofosbuvir,<br />

With or Without Ribavirin, in Patients With HCV<br />

Genotype 3 Infection<br />

David Bernstein 1 , Charles S. Landis 2 , Eric Lawitz 3 , Anne Luetkemeyer<br />

4 , Melissa Harris 5 , Rafia Bhore 5 , Eugene S. Swenson 6 ,<br />

Peter Ackerman 6 , Khurram Rana 6 , Douglas Dieterich 7 ; 1 Hofstra<br />

North Shore-Long Island Jewish School of Medicine, Manhasset,<br />

NY; 2 University of Washington School of Medicine, Seattle, WA;<br />

3 Texas Liver Institute, University of Texas Health Science Center,<br />

San Antonio, TX; 4 University of California and San Francisco General<br />

Hospital, San Francisco, CA; 5 Bristol-Myers Squibb, Princeton,<br />

NJ; 6 Bristol-Myers Squibb, Wallingford, CT; 7 Icahn School of Medicine<br />

at Mount Sinai, New York, NY<br />

Background: SVR12 rates of up to 98% have been achieved<br />

in the three phase 3 ALLY <strong>studies</strong> with the pangenotypic combination<br />

of daclatasvir (DCV) plus sofosbuvir (SOF), with or<br />

without ribavirin (RBV). In patients with genotype (GT)3 infection<br />

enrolled in these <strong>studies</strong>, SVR12 rates were 88–100%. The<br />

safety profile of DCV+SOF±RBV in patients with GT3 infection<br />

was evaluated using integrated data from the ALLY <strong>studies</strong>.<br />

Methods: Integrated data were derived from 182 GT3-infected<br />

patients enrolled in ALLY-1 (advanced cirrhosis or post-liver<br />

transplant recurrence, N=17), ALLY-2 (HIV-HCV coinfection,<br />

N=13), and ALLY-3 (GT3 infection, N=152). Patients were<br />

treated with DCV 60mg (dose-adjusted for concomitant antiretroviral<br />

therapy in HIV coinfected patients) +SOF 400mg<br />

for 8 (ALLY-2) or 12 weeks (ALLY-2, ALLY-3) or with DCV+-<br />

SOF+RBV for 12 weeks (ALLY-1). Integrated safety data were<br />

analyzed for on-treatment adverse events (AEs), serious AEs<br />

(SAEs), AE-related discontinuations, and grade 3/4 AEs and<br />

lab abnormalities. Results: Patients were 64% male, 91%<br />

white, and 64% HCV treatment-naive; the median age was<br />

55 years and 21% had cirrhosis. In this cohort, there were<br />

no treatment-related SAEs or deaths. There were 7 grade 3/4<br />

AEs and 16 grade 3/4 lab abnormalities. Four AEs led to<br />

discontinuation of study drug (all treatment-related): 3 events<br />

were associated with RBV and led to discontinuation of only<br />

RBV in patients with advanced cirrhosis; 1 event (headache)<br />

in a liver transplant recipient led to discontinuation of all HCV<br />

therapy after 4 weeks, but SVR12 was still achieved. Overall,<br />

the most common AEs (any grade) in the ALLY <strong>studies</strong><br />

were fatigue (19%), headache (17%), and nausea (12%). As<br />

expected, safety events were more frequent in patients with<br />

advanced cirrhosis that received RBV; however, none of these<br />

patients required discontinuation of all study therapies and<br />

5/6 achieved SVR12. Conclusions: Safety events with DCV+-<br />

SOF±RBV were uncommon in patients with HCV GT3 infection.<br />

HIV coinfection had no notable impact on safety parameters;<br />

events were more frequent in patients with advanced cirrhosis<br />

that received RBV, but had minimal impact on SVR12. These<br />

results support the use of DCV+SOF±RBV therapy in a broad<br />

spectrum of patients with GT3 infection.<br />

Disclosures:<br />

David Bernstein - Advisory Committees or Review Panels: Gilead; Consulting:<br />

Abbvie, BMS, Merck, Janssen; Grant/Research Support: Gilead, Abbvie, BMS,<br />

Merck, Janssen, Genentech; Speaking and Teaching: Abbvie, BMS, Merck,<br />

Gilead<br />

Charles S. Landis - Grant/Research Support: Gilead, Abbvie, BMS<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Anne Luetkemeyer - Grant/Research Support: Gilead, Abbvie, Pfizer, Bristol<br />

Myers Squibb, Merck<br />

Melissa Harris - Employment: Bristol-Myers Squibb<br />

Eugene S. Swenson - Employment: Bristol-Myers Squibb<br />

Peter Ackerman - Employment: Bristol-Myers, Squibb<br />

Khurram Rana - Employment: Bristol-Myers Squibb<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

The following authors have nothing to disclose: Rafia Bhore<br />

727<br />

Projected Long-Term Impact of Grazoprevir (GZR,<br />

MK-5172)/Elbasvir (EBR, MK-8742) in Treatment-Naive<br />

and Treatment-Experienced Patients with Hepatitis C<br />

Virus Genotype 1 Infection and Chronic Kidney Disease<br />

Shannon A. Ferrante, Elamin Elbasha, Wayne Greaves, Chizoba<br />

Nwankwo; Merck & Co., Inc, North Wales, PA<br />

Background: Hepatitis C virus (HCV) infection is an important<br />

cause of morbidity, liver disease-related deaths from complications<br />

of decompensated cirrhosis (DC) and hepatocellular<br />

carcinoma (HCC), and cardiovascular mortality in chronic kidney<br />

disease (CKD) patients in the United States. GZR (NS3/4A<br />

protease inhibitor) and EBR (NS5A inhibitor) were recently<br />

studied in C-SURFER, a phase 2/3 double-blind, placebo-control<br />

trial in HCV genotype 1 patients with CKD4/5. Aims: To<br />

translate short-term findings from the C-SURFER trial into longterm<br />

predictions of the impact of GZR/EBR on incidence of liver-related<br />

morbidity and mortality compared with no treatment<br />

(NoTX) and pegylated interferon plus ribavirin (peg-IFN/RBV).<br />

Methods: A computer-based mathematical model of the natural<br />

history of chronic HCV genotype 1 infection and chronic<br />

kidney and liver disease was developed to project lifetime<br />

cumulative incidence of DC, HCC, liver-transplant (LT) and<br />

liver-related death. Efficacy of GZR/EBR was obtained from<br />

C-SURFER trial. In the pre-specified primary population, the<br />

proportion of patients achieving sustained viral response 12<br />

weeks after the completion of therapy was 0.99 (0.95–1.00).<br />

Based on the results of a meta-analysis, we assumed an efficacy<br />

of 0.60 (0.47–0.71) for peg-IFN/RBV. Data on baseline<br />

characteristics of the simulated patients were obtained from<br />

the National Health and Nutrition Examination Survey. Natural<br />

history parameters were estimated from published <strong>studies</strong>.


572A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

We estimated base case values for incidence and conducted<br />

sensitivity analyses. Results: GZR/EBR was projected to reduce<br />

lifetime cumulative incidence of DC to 3.8% from 22.0% and<br />

11.0% compared with NoTX and Peg-IFN/RBV, respectively<br />

(Table 1). The incidence of HCC was projected to be 26.7%<br />

in patients who received NoTx and 11.2% with peg-IFN/<br />

RBV compared with 1.1% when GZR/EBR was used. As a<br />

result, GZR/EBR reduced liver-related mortality from 35.7%<br />

with NoTX and 14.3% with Peg-IFN/RBV to 0.3%. GZR/EBR<br />

extended life expectancy by 7.6 and 3.0 years compared with<br />

NoTx and peg-IFN/RBV, respectively. The results were sensitive<br />

to assumed patient characteristics. Conclusion: Use of GZR/<br />

EBR was projected to substantially reduce the incidence of<br />

liver-related complications and mortality in treatment-naïve and<br />

treatment-experienced patients with hepatitis C virus genotype<br />

1 infection and chronic kidney disease.<br />

Table 1. Lifetime incidence of liver complications and life expectancy<br />

by regimen<br />

Disclosures:<br />

Shannon A. Ferrante - Employment: Merck & Co, Inc<br />

Elamin Elbasha - Employment: Merck & Co., Inc; Stock Shareholder: Merck &<br />

Co., Inc<br />

Wayne Greaves - Employment: Merck<br />

Chizoba Nwankwo - Employment: Merck<br />

728<br />

Daclatasvir exposure alone does not explain HCV<br />

relapse in HIV-HCV coinfected patients receiving daclatasvir<br />

plus sofosbuvir with ritonavir-boosted darunavir in<br />

the ALLY-2 study<br />

Tushar Garimella, Yash Gandhi, Reena Wang, Peter Ackerman,<br />

Zhaohui Liu, Joseph J. Raybon, Brenda Cirincione, Frank LaCreta,<br />

Timothy Eley; Bristol-Myers Squibb Research and Development,<br />

Princeton, NJ<br />

Background: ALLY-2 evaluated daclatasvir (DCV; 30, 60 or<br />

90 mg daily, adjusted for type of antiretroviral therapy) plus<br />

sofosbuvir (SOF; 400 mg daily) for 12 or 8 weeks in HCV<br />

treatment-naive or -experienced patients coinfected with HIV-1<br />

and HCV genotypes 1- 4. Sustained virologic response at<br />

posttreatment week 12 (SVR12) was lower after 8 weeks of<br />

DCV+SOF than after 12 weeks (76% vs 97%); mostly due to<br />

relapse. Three-quarters (9/12) of all relapsers were receiving<br />

ritonavir-boosted darunavir (DRV/r). In ALLY-2, the standard<br />

DCV dose of 60 mg QD was reduced a priori to 30 mg QD<br />

when coadministered with permissible ritonavir-boosted protease<br />

inhibitors, based on results from a drug-drug interaction<br />

(DDI) study showing ≈2-fold increase in DCV exposure with<br />

boosted atazanavir (ATV/r). Subsequent data from a dedicated<br />

DDI study in healthy subjects (HS) indicate that a DCV<br />

dose-adjustment is unnecessary when used with darunavir/<br />

ritonavir (DRV/r) or lopinavir/ritonavir (LPV/r); the standard 60<br />

mg dose is now recommended. Thus, the relationship between<br />

DCV exposure and SVR12 was explored in patients receiving<br />

DRV/r in ALLY-2. Methods: Composite plasma C trough<br />

concentration<br />

values in ALLY-2 patients were derived from individual<br />

DCV C trough<br />

data at dosing Weeks 1, 2, 4, 8 and 12 (for the<br />

12-week treatment arms). Data between those who did or did<br />

not receive DRV/r and did or did not achieve SVR12 were<br />

visually compared. Population PK (PopPK)-modeled DCV exposures<br />

for ALLY-2 patients were then compared with previously<br />

observed HS DDI data. Results: SVR12 among those on DRV/r<br />

in the 12-week arms was 93% (28/30), comparable to other<br />

regimens in the study, versus 67% (14/21) in the 8-week arm.<br />

Although a trend towards lower DCV C trough<br />

was observed<br />

in patients on DRV/r who did not achieve SVR12, there was<br />

substantial overlap with those who did achieve SVR12 on<br />

DRV/r (82% [42/51] of all patients on DRV/r) or LPV/r (100%<br />

[12/12]). DCV trough concentrations for DRV/r and LPV/r<br />

were comparable in the 12- and 8-week arms, and comparable<br />

to patients receiving 60 mg DCV. PopPK data for ALLY-2<br />

patients was consistent with HS data for boosted PIs. Baseline<br />

HCV RNA was >2 million IU/mL in 15/21 DRV/r patients<br />

in the 8 week arm, including all 7 relapsers. The two DRV/r<br />

relapsers in the 12-week arms each had baseline HCV RNA<br />

>10 million IU/mL and cirrhosis. Conclusions: Relapse among<br />

patients receiving DRV/r with 30 mg DCV in ALLY-2 may be<br />

associated primarily with shorter (8-week) treatment duration<br />

and high baseline HCV RNA, and cannot be solely explained<br />

by slightly lower DCV exposures. DCV should be dosed at 60<br />

mg daily with DRV/r or LPV/r based on new DDI data.<br />

Disclosures:<br />

Tushar Garimella - Employment: Bristol Myers-Squibb; Stock Shareholder: Abbvie,<br />

Bristol Myers-Squibb<br />

Yash Gandhi - Employment: Bristol-Myers Squibb<br />

Peter Ackerman - Employment: Bristol-Myers, Squibb<br />

Brenda Cirincione - Employment: BMS<br />

Frank LaCreta - Employment: Bristol-Myers Squibb; Management Position: Bristol-Myers<br />

Squibb; Stock Shareholder: Bristol-Myers Squibb<br />

Timothy Eley - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Reena Wang, Zhaohui Liu,<br />

Joseph J. Raybon<br />

729<br />

C-EDGE Co-Infection: Impact Of 12-Week Oral Regimen<br />

Of Grazoprevir (GZR, MK-5172)/Elbasvir (EBR,<br />

MK-8742) On Patient-Reported Outcomes (PROs) In<br />

Treatment-Naïve Patients With HCV/HIV Co-infection<br />

Jean Marie Arduino, Zhiwei Jiang, Melissa Shaughnessy, Shazia<br />

Khawaja, Elisa Martinez, Heather L. Platt, Chizoba Nwankwo, T.<br />

Christopher Mast; Merck & Co., Inc., Kenilworth, NJ<br />

Background: HCV/HIV co-infection negatively impacts patients’<br />

health and well-being. Treatment with interferon (IFN) and ribavirin<br />

(RBV) further diminishes patient-reported health states.<br />

This international, open-label, single arm study evaluated the<br />

fixed-dosed combination of GZR 100 mg/EBR 50 mg once<br />

daily for 12 weeks among treatment- naïve patients with HCV<br />

GT1-,4-, or 6- co-infection with HIV. GZR/EBR was highly effective<br />

(SVR12: 95.0% (95% CI: 91.2%, 97.5%)) and generally<br />

well-tolerated. Our aim was to evaluate whether treatment<br />

with GZR/EBR altered PROs during treatment and follow-up.<br />

Method: Subjects completed the following PROs at baseline,<br />

treatment week 4 (TW4), TW12, and follow-up week 4 (FW4):<br />

SF-36v2® Health Survey, EuroQol-Visual Analogue Scale<br />

(EQ-VAS), FACIT-Fatigue Scale, HCV-specific Chronic Liver Disease<br />

Questionnaire (CLDQ-HCV) and Work Productivity and<br />

Activity Impairment due to Hepatitis C. Mean change from<br />

baseline in the PRO scores, with 95% CIs, were estimated.<br />

Results: Mean improvement from baseline scores (95% CIs<br />

exclude 0) at TW12 occurred in general (Bodily Pain, General<br />

Health, Vitality, Mental Component Summary (MCS), EQ-VAS)<br />

and HCV-specific HRQOL [Figure]. The largest mean improvements<br />

occurred in General Health and Vitality domains. These<br />

improvements were also reflected in mean improvements in


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 573A<br />

fatigue levels. In addition, patients reported considerably less<br />

activity impairment (


574A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

mately 10 hours (Day -1) and 2 hours (Day 1) prior to SD 100<br />

mg GZR/50 mg EBR FDC on Day 1. In Period 3, 40 mg PAN<br />

was administered QD on Days 1 to 5 followed by administration<br />

of SD 100 mg GZR/50 mg EBR FDC 2 hours after PAN<br />

administration on Day 5. There was at least a 10-day washout<br />

between periods. PK parameters were determined for GZR and<br />

EBR on Day 1 of Periods 1 and 2, and on Day 5 of Period 3.<br />

Results Coadministration of FAM 10 and 2 hours prior to FDC<br />

of GZR/EBR did not have a clinically relevant effect on SD<br />

GZR AUC0-∞, Cmax, and C24, with geometric mean ratios<br />

(GMRs) [90% confidence intervals (CIs)] of 1.10 [0.95, 1.28,<br />

0.89 [0.71, 1.11], and 1.12 [0.97, 1.30], respectively. FAM<br />

also did not have a clinically relevant effect on EBR AUC0-∞,<br />

Cmax, and C24, with GMRs [90% CIs] of 1.05 [0.92, 1.19],<br />

1.11 [0.98, 1.26], and 1.03 [0.91, 1.17], respectively. Similarly,<br />

coadministration of PAN with GZR/EBR FDC did not<br />

have a clinically relevant effect on SD GZR AUC0-∞, Cmax,<br />

and C24, with GMRs [90% CIs] of 1.12 [0.96, 1.30], 1.10<br />

[0.89, 1.37], and 1.17 [1.02, 1.34], respectively. PAN also<br />

did not have a clinically relevant effect on EBR AUC0-∞, Cmax,<br />

and C24, with GMRs [90% CIs] of 1.05 [0.93, 1.18], 1.02<br />

[0.92, 1.14], and 1.03 [0.92, 1.17], respectively. Conclusions<br />

Coadministration of GZR/EBR FDC with FAM or PAN<br />

had no clinically relevant effect on the PK of GZR or EBR. These<br />

results demonstrate that acid-reducing agents, such as H2-receptor<br />

antagonist and proton-pump inhibitors, may be coadministered<br />

with the GZR/EBR FDC in HCV-infected patients<br />

without restriction.<br />

Disclosures:<br />

Hwa-Ping Feng - Employment: Merck<br />

Pavan Vaddady - Employment: Merck & Co., Inc.<br />

Deborah L. Panebianco - Employment: Merck<br />

Luzelena Caro - Employment: Merck & Co., Inc.<br />

Joan R. Butterton - Employment: Merck Sharp & Dohme Corp.; Stock Shareholder:<br />

Merck Sharp & Dohme Corp.<br />

Marian Iwamoto - Employment: Merck Sharp & Dohme Corp.<br />

Wendy W. Yeh - Employment: Merck<br />

The following authors have nothing to disclose: Patrice Auger, Xiaobi Huang,<br />

Fang Liu, Chun Feng, Nadia Cardillo Marricco, Vanessa Levine, David Goblet,<br />

Michael Gartner, Daria Stypinski<br />

732<br />

Viral suppression with IFN-free therapies ameliorates<br />

portal hypertension in patients with hepatitis C-related<br />

cirrhosis<br />

Mattias Mandorfer, Karin Kozbial, Philipp Schwabl, Clarissa Freissmuth,<br />

Remy Schwarzer, Rafael Stern, Simona Bota, Thomas<br />

Reiberger, Albert Stättermayer, Wolfgang Sieghart, Sandra Beinhardt,<br />

Michael Trauner, Harald Hofer, Arnulf Ferlitsch, Peter Ferenci,<br />

Markus Peck-Radosavljevic; Department of Internal Medicine<br />

III, Medical University of Vienna, Vienna, Austria<br />

Background: Interferon (IFN)-free regimens are highly effective<br />

and generally well tolerated. Portal pressure, assessed by<br />

hepatic venous pressure gradient (HVPG) measurement, drives<br />

the development of liver-related events and mortality. Since a<br />

decrease in HVPG translates into an immediate clinical benefit,<br />

it is an excellent surrogate endpoint. Aim: To investigate the<br />

effect of IFN-free treatments on portal pressure in patients with<br />

hepatitis C-related cirrhosis. Methods: Our study comprises 25<br />

patients with hepatitis C-related cirrhosis and portal hypertension<br />

(HVPG ≥6mmHg) at baseline (BL) who underwent HVPG<br />

measurement after 12-24 weeks of sofosbuvir-based IFN-free<br />

therapy (FU). Concomitant medication known to have an effect<br />

on HVPG, including but not limited to beta blockers, were<br />

paused prior to the HVPG measurements. Results are reported<br />

as median[interquartile range] or mean±standard deviation.<br />

Results: The majority of patients had Child-Pugh stage (CPS)<br />

A (80%[20/25]) and compensated cirrhosis (92%[23/25]).<br />

Ascites was the reason for decompensation in both patients<br />

and none of the patients had CPS C cirrhosis. Forty-eight percent<br />

(12/25) of patients had varices (small: 67%[8/12]/large:<br />

33%[4/12]) and two patients (8%[2/25]) had a history of variceal<br />

bleeding. All patients had undetectable HCV-RNA at the<br />

time of FU HVPG measurement. Antiviral therapy significantly<br />

decreased HVPG (BL: 16[10.5] vs. FU: 13[11.5]mmHg; mean<br />

of differences: -2.4±3.15mmHg; P20% or<br />

to ≤12mmHg) was observed in 40%(6/15). Sixty-one percent<br />

(11/18) of patients achieved a clinically meaningful HVPG<br />

decrease defined as (A) in patients with CSPH but without varices,<br />

or (B) in patients with a HVPG >12mmHg and varices.<br />

Conclusions: While Afdhal and co-workers did not observe an<br />

effect of viral suppression on HVPG in their study population<br />

comprised of a higher proportion of decompensated CPS B/C<br />

patients, effective antiviral therapy led to an immediate and<br />

clinically meaningful decrease in HVPG in our study. Thus,<br />

IFN-free therapies are likely to become a cornerstone in the<br />

treatment of portal hypertension, especially in patients with<br />

compensated cirrhosis. The results will be updated to include<br />

additional patients and information on potential predictors of<br />

HVPG-response.<br />

Disclosures:<br />

Mattias Mandorfer - Consulting: Janssen; Speaking and Teaching: AbbVie, Gilead,<br />

Janssen, Boehringer Ingelheim, Bristol-Myers Squibb, Roche<br />

Simona Bota - Speaking and Teaching: Janssen Pharmaceutica, Bristol-Myers<br />

Squibb<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

Wolfgang Sieghart - Grant/Research Support: Bayer Schering Pharma, Bayer<br />

Schering Pharma, Bayer Schering Pharma, Bayer Schering Pharma; Speaking<br />

and Teaching: Bayer Schering Pharma, Bayer Schering Pharma, Bayer Schering<br />

Pharma, Bayer Schering Pharma<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Harald Hofer - Advisory Committees or Review Panels: Gilead, Abbvie; Speaking<br />

and Teaching: Janssen, BMS, Gilead, Abbvie<br />

Peter Ferenci - Advisory Committees or Review Panels: Idenix, Gilead, MSD,<br />

Janssen, Salix, AbbVie, BMS; Patent Held/Filed: Madaus Rottapharm; Speaking<br />

and Teaching: Gilead, Roche<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

The following authors have nothing to disclose: Karin Kozbial, Philipp Schwabl,<br />

Clarissa Freissmuth, Remy Schwarzer, Rafael Stern, Albert Stättermayer, Sandra<br />

Beinhardt, Arnulf Ferlitsch


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 575A<br />

733<br />

Propranolol for Prevention of Recurrence of Varices<br />

After Endoscopic Eradication<br />

Housam Helmy 1 , Hassan E. Zaghla 1 , Wael Abdel-Razek 1 , Mohammed<br />

Abbasy 1 , Hasan A. Elzohry 1 , Ashraf Y. El-Fert 2 , Enas Korayem<br />

3 , Gamal. A. Badra 1 , Imam Waked 1 ; 1 Hepatology, National<br />

Liver Institute, Menoufiya, Egypt; 2 Biochemistry, National Liver Institute,<br />

Shebeen El Kom, Egypt; 3 Radiology, National Liver Institute,<br />

Shebeen El Kom, Egypt<br />

Background and aim: Acute variceal bleeding is a life-threatening<br />

complication of portal hypertension associated with mortality.<br />

Endoscopic band ligation can eradicate varices without<br />

an effect on portal pressure. Non-selective beta-blockers are<br />

effective in reducing portal pressure and preventing primary<br />

and secondary variceal bleeding. Whether beta-blockers can<br />

prevent recurrence of varices after eradication by band ligation<br />

is not settled. The purpose of this study was to evaluate<br />

whether propranolol can prevent the recurrence of varices after<br />

endoscopic eradication. Methods: Ninety patients in whom varices<br />

were eradicated by band-ligation (70 to prevent variceal<br />

re-bleeding and 20 for primary prophylaxis) were randomized<br />

to receive for 24 months the maximum tolerable dose of propranolol<br />

(n=47) or follow-up (n=43). Patients were excluded<br />

if they had hepatocellular or extrahepatic malignancy, portal<br />

vein thrombosis, refractory ascites, hepatorenal syndrome,<br />

Child-Turcotte-Pugh (CTP) class C, advanced systemic disease,<br />

or contra-indication to propranolol. All enrolled patients were<br />

followed endoscopically every 3 months. The primary endpoint<br />

was the recurrence of varices, and varices that recurred were<br />

re-ligated. Results: Both groups were comparable regarding<br />

age (propranolol: 46.4±11.2, follow-up: 50.1±7.3, P=0.07),<br />

gender [propranolol: males: 29 (61.7%), follow-up: 27<br />

(62.8%), P=0.915] and history of variceal bleeding [propranolol:<br />

35 (74.5%), follow-up: 35 (81.4%), P=0.459). The median<br />

dose of propranolol was 40 (20-120) mg/day. Median CTP<br />

and MELD scores did not differ between groups at baseline<br />

(CTP: 5 vs. 5, P=0.272; MELD: 6 vs. 6, P=0.840) or at the<br />

end of follow-up (CTP: 6 vs. 5.5, P=0.280; MELD: 6.43 vs. 6,<br />

P=0.264). Number of patients with recurrence of varices [propranolol:<br />

31 (66%), follow-up: 35 (81.4%), P=0.098] did not<br />

differ between groups. However, time to recurrence of varices<br />

was significantly longer in the propranolol group (20.9±1.2<br />

months vs. 15.3±1.2 months, P=0.008). No severe adverse<br />

events, rebleeding or mortality occurred during the study and<br />

no patients needed drug discontinuation. Conclusion: Propranolol<br />

use safely and significantly delayed, but did not reduce,<br />

the recurrence of esophageal varices.<br />

Disclosures:<br />

Imam Waked - Advisory Committees or Review Panels: Janssen; Speaking and<br />

Teaching: Hoffman L Roche, Merck, BMS, Gilead, AbbVie<br />

The following authors have nothing to disclose: Housam Helmy, Hassan E.<br />

Zaghla, Wael Abdel-Razek, Mohammed Abbasy, Hasan A. Elzohry, Ashraf Y.<br />

El-Fert, Enas Korayem, Gamal. A. Badra<br />

734<br />

Factors Influencing The First Decompensation Of Cirrhosis<br />

With Portal Hypertension And Varices In Primary<br />

Prophylaxis With Β-Blockers<br />

Edilmar A. Alvarado 1 , Oana Pavel 1 , Alba Ardevol 1 , Miguel<br />

Martinez 1 , Natali Escajadillo 1 , Miguel Rios 1 , Maria Poca 1 , Mar<br />

Concepción 1 , Torras Xavier 1,2 , Carlos Guarner 1,2 , Càndid Villanueva<br />

1,2 ; 1 Hospital de la santa creu i sant pau, Barcelona, Spain;<br />

2 Centro de investigación Biomédica en red en el Area temática de<br />

Enfermedades Hepáticas y Digestivas, Barcelona, Spain<br />

Patients with clinically significant portal hypertension (CSPH)<br />

and varices have a higher risk of decompensation.Hepatic<br />

venous pressure gradient (HVPG) response induces a reduction<br />

in risk of variceal bleeding.Factors such as HVPG non-response,<br />

obesity or alcohol have been associated with higher risk of<br />

decompensation.However, determinants of decompensation in<br />

patients with CSPH and varices have not been clarified.The<br />

aim of this study was to investigate the main determinants of<br />

decompensation in compensated cirrhosis treated with β-blockers<br />

for primary prophylaxis. METHODS: Cirrhotic patients<br />

referred for primary prophylaxis, were included.Hepatic<br />

and systemic hemodynamic assessment was performed.Portal<br />

pressure was estimated by HVPG measured before and<br />

after i.v. administration of propranolol (0.15 mg/kg).Nadolol<br />

was administered to prevent bleeding.Parameters which<br />

may influence the development of decompensation such as<br />

acute response to β-blockers(βB), obesity, metabolic syndrome<br />

or hepatic function,were evaluated in patients without previous<br />

decompensation of cirrhosis. RESULTS: 285 patients were<br />

included. Alcohol and HCV were the most common ethiology<br />

(33%,44%) and 161(56%)/103(36%)/ 21(7%)were Child-<br />

Pugh class A/B/C.In acute β-blockers test 49% of patients had<br />

a decrease HVPG ≥10% from baseline,137(48%)patients had<br />

no previous decompensation.Among compensated patients,decompensation<br />

occurred in 89(59%)after a mean follow-up of<br />

48±38 months:73(53%)had ascites,22(16%)SBP,11(8%)hepatorenal<br />

syndrome,22(16%)variceal bleeding,37(27%)encephalopathy,26(19%)hepatocellular<br />

carcinoma and 40(29%)died.<br />

Acute response to BB was the parameter which better discriminate<br />

decompensation,that occurred in 45% of responders vs<br />

71% of non-reponders(P= 0.01) while death occurred in 19%<br />

vs 43% respectively(P< 0.01).Cox regression analysis, acute<br />

hemodynamic response(HR= 2.4, 95% CI= 1.3-4.1),Child-<br />

Pugh score(HR= 1.9, 95% CI= 13-2.6) and active alcohol<br />

intake(HR= 2.2, 95% CI= 1.1-5.2) independently predicted the<br />

development of decompensation, while obesity, metabolic syndrome<br />

and MELD did not.Acute hemodynamic response(HR=<br />

2.5, 95% CI= 1.2-5.2) and development of decompensation(HR=<br />

2.4, 95% CI= 1.0-6.9) independently predicted mortality.<br />

CONCLUSIONS: In compensated cirrhosis on primary<br />

prophylaxis with BB, acute hemodynamic response is a strong<br />

predictor of decompensation while Child-Pugh and active alcoholism<br />

also have independent prognostic value. Acute hemodynamic<br />

response and decompensation independently predict<br />

mortality.These results suggest that etiologic therapies and<br />

treatments to improve hemodynamic response can be prevent<br />

decompensation and improve survival.<br />

Disclosures:<br />

The following authors have nothing to disclose: Edilmar A. Alvarado, Oana<br />

Pavel, Alba Ardevol, Miguel Martinez, Natali Escajadillo, Miguel Rios, Maria<br />

Poca, Mar Concepción, Torras Xavier, Carlos Guarner, Càndid Villanueva


576A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

735<br />

Obesity is associated with increased mortality and need<br />

for transplant in patients with alcoholic liver disease<br />

Mohamed G. Shoreibah 2 , Donny Kakati 1 , Sudha Kodali 2 , Ashwani<br />

K. Singal 2 ; 1 Internal Medicine, University of Alabama at Birmingham,<br />

Birrmingham, AL; 2 Gastroenetrology and Hepatology,<br />

University of Alabam at Birmingham, Birmingham, AL<br />

Obesity is associated with increased mortality and need for<br />

liver transplant in patients with alcoholic liver disease Background<br />

and aims: Alcoholic liver disease (ALD) is the second<br />

most common cause of cirrhosis and the need for liver transplant.<br />

Increasing prevalence of obesity has also resulted in a<br />

growing population of obese alcoholics. Data on the impact of<br />

obesity on the natural history of ALD are scarce. Methods: We<br />

performed a retrospective study on 286 well characterized ALD<br />

patients. After excluding 28 patients without body mass index<br />

(BMI) data at presentation, 258 ALD patients were stratified<br />

into three groups based on BMI as Group A: normal weight<br />

(BMI40 g/d in women and >60<br />

g/d in men for >5 years. Results: Group A (N=97; mean age<br />

47) differed compared to overweight alcoholics (N=78; mean<br />

age 46 y) and obese alcoholics (N=83; mean age 43 y) for<br />

male gender (53% vs. 58% vs. 70%; P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 577A<br />

737<br />

Clinical and histologic correlates of the hepatic venous<br />

pressure gradient (HVPG) in patients with compensated<br />

cirrhosis due to nonalcoholic steatohepatitis (NASH)<br />

Arun J. Sanyal 1 , Zachary D. Goodman 2 , Manal F. Abdelmalek 3 ,<br />

Stephen A. Harrison 4 , Don C. Rockey 5 , Anna Mae Diehl 3 , Stephen<br />

H. Caldwell 6 , Mitchell L. Shiffman 7 , Robert P. Myers 8 , Raul E.<br />

Aguilar Schall 8 , Mani Subramanian 8 , John G. McHutchison 8 , Vlad<br />

Ratziu 9 , Nezam H. Afdhal 10 , Jaime Bosch 11 ; 1 Virginia Commonwealth<br />

University, Richmond, VA; 2 Inova Fairfax Hospital, Falls<br />

Church, VA; 3 Duke Clinical Research Institute, Durham, NC; 4 Fort<br />

Sam Houston, San Antonio Military Medical Center, San Antonio,<br />

TX; 5 Medical University of South Carolina, Charleston, SC; 6 University<br />

of Virginia, Charlottesville, VA; 7 Liver Institute of Virginia,<br />

Richmond, VA; 8 Gilead Sciences, Inc., Foster City, CA; 9 Hôpital<br />

Pitié-Salpêtrière, Paris, France; 10 Beth Israel Deaconess Medical<br />

Center and Harvard Medical School, Boston, MA; 11 University of<br />

Barcelona, Barcelona, Spain<br />

Background: Clinical complications in patients with cirrhosis<br />

are predominantly linked to the severity of portal hypertension.<br />

Our objective was to examine clinical and histologic correlates<br />

of the HVPG in patients with compensated cirrhosis due to<br />

NASH. Methods: The study included adults with compensated<br />

cirrhosis (Ishak 5 or 6) due to NASH who were enrolled in a<br />

phase 2b trial of simtuzumab, a monoclonal antibody against<br />

lysyl oxidase-like-2 (LOXL2). Liver biopsies were graded centrally<br />

according to the NAFLD Activity Score (NAS) and hepatic<br />

collagen was quantified via computer-assisted morphometry.<br />

HVPG was measured according to a standardized protocol<br />

and reviewed centrally. Serum LOXL2 (sLOXL2) was measured<br />

using an immunoassay (VIDAS® LOXL2; bioMérieux, Marcy<br />

L’Etoile, France). The associations between HVPG and age,<br />

sex, body mass index (BMI), diabetes, use of non-selective beta<br />

blockers (NSBBs), esophageal varices, laboratory variables,<br />

fibrosis stage, hepatic collagen, and NAS and its elements<br />

were determined. Stepwise-forward, logistic regression models<br />

evaluated independent predictors of clinically significant<br />

portal hypertension (CSPH, defined as an HVPG ≥10 mmHg).<br />

Results: 230 of 258 randomized patients (89.1%) with HVPG<br />

and complete histologic data were included. The median age<br />

was 56 years (IQR 51-61), 63% were female, 67% had stage<br />

6 fibrosis, 40% had a NAS ≥5, and the median hepatic collagen<br />

content was 12.5% (n=218 with complete data; IQR<br />

8.1-19.4%). The median HVPG was 12 mmHg (IQR 9-16.5)<br />

and 69% had CSPH. The HVPG was weakly correlated with<br />

hepatic collagen content (Spearman ρ=0.19; P=0.004) and<br />

moderately correlated with sLOXL2 (ρ=0.50; P


578A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Demographic data were comparable between both groups.<br />

After a mean follow-up of 17.3±13.1 months in group A, 11<br />

patients (23.4%) bled and 13 patients (27.7%) died. In group<br />

B, 6 patients bled (12.2%) and 9 patients (18.4%) died after<br />

a mean follow-up of 18.9±14.2 months. The rebleeding and<br />

survival rates analyzed by the Kaplan-Meier method with logrank<br />

test were not significantly different between both groups<br />

(P=0.09 and 0.3, respectively). Age was the independent factor<br />

of rebleeding (OR: 1.1, 95% CI: 1.0 – 1.1, P=0.03). Hepatoma<br />

(OR: 5.6, 95% CI: 2.3 – 13.3, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 579A<br />

740<br />

Hemostatic profiles in liver disease: novel evidence of<br />

mixed phenotype in critically-ill patients with cirrhosis<br />

and liver failure and implications for clinical management<br />

of hemostatic disturbance.<br />

Vishal C. Patel 1 , Arjuna Singanayagam 1 , Jelle Adelmeijer 2 , Julia<br />

Wendon 1 , William Bernal 1 , Ton Lisman 2 ; 1 Liver Intensive Therapy<br />

Unit, Kings College Hospital, London, United Kingdom; 2 Surgical<br />

Research Laboratory, University Medical Center Groningen, Groningen,<br />

Netherlands<br />

Background Hemostatic disturbance is ubiquitous in critically-ill<br />

patients with cirrhosis and liver failure but there is lack of<br />

evidence to guide management. Studies in patients with compensated<br />

cirrhosis suggests a ‘rebalanced’ or pro-thrombotic<br />

state, but very limited data exists as to that present in those with<br />

decompensated disease or acute on chronic liver failure. We<br />

performed detailed evaluation of hemostatic profiles in patients<br />

with liver failure of varying severity, comparing to those in<br />

healthy controls. Patients and Methods Hemostatic profiles<br />

were determined in patients on admission to a single centre<br />

with stable (SC, n=4: median age 55 yrs (IQR 42-76); male<br />

75%) and decompensated (DC, n=22; 58 yrs (56-63); M 59%)<br />

cirrhosis and Acute on Chronic Liver Failure (ACLF, n=5; 60<br />

yrs (43-67), M 80%) ) and Healthy Controls (HC, n=19; 38 yrs<br />

(33-41); M 47%). MELD score was 8 (7-11) in SC, 14 (10-27)<br />

in DC and 38 (23-40) in ACLF. Assays included plasma von<br />

Willebrand Factor (VWF), its regulator ADAMTS13, Thrombin<br />

generation in the presence and absence of thrombomodulin<br />

(TM) to activate protein C, and clot lysis time to test fibrinolytic<br />

capacity. Results As compared to HC, VWF was elevated in<br />

all patient groups, with levels increasing with disease severity:<br />

SC 247% [124-467], p=ns, DC 591% [389-842], P


580A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

742<br />

External validation of risk-stratification criteria in variceal<br />

bleeding (ChildC-C1, MELD19) for the optimization<br />

of the use of early TIPS<br />

Salvador Augustin 1 , Juan Abraldes 2 , Lucio Amitrano 3 , Alba<br />

Cachero 4 , Maria Anna Guardascione 3 , Jose Castellote 4 , Puneeta<br />

Tandon 2 , Joan Genescà 1 ; 1 Hospital Universitari Vall d’Hebron,<br />

Barcelona, Spain; 2 University of Alberta, Edmonton, AB, Canada;<br />

3 Ospedale Cardarelli, Naples, Italy; 4 Hospital Bellvitge, Barcelona,<br />

Spain<br />

Early-TIPS is currently considered the treatment of choice for<br />

high-risk cirrhotic patients with acute variceal bleeding (AVB).<br />

However, recent <strong>studies</strong> have shown that the criteria used to<br />

define high-risk should be refined. Several alternatives to early-TIPS<br />

criteria (Child B + active bleeding and Child C) have<br />

been proposed (Child C + creatinine >1 mg/dL, ChildCC1;<br />

and MELD ≥19). The aim of the present study was to evaluate<br />

the external validity of three classification rules (early-TIPS trial<br />

criteria, ChildCC1 and MELD19) to identify high-risk patients<br />

eligible for early-TIPS after an ABV. In the present observational,<br />

retrospective, multicentric, international study, we<br />

reviewed already collected data from 915 patients with AVB<br />

from 4 centers which have previously published <strong>studies</strong> on the<br />

issue (Naples-Italy; Alberta-Canada; Bellvitge and Vall d’Hebron-Spain).<br />

All patients were treated with current standard of<br />

care (drugs + ligation + antibiotics, with TIPS as rescue therapy).<br />

Most patients had viral (39%) or alcohol (35%) cirrhosis<br />

(12% had both). Median age was 59 (range 23-90), 64%<br />

were male. 6-week mortality in the whole cohort was 18%.<br />

Excluding patients non-eligible for TIPS (age ≥75, creatinine<br />

≥3 mg/dL, Child score ≥14, HCC beyond Milano criteria or<br />

portal thrombosis, N=673) mortality was 15%. Mortality for<br />

Child class and each classification rule for these patients are<br />

shown in the table. All 3 rules discriminated well high vs. lowrisk<br />

patients. However, mortality for Child B was significantly<br />

lower than for Child C patients, regardless of active bleeding.<br />

Within Child C, mortality was overall high but significantly<br />

different depending on baseline creatinine (23% in


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 581A<br />

The following authors have nothing to disclose: Nicole Loo, Khurram Bari, Salvador<br />

Augustin, Maria Ciarleglio, Yanhong Deng, Mario Strazzabosco<br />

744<br />

Inter-observer variability is high for the assessment of<br />

active bleeding at endoscopy in cirrhotic patients with<br />

variceal bleeding<br />

Simona Tripon 1 , Marika Rudler 1 , Maxime Mallet 1 , Christophe<br />

Bureau 2 , Dominique Thabut 1 ; 1 Hepatology ICU, AP-HP, hôpital<br />

Pitié-Salpêtrière, Paris, France; 2 Service d’Hépato-Gastroentérologie,<br />

Hôpital Purpan, Toulouse, France<br />

Background/Aim: Guidelines stipulate that an early-TIPS<br />

placement must be considered in patients with Child-Pugh B<br />

(CPB) cirrhosis and variceal bleeding (VB) with active bleeding<br />

at initial endoscopy. However, there are no strong data<br />

confirming that active bleeding is predictive of bad survival<br />

in CPB patients. Moreover, it is a subjective criterion, which<br />

inter-observer variability has never been evaluated. The aim<br />

of this study was to assess the inter-observer variability of the<br />

endoscopic finding of active bleeding in patients with cirrhosis<br />

and portal hypertension-related gastrointestinal bleeding<br />

(GIB), using video-recording of endoscopies. Methods: We<br />

prospectively included all consecutive patients with cirrhosis<br />

and upper GIB admitted in Hepatological ICU between October<br />

2014 and May 2015. Endoscopies were performed by<br />

a senior hepatologist in charge of the patient. Patients were<br />

treated according to Baveno V recommendations. Endoscopies<br />

were recorded and analyzed by 4 other senior hepatologists<br />

who were blinded to patients’ diagnosis and outcomes<br />

and to each other findings. Interobserver agreement for endoscopic<br />

findings was determined using an agreement rate and<br />

Kappa statistics. Results: Overall, 224 evaluations were performed<br />

in 56 patients with upper GIB and cirrhosis Patients<br />

characteristics: male gender: 66.1%; age 60 ± 11.9 yrs ;<br />

etiology of cirrhosis: alcohol= 52%, viral=20%, other=28%<br />

; Child-Pugh A/B/C= 12./36/52% ; MELD score 18.8 ± 7<br />

; shock at admission 23%, source of bleeding according to<br />

physician in charge of the patient: oesophageal varices 50 %,<br />

cardial varices 11 %, ulcer 5 %, portal gastropathy 5%, other<br />

29%). Interobserver agreement for assessment of the source<br />

of bleeding was excellent (kappa=0.76, CI95%: 0.54-0.88)<br />

whereas interobserver agreement for the diagnosis of active<br />

bleeding at endoscopy was poor (kappa=0.55, CI95%: 0.35-<br />

0.70). Early-TIPS was indicated in 6 CPB patients because of<br />

active bleeding at endoscopy. Interobserver agreement was<br />

also poor for the diagnosis of active bleeding at endoscopy in<br />

those patients (kappa=0.45, CI95%: 0.24-0.60). Conclusion:<br />

Inter-observer variability in the evaluation of active bleeding at<br />

endoscopy is very high in patients with cirrhosis and variceal<br />

bleeding. This criterion should not be used anymore to indicate<br />

early-TIPS placement.<br />

Disclosures:<br />

Marika Rudler - Speaking and Teaching: Gilead, Mayoly<br />

Christophe Bureau - Grant/Research Support: Gore; Speaking and Teaching:<br />

Gore<br />

The following authors have nothing to disclose: Simona Tripon, Maxime Mallet,<br />

Dominique Thabut<br />

745<br />

Prevention of Recurrent Variceal Hemorrhage: Individual<br />

Patient Data (IPD) Meta-Analyses Of Outcomes With<br />

Current Therapy Based On Severity Of Cirrhosis<br />

Agustin Albillos 1 , Javier M. González 1 , David Arroyo-Manzano 2 ,<br />

Javier Zamora 2 , Irfan Ahmad 3 , Joaquin de la Peña 4 , Juan Carlos<br />

Garcia-Pagan 5 , Gin-Ho Lo 6 , Shiv K. Sarin 7 , Jaime Bosch 5 , Juan<br />

Abraldes 8 , Guadalupe Garcia-Tsao 9 ; 1 Hospital Universitario<br />

Ramon y Cajal, IRYCIS, CIBEREHD, Madrid, Spain; 2 Hospital<br />

Universitario Ramon y Cajal, IRYCIS, CIBERESP, Madrid, Spain;<br />

3 Sheikh Zayed Medical College/Hospital, Rahim Yar Khan, Pakistan;<br />

4 Hospital Universitario Marqués de Valdecilla, Santander,<br />

Spain; 5 Hospital Clinic, IDIBAPS, CIBEREHD, University of Barcelona,<br />

Barcelona, Spain; 6 E-DA Hospital, Kaohsiung, Taiwan;<br />

7 Institute of Liver and Biliary Sciences, New Delhi, India; 8 University<br />

of Alberta, Alberta, AB, Canada; 9 Yale University School of<br />

Medicine, New Haven, CT<br />

Current standard of care to prevent variceal rebleeding, i.e.<br />

endoscopic variceal ligation plus beta-blockers (EVL+BB), is<br />

recommended for all patients. Since risk stratification is essential<br />

in individualizing care, we aimed to investigate whether<br />

outcomes of this therapy would differ depending on cirrhosis<br />

severity. Because data were unavailable from published <strong>studies</strong>,<br />

we performed meta-analyses using IPD from fully published<br />

RCT that compared EVL+BB vs. each therapy alone (BB or EVL)<br />

and that reported results on all-source rebleeding and mortality<br />

in patients with cirrhosis who had recovered from acute<br />

variceal hemorrhage. We performed two meta-analyses: one<br />

using IPD from 3 trials (389 patients) comparing EVL+BB vs.<br />

BB (+nitrates if tolerated) (Ahmad 2009, Garcia-Pagan 2009,<br />

Lo 2009) and the second using IPD from 4 trials (409 patients)<br />

comparing EVL+BB vs. EVL alone (Ahmad 2009, De la Peña<br />

2005, Kumar 2009 and Lo 2000). We excluded 4 trials: 2<br />

abstracts, 1 with HVPG-guided drug therapy, and 1 not reporting<br />

mortality. Multilevel logistic regression models were fitted to<br />

perform one-step IPD meta-analyses, with rebleeding and mortality<br />

as dependent variables and severity of cirrhosis (Child<br />

A vs. B/C, ascites, bilirubin, albumin) and treatment modality<br />

as independent variables. Interaction terms between treatment<br />

modality and covariates were included in the model to test if<br />

treatment effect varied between subgroups. Results: The first<br />

meta-analysis (BB control group) failed to show superiority of<br />

combination therapy vs. BB alone, with a trend for a significant<br />

benefit in Child A patients. The second meta-analysis<br />

(EVL control group) shows that combination therapy EVL+BB is<br />

superior to EVL alone as it reduces rebleeding in all patients,<br />

with a significant reduction in mortality, particularly in Child<br />

B/C pts. Conclusions: Results suggest that BB alone are equivalent<br />

to combination therapy, however these results should be<br />

taken cautiously given large confidence intervals. In contrast,<br />

results clearly show that EVL alone carries an increased risk of<br />

rebleeding and death, indicating that BB are the cornerstone<br />

of combination therapy in the prevention of recurrent variceal<br />

hemorrhage.


582A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Juan Carlos Garcia-Pagan - Consulting: Novartis; Grant/Research Support:<br />

GORE<br />

Jaime Bosch - Consulting: Falk, Gilead Science, Intercept Therapeutics, Conatus<br />

Pharmaceuticals, Exalenz, Almirall, Chiasma<br />

Guadalupe Garcia-Tsao - Advisory Committees or Review Panels: Abbvie, Fibrogen<br />

The following authors have nothing to disclose: Agustin Albillos, Javier M.<br />

González, David Arroyo-Manzano, Javier Zamora, Irfan Ahmad, Joaquin de la<br />

Peña, Gin-Ho Lo, Shiv K. Sarin, Juan Abraldes<br />

746<br />

Serum lysyl oxidase-like-2 (sLOXL2) is correlated with<br />

the hepatic venous pressure gradient (HVPG) in patients<br />

with cirrhosis due to hepatitis C<br />

Nezam H. Afdhal 1 , Gregory T. Everson 2 , José Luís Calleja 3 , Geoff<br />

McCaughan 4 , Tarik Asselah 5 , Robert P. Myers 6 , Raul E. Aguilar<br />

Schall 6 , Jill M. Denning 6 , Diana M. Brainard 6 , Mani Subramanian<br />

6 , John G. McHutchison 6 , Michael R. Charlton 7 , Edward J.<br />

Gane 8 , Xavier Forns 9 , Jaime Bosch 9 ; 1 Beth Israel Deaconess Medical<br />

Center and Harvard Medical School, Boston, MA; 2 University<br />

of Colorado Denver, Aurora, CO; 3 Hospital Puerta de Hierro,<br />

Madrid, Spain; 4 Royal Prince Alfred Hospital, University of Sydney,<br />

Sydney, NSW, Australia; 5 Service d’Hépatologie, Hôpital<br />

Beaujon, Clichy, France; 6 Gilead Sciences, Inc., Foster City,<br />

CA; 7 Intermountain Medical Center, Murray, UT; 8 University of<br />

Auckland, Auckland, New Zealand; 9 Liver Unit, Hospital Clinic,<br />

IDIBAPS and CIBEREHD, Barcelona, Spain<br />

Background: Noninvasive predictors of clinically significant<br />

portal hypertension (CSPH) are limited. Our aim was to determine<br />

the association between sLOXL2, an enzyme involved in<br />

collagen cross-linkage, and the HVPG in patients with portal<br />

hypertension due to hepatitis C virus (HCV)-related cirrhosis.<br />

Methods: The study included patients with portal hypertension<br />

due to HCV-related cirrhosis who achieved a sustained virologic<br />

response (SVR12) to a 48-week course of sofosbuvir (SOF) and<br />

ribavirin (RBV) in study GS-US-334-0125. HVPG was measured<br />

according to a standardized protocol with centralized<br />

review at baseline and the end of treatment (EOT). A hemodynamic<br />

response was defined as a reduction in HVPG to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 583A<br />

fixed effects model to assess the primary outcome (incidence<br />

of gastric variceal rebleeding) and secondary outcomes (complete<br />

obliteration of gastric varices, aggravation of esophageal<br />

varices and mortality rate). Review Manager 5.3 software program<br />

was utilized for statistical analysis. Results: The search<br />

yielded three <strong>studies</strong> involving 206 patients. The mean percentage<br />

of patients with gastric variceal rebleeding was 21%<br />

in the EIT group and 4% in the BRTO group, with a mean odds<br />

ratio of 0.13, P < 0.01 (95% CI: 0.03-0.47).The percentage of<br />

patients with complete obliteration of gastric varices was 74%<br />

in the EIT group and 95% in the BRTO group: odds ratio 9.40,<br />

P < 0.01 (95% CI: 2.32-38.02). There was no significant difference<br />

for aggravation of esophageal varices and mortality<br />

rate between the two groups. Conclusions: Balloon-occluded<br />

retrograde transvenous obliteration was more effective than<br />

endoscopic injection therapy in the eradication of GV and<br />

prevention of rebleeding.<br />

Rebleeding rate BRTO vs EIT<br />

(6.1% vs 6.5%, p=1.000) and 5-day mortality (5.3% vs 4.9%,<br />

p=1.000) in PPI group and control group. PPI group showed<br />

less 6-week rebleeding rate (12.1 % vs 22.8%, p=0.031) than<br />

control group while 6-week mortality was similar between<br />

two groups (7.6% vs 6.5% respectively, p=0.810). PPI group<br />

showed better event free survival compared to control group<br />

over control group (p=0.019 by Log Rank test). Multivariate<br />

analysis identified Child-Pugh score as a predictor for 5-day<br />

treatment failure (HR 1.648, 95%CI 1.139-2.384, p=0.008).<br />

Child-Pugh score and PPI therapy were identified as independent<br />

predictors for 6-week treatment failure (HR, 1.328, 95%CI<br />

1.096-1.610 p=0.004 and HR, 0.465, 95%CI 0.233-0.931,<br />

p=0.031). The occurrence of serious infection was similar<br />

between two groups (3.8% and 6.5%, respectively p=0.399)<br />

Conclusion: PPI therapy can offer advantage of decrease the<br />

rate of recurrent bleeding in cirrhotic patients with acute variceal<br />

bleeding. PPI could be safely used in patients with liver<br />

cirrhosis by careful selection of candidate.<br />

Cox’s univariate and multivariate analysis of predictors of 6-week<br />

treatment failure<br />

Disclosures:<br />

The following authors have nothing to disclose: Dina Ahmad, Mohammad<br />

Esmadi, Shadi Hamdeh, Kunut Kijsirichareanchai, Chijioke U. Enweluzo, Saurabh<br />

Kapur, Marco A. Olivera-Martinez<br />

749<br />

Prospective validation of the efficacy of proton pump<br />

inhibitors in acute gastroesophageal variceal bleeding<br />

in patients with cirrhosis: A single center case-control<br />

study<br />

Soo Young Park 1 , Se Young Jang 1 , Su Hyun Lee 1 , Yu Rim Lee 1 , Sun<br />

Kyung Jang 1 , Won Young Tak 1 , Young Oh Kweon 1 , Jeong Heo 2 ,<br />

Hyun Young Woo 2 , Won Lim 2 , Jung Gil Park 3 , Keun Hur 4 , Gyeonghwa<br />

Kim 4 , Yong-Hun Choi 4 ; 1 Hepatology, Kyungpook National<br />

University Hospital, Daegu, Korea (the Republic of); 2 Department<br />

of Internal Medicine, Pusan National University Hospital, Pusan,<br />

Korea (the Republic of); 3 Internal Medicine, CHA University Gumi<br />

CHA Medical Center, Gumi, Korea (the Republic of); 4 Department<br />

of Biochemistry, Kyungpook National University, Daegu, Korea<br />

(the Republic of)<br />

Background and aims: The efficacy and safety of PPI in patient s<br />

with acute variceal bleeding have not evaluated yet. Therefore,<br />

we aimed to determine the efficacy and safety of long-term<br />

proton pump inhibitors in acute gastroesophageal bleeding<br />

by reviewing prospective cohort of patients with liver cirrhosis.<br />

Patients and methods: Patients with acute variceal bleeding<br />

were prospectively evaluated for 5-day in-hospital failure,<br />

6-week failure and recurrent bleeding afterwards. A total<br />

of 353 patients with liver cirrhosis visited for acute variceal<br />

bleeding from 2008 to 2013. We enrolled 255 patients by<br />

excluding 47 patients with non-variceal bleeding and 51 with<br />

HCC with portal vein thrombosis. Patients were prospectively<br />

evaluated according to previous protocols evaluating treatment<br />

outcome in acute gastroesophageal variceal bleeding (Hepatology<br />

2015;61:1033) Five day in-hospital treatment failures<br />

were determined by Baveno V criteria and rebleeding within<br />

6 weeks were prospectively evaluated Results: No differences<br />

were observed regarding baseline characteristics of patients<br />

with PPI group (n=132) and control group (n=123). There<br />

were no significant differences in 5-day treatment failure rate<br />

Disclosures:<br />

Won Young Tak - Advisory Committees or Review Panels: Gilead Korea; Grant/<br />

Research Support: SAMIL Pharma; Speaking and Teaching: Bayer Korea<br />

Jeong Heo - Advisory Committees or Review Panels: Abbvie; Grant/Research<br />

Support: Roche<br />

The following authors have nothing to disclose: Soo Young Park, Se Young Jang,<br />

Su Hyun Lee, Yu Rim Lee, Sun Kyung Jang, Young Oh Kweon, Hyun Young Woo,<br />

Won Lim, Jung Gil Park, Keun Hur, Gyeonghwa Kim, Yong-Hun Choi<br />

750<br />

Atorvastatin and propranolol combination cause greater<br />

reduction of HVPG than propranolol alone in patients of<br />

cirrhosis with portal hypertension<br />

Saptarshi Bishnu 1 , Avik Sarkar 2 , Saswata Chatterjee 1 , Sk Mahiuddin<br />

Ahammed 1 , Kshaunish Das 3 , Kausik Das 1 , Gopal K. Dhali 3 ,<br />

Abhijit Chowdhury 1 ; 1 Department of Hepatology, School of Digestive<br />

and Liver Diseases, Institute of Postgraduate Medical Education<br />

and Research, Kolkata, India; 2 Department of GI Radiology,<br />

School of Digestive and Liver Diseases, Institute of Postgraduate<br />

Medical Education and Research, Kolkata, India; 3 Department of<br />

Gastroenterology, School of Digestive and Liver Diseases, Institute<br />

of Postgraduate Medical Education and Research, Calcutta, India<br />

Introduction:Statins can modulate portal microvascular dynamics<br />

in cirrhotic patients. We present preliminary data from a<br />

study aimed at comparing combination of propranolol (Ppnl)<br />

and atorvastatin versus Ppnl alone in reducing portal pressure<br />

in cirrhosis. Methods:In this open-label study, 23 consecutive<br />

cirrhotic patients were randomized into Arm A (atorvastatin<br />

20 mg daily with Ppnl in incremental dose, n=11) or Arm B<br />

(incremental dose Ppnl, n=12). Hepatic venous pressure gradient<br />

(HVPG) was estimated at baseline and after 30 days.<br />

Results:The 2 arms were matched with respect to etiology,<br />

Child status and baseline HVPG (Table 1). Decrease of wedged<br />

hepatic venous pressure, free hepatic venous pressure and<br />

HVPG in Arm A and Arm B after 30 days were 6.09±3.56 vs<br />

4.67±2.57 (p=0.290), 1.27±1.67 vs 1.83±2.62 (p=0.546)<br />

and 4.81±2.82 vs 2.58±1.88 mm Hg (p=0.041) respectively.<br />

Conclusion:HVPG decrease in cirrhotics treated with atorvasta-


584A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tin and Ppnl is significantly more than those treated with only<br />

Ppnl. Atorvastatin, with its pleiotropic effects, may be useful in<br />

portal hypertension in cirrhosis. Larger datasets are required<br />

for ratification.<br />

Table 1: Baseline characteristics<br />

criteria and 81 patients were excluded later because of history<br />

of EVs endoscopic treatment. Finally, 1269 patients were analyzed.<br />

The prevalence of EVs/large EVs is 43.4/25.5% in all<br />

patients and is 32.0/15.4%, 69.3/47.7%, 81.6/61.8% in<br />

child A, B, C patients, respectively. Platelet


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 585A<br />

serelaxin was well tolerated. These results support further evaluation<br />

of the effect of serelaxin on HVPG in patients with PHT.<br />

Percent reductions in PPG and PVP from baseline during serelaxin<br />

infusion<br />

All values are median values. TGT – Thromboplastin generation<br />

time done using patients adsorbed plasma with control serum and<br />

control platelets. K – kinetic time, R – Reaction time.<br />

Disclosures:<br />

The following authors have nothing to disclose: Vaibhav S. Somani, Apurva<br />

Shah, Deepak N. Amarapurkar<br />

753<br />

Serelaxin reduced portal pressure gradient and portal<br />

vein pressure in patients with cirrhosis and portal<br />

hypertension<br />

Neil J. Lachlan 1 , Neil Masson 2 , Hamish Ireland 2 , Graeme Weir 2 ,<br />

Christopher Hay 2 , Thomas Severin 3 , Neelesh Dongre 3 , Rajnish<br />

Saini 4 , Judy Pak 4 , David Carr 3 , Denise Yates 5 , John P. Iredale 6 ,<br />

Peter C. Hayes 1 , Jonathan A. Fallowfield 1,6 ; 1 Liver Unit, Royal<br />

Infirmary of Edinburgh, Edinburgh, United Kingdom; 2 Department<br />

of Radiology, Royal Infirmary of Edinburgh, Edinburgh, United<br />

Kingdom; 3 Novartis Pharma AG, Basel, Switzerland; 4 Novartis<br />

Pharmaceuticals Corporation, East Hanover, NJ; 5 Novartis Institutes<br />

for Biomedical Research, Cambridge, MA; 6 MRC Centre for<br />

Inflammation Research, University of Edinburgh, Edinburgh, United<br />

Kingdom<br />

Background: Portal hypertension (PHT) accounts for the most<br />

serious complications of cirrhosis. A reduction in the hepatic<br />

venous pressure gradient (HVPG) to 20% has<br />

been reported to improve clinical outcomes, but current drug<br />

therapy is ineffective in a significant proportion of patients and<br />

adverse effects (AEs) are common. We have recently shown<br />

that serelaxin, a recombinant form of the human peptide hormone<br />

relaxin-2, has anti-fibrotic and portal hypotensive effects<br />

in cirrhotic rats. Direct measurement of portal hemodynamics<br />

in response to vasoactive drugs can be performed in patients<br />

with a transjugular intrahepatic portosystemic shunt (TIPSS) at<br />

the time of portography assessment. In a two-part exploratory<br />

open-label study, Part B examined the safety, tolerability, and<br />

effects of serelaxin on portal and systemic hemodynamics in<br />

patients with cirrhosis, PHT and a TIPSS. Methods: Patients<br />

with alcohol-related cirrhosis, a functioning TIPSS confirmed by<br />

portography, and a portal pressure gradient (PPG = portal vein<br />

pressure (PVP) - inferior vena cava pressure (IVCP)) >5mmHg<br />

were treated with serelaxin (i.v. for 60 minutes at 80ug/kg/<br />

day then at least 60 minutes at 30ug/kg/day). PPG was measured<br />

at baseline and post infusion. PVP and blood pressure<br />

(BP) were measured every 15 and 10 minutes, respectively.<br />

Safety, including AEs, was assessed. Results: 3 male and 3<br />

female patients entered the study (mean age 42.8±7.7 years).<br />

The mean baseline PPG was 8.2mmHg (95% CI 6.1, 11.1).<br />

Following at least 120 minutes of serelaxin infusion there was<br />

a 31.3% (95% CI -66.5, 71.6) reduction in the PPG compared<br />

to baseline. During the infusion there was a progressive reduction<br />

in the PVP reaching a decrease of 25.2% (95% CI -12.7,<br />

50.3) from baseline at the 120 minute time point (Table). The<br />

reduction in PVP started at 30 minutes and continued through<br />

to the 135 minute time point. With serelaxin infusion, there<br />

were no newly occurring liver enzyme abnormalities, no clinically<br />

significant changes in BP, and no discontinuations due<br />

to AEs. Conclusions: In this small exploratory study serelaxin<br />

induced a rapid and potentially clinically significant reduction<br />

in portal pressure. Despite advanced cirrhosis in these patients,<br />

Disclosures:<br />

Thomas Severin - Employment: Novartis Pharma AG; Stock Shareholder: Novartis<br />

Pharma AG<br />

Neelesh Dongre - Employment: Novartis Pharma AG; Stock Shareholder: Novartis<br />

Pharma AG<br />

Rajnish Saini - Employment: Novartis Pharmaceutical Corporation<br />

Judy Pak - Employment: Novartis Pharmaceuticals<br />

Denise Yates - Employment: Novartis<br />

Peter C. Hayes - Advisory Committees or Review Panels: Roche, MSD, Jannsen,<br />

Gilead, ?ONO, Norgine; Grant/Research Support: Novartis; Speaking and<br />

Teaching: Roche, MSD, Pfiser, Gore, Falk, Ferring<br />

The following authors have nothing to disclose: Neil J. Lachlan, Neil Masson,<br />

Hamish Ireland, Graeme Weir, Christopher Hay, David Carr, John P. Iredale,<br />

Jonathan A. Fallowfield<br />

754<br />

Transjugular intrahepatic portosystemic shunt for portal<br />

vein thrombosis with variceal bleeding in liver cirrhosis:<br />

Outcomes and predictors in a prospective cohort study<br />

Xingshun Qi, Chuangye He, Wengang Guo, Zhanxin Yin, Jianhong<br />

Wang, Zhengyu Wang, Jing Niu, Ming Bai, Zhiping Yang,<br />

Daiming Fan, Guohong Han; Xijing Hospital of Digestive Diseases,<br />

Xijing Hospital, Fourth Military Medical University, Xi’an, China<br />

This prospective cohort study aimed to assess the risk factors<br />

associated with transjugular intrahepatic portosystemic shunt<br />

(TIPS) technical success, outcome, and prognosis in cirrhotic<br />

patients with portal vein thrombosis (PVT) and a history of variceal<br />

bleeding. Between May 2009 and April 2011, 51 cirrhotic<br />

patients with PVT who attempted TIPS procedures for the<br />

prevention of variceal rebleeding were enrolled. TIPS success<br />

rate was 84% (43/51). An increased degree of thrombosis<br />

within the portal trunk and portal vein branches was inversely<br />

associated with TIPS success. Median follow-up time was 40.07<br />

months (range: 0.02-56.87). The cumulative risk of rebleeding<br />

was significantly different between TIPS success and failure<br />

group (p=0.002). The univariate analysis also demonstrated<br />

that TIPS failure was the only significant predictor associated<br />

with rebleeding (hazard ratio [HR]=4.174, 95% confidence<br />

interval [CI]: 1.558-11.186). In TIPS success group, the cumulative<br />

rates free of shunt dysfunction at the 6th and 12th month<br />

were 79% and 76%, respectively. Absence of total superior<br />

mesenteric vein (SMV) thrombosis was the only independent<br />

predictor (HR=0.189, 95%CI: 0.047-0.755). In TIPS success<br />

group, the 1- and 3-year cumulative survival rates were 77%<br />

and 62%, respectively. Albumin level was the only independent<br />

predictor (HR=0.877, 95% CI: 0.779-0.986). Successful<br />

TIPS insertions could effectively prevent from rebleeding in<br />

cirrhotic patients with PVT and variceal bleeding. Degree of<br />

PVT and SMV thrombosis was associated with TIPS failure and<br />

shunt dysfunction, respectively.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xingshun Qi, Chuangye He,<br />

Wengang Guo, Zhanxin Yin, Jianhong Wang, Zhengyu Wang, Jing Niu, Ming<br />

Bai, Zhiping Yang, Daiming Fan, Guohong Han


586A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

755<br />

Transient Elastography and simple serological tests<br />

as screening tools for esophageal varices in cirrhotic<br />

patients<br />

Simona Bota 1 , Mattias Mandorfer 1 , Remy Schwarzer 1 , Vlad F.<br />

Iovanescu 2 , Philipp Schwabl 1 , Thomas Reiberger 1 , Michael<br />

Trauner 1 , Arnulf Ferlitsch 1 , Markus Peck-Radosavljevic 1 ; 1 Gastroenterology<br />

and Hepatology, Medical University of Vienna, Vienna,<br />

Austria; 2 Gastroenterology and Hepatology, University of Medicine<br />

and Pharmacy, Craiova, Romania<br />

Background:Diagnosis of esophageal varices (EV) in cirrhotic<br />

patients is crucial for timely initiation of prophylaxis of variceal<br />

bleeding but requires upper GI endoscopy. Recently,the Baveno<br />

VI consensus proposed the use of Transient Elastography<br />

(TE) for exclusion of EV for a value under 20kPa. Our aim<br />

was to assess the diagnostic value of serological scores (initially<br />

developed for fibrosis evaluation) and transient elastrography(TE)<br />

as screening tool for exclusion of EV. Methods:Our<br />

study included 742 consecutive cirrhotic patients evaluated in<br />

our Hemodynamic Labor undergoing upper GI endoscopy,<br />

HVPG measurements, TE and serological tests between 2007<br />

and 2014. Simple serological scores were calculated: Fibrosis<br />

Index(FI), FIB-4, Forns score and Lok score. According to the<br />

newly proposed TE quality criteria, ten valid measurements with<br />

a median liver stiffness value≥7.1kPa and IQR/median>30%<br />

were considered poorly reliable and not included into the<br />

analysis. TE cases without 10 valid measurements were also<br />

excluded. For TE was used the cut-off proposed by Baveno<br />

VI consensus und for the other test the cut-offs with the best<br />

negative predictive value. Results: The etiology of liver cirrhosis<br />

was: viral-340 patients (45.8%), alcoholic-250 (33.7%),<br />

others etiologies-152 (20.5%). The incidence of EV in different<br />

etiologies of liver cirrhosis was: viral-153/340 patients (45%),<br />

alcoholic-198/250 (79.2%), others-90/152 (59.2%). The proportion<br />

of Child-Pugh A, B and C patients in each cohort of<br />

patients was: viral etiology-60%, 30%, 10%; alcoholic etiology<br />

-31.6%, 37.6%, 30.8% and others etiology -32.9%, 44.1%,<br />

23%. For a cut-off value


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 587A<br />

757<br />

Acute kidney injury predicts mortality in cirrhotic<br />

patients with gastric variceal bleeding<br />

Yun-Cheng Hsieh, Kuei-Chuan Lee, Ping-Hsien Chen, Chien-Wei<br />

Su, Ming-Chih Hou, Han-Chieh Lin; Taipei Veterans General hospital,<br />

Taipei, Taiwan<br />

Backgrounds and Aims: Acute kidney injury (AKI) is an important<br />

complication in patients with cirrhosis. However, the role<br />

of AKI in patients with gastric variceal (GV) bleeding remained<br />

unknown. Recently, the international club of ascites (ICA) proposed<br />

the new definition of AKI in cirrhotic patients. We aim at<br />

evaluating the ICA criteria and their association with the prognosis<br />

of cirrhotic patients with GV bleeding. Methods: From<br />

January 2005 to August 2011, 103 cirrhotic patients admitted<br />

to Taipei Veterans General hospital due to acute GV bleeding<br />

were analyzed retrospectively. Results: Forty-one (39.8%)<br />

patients fulfilled the ICA criteria of AKI, 26 patients (25.2%) in<br />

stage1, 13 patients (12.6%) in stage 2 and 2 patients (1.9%)<br />

in stage3, respectively. The severity of liver disease (Child-Pugh<br />

scores or its component, and MELD scores), shock at admission,<br />

hematemesis, increased transfusion blood units, elevated<br />

CRP and the development of infection were associated with the<br />

occurrence of AKI. Among patients with AKI stage 1 and 2, 8<br />

(20.5%) progressed to a higher AKI stage. The patients with<br />

progressed AKI stages had higher child-pugh scores, lower<br />

serum sodium and higher CRP level. The overall 6-week and<br />

3-month mortality were 15.5% (n=16) and 20.4% (n=21),<br />

respectively. In multivariate analysis, the occurrence of AKI<br />

was significantly associated with a higher 6-week mortality<br />

rate (34.1% vs. 3.2%, p=0.049). AKI stages were independent<br />

predictors of 3-month survival (8.1% in patients without AKI,<br />

30.1% in stage 1, 53.3% in stage 2+3, p=0.002). Conclusions:<br />

The occurrence of AKI defined by the ICA criteria is high<br />

in cirrhotic patients with acute GV bleeding. The presence AKI<br />

was associated with 6-week mortality and the stages of AKI<br />

further predict 3-month survival.<br />

Six-week (A) and 3-month (B) mortality in patients with gastric variceal<br />

bleeding stratified by acute kidney injury (AKI)<br />

Disclosures:<br />

The following authors have nothing to disclose: Yun-Cheng Hsieh, Kuei-Chuan<br />

Lee, Ping-Hsien Chen, Chien-Wei Su, Ming-Chih Hou, Han-Chieh Lin<br />

758<br />

Argon Plasma Coagulation Versus Endoscopic Band<br />

Ligation for Management of Severe Portal Hypertensive<br />

Gastropathy<br />

Mohamed M. El-Saadany; Internal Medicine, Mansoura University,<br />

Mansoura City, Egypt<br />

INTRODUCTION: Portal hypertensive gastropathy ( PHG )<br />

occurs as a complication of cirrhotic or non-cirrhotic portal<br />

hypertension. PHG is clinically important since it may cause<br />

insidious or acute severe blood loss. It is characterized<br />

endoscopically by presence of gastric mucosa abnormality<br />

described as mosaic-like pattern of snake skin with or without<br />

red spots. Argon Plasma Coagulation ( APC ) has been used<br />

as the first therapeutic endoscopic treatment of severe PHG.<br />

AIM OF STUDY: In this work, Endoscopic Band Ligation ( EBL<br />

) of affected gastric mucosa has been evaluated as a new<br />

modality for the treatment of bleeding PHG in comparison with<br />

the commonly used APC. PATIENTS and METHOD: This is a<br />

single center, prospective and randomized study included 18<br />

patients with post-HCV cirrhosis (100%) presented with upper<br />

GI bleeding admitted at Mansoura University Emergency Hospital<br />

form Jan.1 st 2013 to Feb.28 th 2014 and were followed<br />

up for 6 months after complete eradication of PHG. The 18<br />

patients were randomized 2:1 into group 1 (APC) including 12<br />

patients: 7 males and 5 females with age (meana± SD) 58.33<br />

±9.61years and group 2 (EBL) including 6 patients 1 male and<br />

5 females with age (mean± SD) 62.67±4.55 years. RESULTS:<br />

The efficacy to eradicate PHG was 83.3% in APC group and<br />

100% in EBL group (ρ. > 0.05). The number of endoscopic sessions<br />

needed to eradicate PHG was significantly (ρ. 0.03) less<br />

in EBL group (3.17±0.75) than in APC group (4.08±0.79) and<br />

the duration of endoscopic procedures was also significantely<br />

( ρ.0.001) shorter in EBL group (275.5±35.5 sec.) than in<br />

APC group (469.75±60.22 sec.). Adverse events like duration<br />

of distention was significantly (ρ.0.001) less in EBL group<br />

(3.0 ±3.29 min.) than in APC group (25.0±12.75 min.) and<br />

duration of nausea was significantly (0.02) less in EBL group<br />

(3.0±3.29 min.) than APC group (12.0±8.0 min.). However,<br />

gastric ulcers were found in 2 patients in EBL group and none<br />

in APC group (ρ.0.03) while gastric polyps were more common<br />

in APC group (41.7%) than in EBL group (ρ.0.06). CON-<br />

CLUSION: There is no significant difference in the efficacy<br />

to ablate PHG using either APC or EBL. However, EBL could<br />

achieve early ablation of PHG with less number of endoscopic<br />

sessions and fewer adverse effects.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mohamed M. El-Saadany<br />

759<br />

The impact of esophagogastric varices on the prognosis<br />

of patients with hepatocellular carcinoma<br />

Chien-Wei Su 1,2 , Ping-Hsien Chen 3,2 , Yi-Hsiang Huang 1,4 , Teh-Ia<br />

Huo 1,5 , Ming-Chih Hou 3,2 , Han-Chieh Lin 1,2 , Jaw-Ching Wu 4,6 ;<br />

1 Division of Gastroenterology, Department of Medicine, Taipei<br />

Veterans General Hospital, Taipei, Taiwan; 2 Faculty of Medicine,<br />

School of Medicine, National Yang-Ming University, Taipei, Taiwan;<br />

3 Endoscopy Center for Diagnosis and Treatment, Taipei<br />

Veterans General Hospital, Taipei, Taiwan; 4 Institute of Clinical<br />

Medicine, School of Medicine, National Yang-Ming University,<br />

Taipei, Taiwan; 5 Institute of Pharmacology, National Yang-Ming<br />

University, Taipei, Taiwan; 6 Division of Translational Research,<br />

Department of Medical Research, Taipei Veterans General Hospital,<br />

Taipei, Taiwan<br />

Background: Esophagogastric varices (EGV) is an important<br />

prognostic factor for patients with advanced liver disease, but<br />

whether it could determine the outcomes of patients with hepatocellular<br />

carcinoma (HCC) is still obscure. Aims: To assess the<br />

impact of EGV on the prognosis of patients with HCC. Methods:<br />

We enrolled 990 treatment-naïve HCC patients who also<br />

received esophagogastroduodenoscopy at the time of HCC<br />

diagnosis from 2007 to 2012. Results: A total of 480 (48.5%)<br />

patients had EGV, including 399 patients had esophageal<br />

varices (EV) alone, 12 had gastric varices (GV) alone, and<br />

69 had both EV and GV. Compared to those without EGV,<br />

patients with EGV were younger in age, and had poorer liver


588A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

functional reserve, such as higher rates of ascites and hepatic<br />

encephalopathy, lower platelet counts, lower serum albumin<br />

levels, higher bilirubin, alanine aminotransferase, aspartate<br />

aminotransferase (AST), and alkaline phosphatase levels, and<br />

more prolonged prothrombin time. Besides, patients with EGV<br />

were more with multinodular lesions, tumor vascular invasion,<br />

advanced tumor stage, and lower rates for curative treatment.<br />

After a median follow-up of 13.1 months (25-75 percentiles<br />

3.6-36.9 months), 533 patients died. The cumulative 5-year<br />

survival rates were 24.9% and 46.4% in patients with and<br />

without EGV, respectively (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 589A<br />

majority of patients were treated for splanchnic vein thrombosis<br />

(12/20) whereas in the TAC group the majority of patients<br />

were treated for non-splanchnic venous thrombosis (12/19).<br />

The TAC group had been treated for a mean of 478 days of<br />

therapy compared to 253 days in the DOAC group (p=0.2).<br />

In the DOAC group, 9/20 (45%) were exposed to traditional<br />

anticoagulation prior to initiation of DOAC. There were 3 documented<br />

bleeding events in the TAC and 4 bleeding events<br />

in DOAC (p=0.9). There were 2 major bleeding events in the<br />

TAC group and 1 major bleeding event in the DOAC group.<br />

There was 1 moderate bleeding event in each group and 2<br />

mild bleeding events in the DOAC group. On univariate regression<br />

and Cox proportional hazard modeling there were no significant<br />

predictors of bleeding, including concurrent antiplatelet<br />

therapy. Conclusions: DOAC display similar safety characteristics<br />

compared to TAC in cirrhosis patients. We did not identify<br />

any characteristic that predicted a bleeding event while on<br />

anticoagulation. Larger, prospective <strong>studies</strong> with anticoagulant<br />

agents are needed to further define safety, pharmacokinetics<br />

and efficacy in patients with cirrhosis.<br />

Bleeding events<br />

ICH- intracranial hemorrhage<br />

Disclosures:<br />

Curtis K. Argo - Independent Contractor: Salix Pharmaceuticals<br />

Patrick Northup - Consulting: Janssen<br />

Stephen H. Caldwell - Advisory Committees or Review Panels: Vital Therapy;<br />

Grant/Research Support: Genfit, Gilead Sciences, Immuron, Hyperion, Immuron,<br />

NGM<br />

The following authors have nothing to disclose: Nicolas Intagliata, Zachary<br />

Henry, Hillary Maitland, Neeral L. Shah<br />

762<br />

Transient elastography evaluation of hepatic and spleen<br />

stiffness in patients with hepatosplenic schistosomiasis<br />

Zulane D. Veiga 1 , Cristiane A. Villela Nogueira 2 , Renata D.<br />

Perez 2 , Homero S. Fogacas 2 , Flavia F. Fernandes 1 , Gustavo<br />

Pereira 1 , João Luiz Pereira 1 , Carlos Eduardo C. Marques 2 , Marta<br />

Cavalcanti 2 ; 1 Gastroenterologia e Hepatologia, Hospital Federal<br />

de Bonsucesso, Rio de Janeiro, Brazil; 2 Hepatologia, Hospital Universitário<br />

Clementino Fraga Filho, Rio de Janeiro, Brazil<br />

Hepatosplenic schistosomiasis (HES) is characterized by portal<br />

hypertension and periportal fibrosis. Transient elastography<br />

(TE) has been used to evaluate liver and spleen stiffness in<br />

many liver diseases. To our knowledge, schistosomiasis has not<br />

been evaluated by TE so far, and its correlation with ultrassonographic<br />

(US) findings related to portal hypertension remains<br />

to be defined. Objective: To assess liver and spleen stiffness by<br />

TE in patients with HES in comparison to compensated cirrhotic<br />

patients and its relation to US findings. Methods: Sixty nine<br />

patients were prospectively included in the study: 29 had HES,<br />

23 had HCV-related liver cirrhosis with esophageal varices<br />

and 17 were healthy controls. Liver and spleen stiffness values<br />

were assessed using the FibroScan® 502 equipment (Echosens,<br />

Paris, France) by an experienced operator. Ten successful<br />

measurements were carried out for each patient. Liver and<br />

spleen stiffness values were considered adequate if the success<br />

rate were at least 60% and the interquartile range (IQR) were<br />

< 30% of the median value. TE range values are 3.5 to 75 kPa.<br />

Patients with liver stiffness (LS) values > 9.5 Kpa were classified<br />

as having significant liver stiffness and spleen stiffness (SS) =<br />

75 kPa as high spleen stiffness. Patients with HES were submitted<br />

to ultrasound with dopplerfluxometry (Doppler Hitachi<br />

EUB 5500 3,5 MHz probe) by a well-trained examinator in<br />

World Health Organization (WHO) protocol for US assessment<br />

of schistosomiasis. Values are expressed as percentage<br />

and mean±SD. Results: Patients with HES had LS significantly<br />

higher than controls and lower than cirrhotic patients: 16.3 ±<br />

19,9 vs 3,7 ± 0,7 vs 29.7±15,4 kPa (p= 0,001). SS values<br />

were comparable between HES and cirrhotic patients: 57.5<br />

±18,5 vs 54 ± 21,6 kPa (p=0,4) and were significantly higher<br />

than controls (p= 0,001). Significant LS was observed in 52%<br />

of HES patients and was associated with higher values of aminotransferases:<br />

AST 45,7 ± 21 vs 32 ± 9,6 ALT 48 ± 22 vs<br />

31,7±10,8 (P< 0,05). SS values were = 75 kPa in 43% of<br />

HES patients while 24% of cirrhotic and no controls presented<br />

this value (p=0,06). In HES patients, high SS was associated<br />

to higher values of splenic artery resistance index (p=0,003),<br />

portal vein diameter( p=0,09), portal vein congestion index<br />

(p= 0,05) and splenic diameter (p= 0,09). Conclusion: HES<br />

patients present a distinct pattern of liver and spleen elastography<br />

in comparison to cirrhotic patients, with lower frequency<br />

of significant LS and higher frequency of high SS. In HES population,<br />

elevated spleen elastography may be a surrogate of<br />

splenic artery resistance and portal vein congestion.<br />

Disclosures:<br />

The following authors have nothing to disclose: Zulane D. Veiga, Cristiane A. Villela<br />

Nogueira, Renata D. Perez, Homero S. Fogacas, Flavia F. Fernandes, Gustavo<br />

Pereira, João Luiz Pereira, Carlos Eduardo C. Marques, Marta Cavalcanti<br />

763<br />

Shunt or meso-portal bypass for portal hypertension in<br />

children<br />

Neslihan Celik, Geoffrey Bond, Kyle A. Soltys, Rakesh Sindhi,<br />

George V. Mazariegos; Hillman Center for Pediatric Transplantation,<br />

Children’s Hospital of Pittsburgh of UPMC, Pittsburgh, PA<br />

Purpose To review surgical porto-systemic or meso-portal<br />

bypass techniques and post-surgical short and long term outcomes<br />

for the children with portal hypertension (PH). Methods<br />

Children who underwent porto-systemic or meso-portal bypass<br />

surgery for PH were reviewed retrospectively in terms of underlying<br />

or concomitant pathology, patient characteristics, surgical<br />

techniques, radiologic patency and clinical outcomes. Results<br />

Between 2002 and 2014, 40 patients underwent 46 bypass<br />

surgeries consisting 26 spleno-renal shunt (SRS), 9 meso-caval<br />

shunt (MCS), 8 meso-Rex bypass (MRB), 1 MRB-SRS and<br />

1 spleno-adrenal shunt. 34 patients with long term follow up<br />

(median 2.85 year, range 0.49-13.28 years) were analyzed.<br />

Extra-hepatic portal vein obstruction (EHPVO) and variceal<br />

bleeding were the indications for surgery. Porto-systemic shunts<br />

were done when MRB was not technically feasible. There was<br />

underlying liver disease in 17 patients (Group-A) and normal<br />

liver parenchyma with EHPVO in 17 patients (Group-B). The<br />

liver pathologies in Group-A were post-liver transplantation<br />

portal vein thrombosis (n=6), idiopathic liver fibrosis (n=5),<br />

total parenteral nutrition induced liver fibrosis (n=3), biliary<br />

atresia (n=2) and primary sclerosing cholangitis (n=1). Thrombophilia<br />

(n=6) and umbilical vein catheterization (n=5) were<br />

the concomitant pathologies in Group-B with idiopathic subgroup<br />

(n=6) presenting no other abnormal finding. Median<br />

patient age was 6.65 years (range 0.65-18.55 years). Three<br />

patients had re-shunt surgery due to shunt thrombosis. Three<br />

patients underwent liver transplantation because of end stage<br />

liver disease (n=2) and porto-pulmonary hypertension (n=1) in<br />

Group-A. One syndromic patient with thrombophilia died 5


590A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

years after shunt surgery because of pneumonia in Group B.<br />

The patient outcomes and comparison between 2 groups were<br />

listed in the Table below: Conclusion Extra-hepatic<br />

portal vein obstruction with or without underlying liver disease<br />

is the major cause of portal hypertension in children. Both porto-systemic<br />

and meso-portal bypass surgeries were very successful<br />

in the control of variceal bleeding demonstrated long<br />

term radiologic patency and clinical outcomes.<br />

Conclusions: Our data show that using proposed cut-offs of<br />

LSM and platelet count to rule out the presence of clinically significant<br />

varices is problematic. A LSM


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 591A<br />

Abdullah M. Al-Osaimi - Advisory Committees or Review Panels: Abbvie, Gilead,<br />

Intercept Pharmaceuticals; Grant/Research Support: Chronic Liver Disease<br />

Foundation (CLDF), Beckman Coulter Inc., Conatus Pharmaceuticals, BioPharma<br />

Alliance, Bayer HealthCare Pharmaceuticals, Inc., Vital Therapies Inc., Roche<br />

Molecular Systems, Inc., Abbvie<br />

The following authors have nothing to disclose: Paul Chang, Daniel Baik<br />

766<br />

Comparison of Beta blockers effects on All-Cause Mortality<br />

in Gastric Varices in an Urban Population<br />

Paul Chang, Daniel Baik, James Langworthy, Jeffrey Barry, Abdullah<br />

M. Al-Osaimi; Hepatology, Temple University Hospital, New<br />

York, NY<br />

Background: Gastric variceal bleeding although less frequent<br />

than esophageal variceal bleeding are associated with higher<br />

mortality and morbidity. Bleeding rates have been reported<br />

to be 16% at one year, 36% at 3 years and 44% at 5 years.<br />

There is a relative paucity of data in regards to the best management<br />

options for these patients. Our aim was to determine<br />

the effects of beta blockers therapy on patients diagnosed with<br />

gastric varices on all-cause mortality and sentinel bleeding in<br />

patients without history of bleeding from gastric varices. Methods:<br />

A retrospective chart review was conducted on subjects<br />

identified as having gastric varices from 2011-2013. Subjects<br />

were identified by our center database (Provation® MD) query.<br />

Once subjects were identified subjects were followed by chart<br />

review (EMR) until the last known follow up, additional deaths<br />

were identified by searching the Social Security database. Initial<br />

MELD, Child-Pugh, and etiology of cirrhosis were calculated<br />

at time of diagnosis, as well as heart rate, and use of beta<br />

blockers. Follow up data included initiation of beta blockers,<br />

discontinuation of beta blockers, bleeding events, source of<br />

bleed, and death. Data was analyzed using SAS V9.4. Results:<br />

A total of 85 patients were identified with no documented<br />

bleeding from gastric varices, with a mean age of 56.54 ±<br />

9.6 years, 61 (71.8%) were male, mean MELD was 12.2 ± 10<br />

with a mean follow up period of 1.82 ± 1.52 years. A total of<br />

28 patients were on beta blockers at the time of diagnosis. An<br />

additional 19 patients were started on beta blockers therapy<br />

after diagnosis, 38 patients received no beta blockers. Figure<br />

1 shows a Kaplan Meier curve for gastric variceal bleeding<br />

free survival. In addition, having a heart rate < 80beats/min<br />

trended towards significance for overall mortality, with 11<br />

(26%) subjects with heart rate < 80 compared to 18 (43%) subjects<br />

with heart rate > 80. Conclusion: Beta blockers appear<br />

to be protective in all-cause mortality in patients with gastric<br />

varices compared to patients who were not on beta blockers.<br />

There is further evidence of this as having a heart rate


592A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

768<br />

Effect of nonselective beta-blockade on prehepatic<br />

portal hypertension evaluated in patients by combined<br />

spleen pulp and hepatic venous pressure gradient measurements<br />

Michael Sørensen 1 , Lars P. Larsen 2 , Gerda E. Villadsen 1 , Niels<br />

Kristian Aagaard 1 , Henning Grønbæk 1 , Peter Ott 1 , Hendrik Vilstrup<br />

1 , Susanne Keiding 1,3 ; 1 Dept. of Hepatology & Gastroenterology,<br />

Aarhus University Hospital, Aarhus, Denmark; 2 Dept. of<br />

Radiology, Aarhus University Hospital, Aarhus, Denmark; 3 Dept.<br />

of Nuclear Medicine & PET Center, Aarhus University Hospital,<br />

Aarhus, Denmark<br />

Prehepatic portal hypertension is a clinically important cause<br />

of esophageal and gastric varices. β-blockers have an established<br />

prophylactic role against development and bleeding<br />

from esophageal and gastric varices in intrahepatic portal<br />

hypertension but the effect on prehepatic portal hypertension<br />

is unknown. Here, we evaluated the effect of β-blockade on the<br />

pressure gradient from spleen pulp to free hepatic vein which<br />

determines the overall risk of developing varices in prehepatic<br />

portal hypertension. Material and Methods: Ten patients underwent<br />

combined measurements of the pressure in the spleen pulp<br />

and hepatic vein pressure gradient (HVPG) while on and off<br />

treatment with propranolol (n=9) or carvedilol (n=1). Results:<br />

The spleen to free hepatic vein pressure gradient decreased<br />

in eight patients, increased slightly in one patient, and was<br />

unchanged in one patient. The mean gradient was 29 mm<br />

Hg without treatment and 24 mm Hg with treatment (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 593A<br />

versus TCS in the management of gastric variceal bleeding. A<br />

meta-analysis was performed using pooled estimates with odds<br />

ratio by fixed effects model to assess the primary outcome (incidence<br />

of gastric variceal rebleeding) and secondary outcomes<br />

(development of encephalopathy and mortality rate). Review<br />

Manager 5.3 software program was utilized for statistical analysis.<br />

Results: Three <strong>studies</strong> met the inclusion criteria. A total of<br />

175 patients were included in the meta-analysis. The use of<br />

TCS for management of gastric variceal bleeding showed a<br />

statistically significant decrease in the incidence of rebleeding<br />

(OR 9.19; 95% CI: 1.96-43.05, p


594A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Simona Bota - Speaking and Teaching: Janssen Pharmaceutica, Bristol-Myers<br />

Squibb<br />

Mattias Mandorfer - Consulting: Janssen; Speaking and Teaching: AbbVie, Gilead,<br />

Janssen, Boehringer Ingelheim, Bristol-Myers Squibb, Roche<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

The following authors have nothing to disclose: Philipp Schwabl, Theresa Bucsics,<br />

Remy Schwarzer, Arnulf Ferlitsch<br />

772<br />

Does Haemodynamic or Clinical Rebound Exist in<br />

patients with Cirrhosis after Abrupt Interruption of<br />

Beta-blockers?<br />

Audrey Payance 2 , Julien Bissonette 2 , Olivier Roux 2 , Laure Elkrief 2 ,<br />

claire francoz 2 , Ouardia Nekachtali 1 , Dominique C. Valla 1 , Didier<br />

Lebrec 1 , Francois Durand 2 , Pierre-Emmanuel Rautou 1 ; 1 Paris VII<br />

University and INSERM, BEAUJON HOSPITAL, Clichy, France;<br />

2 HEPATOLOGY, BEAUJON HOSPITAL, Clichy, France<br />

Background and Aims: Beta-blockers (BBs) may be interrupted<br />

in patients with portal hypertension. The concept of rebound<br />

after abrupt interruption of BBs is based on animal <strong>studies</strong><br />

showing a transient β-adrenergic hypersensitivity state following<br />

propranolol withdrawal and on few case reports of<br />

variceal bleeding following abrupt interruption of this drug.<br />

However, no human study addressed this issue yet. Our aim<br />

was thus to determine if a rebound after abrupt interruption of<br />

BBs exists in patients with cirrhosis both on a hemodynamic<br />

and a clinical point of view. Patients and Methods: This prospective<br />

observational study included all patients with cirrhosis<br />

undergoing right heart and hepatic vein catheterization in our<br />

hemodynamic laboratory. Four groups of patients were a priori<br />

defined: “no BBs” defined as patients not receiving BBs;<br />

“≥ 1day” defined as patients having received BBs within 1<br />

day before catheterization; “2-3 days”, defined as patients<br />

having interrupted BBs 2 or 3 days before catheterization;<br />

and “≥4days” defined as patients having interrupted BBs 4<br />

days or more before catheterization. Results are presented as<br />

median (interquartile range). Results: A total of 150 patients<br />

were included (men, 76%; median age, 57 yrs; alcoholic cirrhosis,<br />

75%; large varices, 73%). Indications for right heart<br />

and hepatic vein catheterization included evaluation prior to<br />

liver transplantation (n=90) or liver surgery (n=8) and assessment<br />

of cirrhosis severity (n=52). There were 81, 44, 11, 14<br />

patients in the “no BBs”, “≥1day”, “2-3 days” and “≥4days”<br />

groups, respectively. MELD and Child-Pugh scores were not<br />

different among the groups. Hepatic venous pressure gradient<br />

was not different between patients in groups “≥ 1day” (18.5<br />

mm Hg, 15-23), “2-3 days” (17 mm Hg, 13-23) and “≥4days”<br />

(19.5 mm Hg, 17-26). Cardiac index remained low up to 3<br />

days after BBs interruption. Patients in the “≥4days” group<br />

had higher cardiac index than those in the “≥1day” group<br />

[4.6 (3.5-5.1) vs 3.4 (2.6-4.0); p=0.001)] or in the “2-3 days”<br />

group [4.6 (3.5-5.1) vs 3.1 (2.7-3.7); p=0.007], but only borderline<br />

significantly higher than patients not receiving BBs [4.6<br />

(3.5-5.1) vs 3.7 (3.0-4.5); p=0.052]. Among the 25 patients<br />

in the groups “2-3 days” and “≥4days”, median duration of<br />

BBs interruption was 4 (3-6) days. No gastrointestinal bleeding<br />

occurred during that period. Conclusions: Abrupt interruption<br />

of BBs may slightly increase cardiac index without rebound<br />

in hepatic venous pressure gradient and with no apparent<br />

increase in the risk of variceal bleeding. Thus, interruption<br />

of BBs in cirrhotic patients with portal hypertension does not<br />

require particular dosing or surveillance.<br />

Disclosures:<br />

Francois Durand - Advisory Committees or Review Panels: Astellas, Novartis,<br />

BMS; Speaking and Teaching: Gilead<br />

Pierre-Emmanuel Rautou - Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Audrey Payance, Julien Bissonette,<br />

Olivier Roux, Laure Elkrief, claire francoz, Ouardia Nekachtali, Dominique<br />

C. Valla, Didier Lebrec<br />

773<br />

High risk of misclassifying liver stiffness using 2D shearwave<br />

and transient elastography during a moderate or<br />

high calorie meal<br />

Maria Kjærgaard 1 , Maja Thiele 1 , Bjørn S. Madsen 1 , Christian<br />

Jansen 2 , Jonel Trebicka 2 , Aleksander Krag 1 ; 1 Odense University<br />

Hospital, Odense, Denmark; 2 University of Bonn, Bonn, Germany<br />

Introduction: Transient Elastography (TE) (FibroScan) and<br />

real time 2D shear wave elastography (2D-SWE) (Aixplorer)<br />

are widely used in the grading and staging of liver fibrosis.<br />

In patients with chronic viral hepatitis a meal of 600 kcal<br />

increases liver stiffness (LS) measured by TE. Thus, food intake<br />

may hamper interpretation of results and clinical decision-making.<br />

This study aimed to investigate the impact of a meal on<br />

LS measured by both TE and 2D-SWE in patients with alcoholic<br />

fibrosis, the role of the meal size and the time it takes<br />

to return to baseline. Methods: Patients with a liver biopsy<br />

or known cirrhosis participated in a two-day protocol with a<br />

250ml/625 kcal liquid meal (49% carbohydrates, 36% fat,<br />

15% protein) and 500ml/1250 kcal meal on day 1 and 2<br />

respectively. TE and 2D-SWE were measured at 0, 20, 40, 60,<br />

120 and 180 minutes. DLS are the highest increase in LS from<br />

baseline. Results: 51 patients participated: 71% male, median<br />

age 61years, 88% with alcoholic etiology, 8% HCV and 4%<br />

NASH. 14 patients showed none or minimal fibrosis (METAVIR<br />

F0-1), 9 patients moderate or severe fibrosis (F2-3) and 28<br />

patients cirrhosis (F4). On day 1 and 2, post-meal LS increased<br />

significantly in all fibrosis groups (see table). The highest DLS<br />

was measured after a median of 60 minutes (range 20-180)<br />

and the largest DLS were observed in patients with cirrhosis<br />

(43% increase). In these patients, LS increased similarly on day<br />

1 and 2, despite twice the volume and calories (P= 0.93). In<br />

patients with F0-F3 fibrosis there was a trend towards a higher<br />

DTE on day 2; F0-1 (P= 0.08) and F2-3 (P=0.07). In patients<br />

with baseline TE


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 595A<br />

2D-SWE (P=0.14). On day 2, the same was true for 72% and<br />

79% (both P18, who underwent a liver biopsy. The<br />

stage of cirrhosis was identified based on liver biopsy according<br />

to the Batts-Ludwig fibrosis model. The data was randomly<br />

separated into two datasets (training 80% and testing 20%)<br />

after excluding cancer or non-liver disease. A stepwise logistic<br />

regression model was used to select candidate predictors from<br />

the variables of routine biomarkers, demographics and behavioral<br />

risk factors in the training dataset. Then, a novel diagnosis<br />

index was developed according to the effect of each candidate<br />

predictor and their correlations. The area under the curve<br />

of receiver operating characteristic (AUROC) was applied to<br />

compare the new index to existing ones (Fibro_Q, FIB4, APRI<br />

AAR), which was also validated in the testing dataset. Results:<br />

A total of 9 routine biomarkers were used to predict cirrhosis in<br />

the new index, including: aspartate aminotransferase, alanine<br />

transaminase, alkaline phosphatase, international normalization<br />

ratio, total protein, platelet count, blood urea nitrogen,<br />

albumin, and bilirubin. The new index had significantly higher<br />

AUROC (0.83, 95%CI 0.79-0.87) than Fibro_Q (0.80, CI<br />

0.76-0.85), FIB4 (0.79, CI 0.74-0.83), APRI (0.74, CI 0.69-<br />

0.78), and AAR (0.72, CI 0.67-0.78). Similar results were<br />

775<br />

Usefulness of Shear Wave Elastography in the evaluation<br />

of liver fibrosis in patients with HCV-related liver<br />

disease: a comparison with Transient Elastography<br />

Silvia Tonello 2 , Paola Bizzotto 2 , Sara Piovesan 3 , Antonietta<br />

Romano 2 , Georgios Anastassopoulos 3 , Alfredo Alberti 3 , Marta<br />

Tonon 2 , Paolo Angeli 2,1 , Giancarlo Bombonato 1 ; 1 Department of<br />

Medicine DIMED, Unit of Medical Emergencies in Liver Transplantation,<br />

University of Padova, Padova, Italy; 2 Department of Medicine<br />

-DIMED, University of Padova, Padova, Italy; 3 Department of<br />

molecular Medicine, University of Padova, Padova, Italy<br />

Aim of the study. To evaluate the applicability and performance<br />

of Shear Wave Elastography (SWE) in the assessment of liver<br />

fibrosis in healthy volunteers and in patients with HCV-related<br />

chronic liver disease, in comparison to transient elastography<br />

(TE). Methods. 48 patients affected by HCV-related chronic<br />

hepatitis or cirrhosis (19 male and 29 female, mean age 61+<br />

11, 25 with a previous diagnosis of chronic hepatitis and 23 of<br />

compensated cirrhosis) and 8 healthy volunteers (males, mean<br />

age 30+ 6) underwent SWE with the US machine Toshiba<br />

Aplio 500 Platinum and TE with Fibroscan (Echosens), on the<br />

same day, performed by two different operators. TE calculates<br />

the velocity of propagation through the liver tissue of a low-frequency<br />

transient shear wave produced by a mechanical probe.<br />

SWE measures the velocity of propagation of shear waves<br />

generated by an US induced-radiation force impulse. Toshiba<br />

Aplio 500 Platinum provides a quantitative color-code elasticity<br />

image in a box presented on the B-mode image; the liver<br />

stiffness was obtained from the average of a circular region<br />

of interest (2 cm in diameter). TE and SWE stiffness values<br />

were expressed in kPa. A semiquantitative assessment of liver<br />

steatosis was also performed according to the hyperecogenicity<br />

and US attenuation of liver parenchyma. Results. Valid<br />

TE and SWE measurements were obtained in all patients and<br />

controls. In all healthy subjects TE and SWE measurements<br />

were within the normal range (TE


596A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

sis may lead to an underestimation of the degree of fibrosis by<br />

TE. Further <strong>studies</strong> are needed to investigate the correlation of<br />

SWE measurements with liver fibrosis and steatosis assessed<br />

with biopsy.<br />

Disclosures:<br />

Alfredo Alberti - Advisory Committees or Review Panels: Merck, roche, Gilead,<br />

Merck, roche, Gilead, Merck, roche, Gilead, Merck, roche, Gilead; Grant/<br />

Research Support: Merck, gilead, Merck, gilead, Merck, gilead, Merck, gilead;<br />

Speaking and Teaching: novartis, BMS, novartis, BMS, novartis, BMS, novartis,<br />

BMS<br />

Paolo Angeli - Advisory Committees or Review Panels: Sequana Medical<br />

The following authors have nothing to disclose: Silvia Tonello, Paola Bizzotto,<br />

Sara Piovesan, Antonietta Romano, Georgios Anastassopoulos, Marta Tonon,<br />

Giancarlo Bombonato<br />

776<br />

ELF Test Thresholds for disease stratification and prognosis<br />

in chronic liver disease.<br />

James W. Day 1 , William M. Rosenberg 1 , Julie Parkes 1,2 ; 1 Institute<br />

for Liver and Digestive Health, University College London, London,<br />

United Kingdom; 2 Public Health Sciences and Medical Statistics,<br />

University of Southampton, Southampton, United Kingdom<br />

BACKGROUND: Clinically meaningful categories of liver fibrosis<br />

have traditionally been based on histological staging of<br />

liver biopsies into mild, moderate and severe fibrosis, and<br />

cirrhosis. Categorization by disease severity is used to stratify<br />

patients and carries prognostic significance. Non-invasive<br />

blood tests for liver fibrosis generate scores that may track<br />

fibrosis across the categorical stages defined by histology. As<br />

they are continuous variables they have the potential to provide<br />

information about fibrosis and prognosis between as well as at<br />

the thresholds defined by histology. We have investigated the<br />

performance of the ELF test both at, and between the manufacturer’s<br />

thresholds and determined the associated prognosis.<br />

In addition we have identified a more specific threshold for<br />

cirrhosis. METHODS: A cohort of 1,000 patients with mixed<br />

chronic liver disease was used to evaluate the performance of<br />

the manufacturer’s recommended thresholds of ELF score. An<br />

additional threshold for detecting cirrhosis with 97% specificity<br />

was derived by analyzing the ROC performance of ELF in this<br />

cohort. These same thresholds were then used to determine the<br />

time to first incidence of a liver related event (bleeding varices,<br />

ascites, encephalopathy, hepatocellular cancer, liver transplantation<br />

or liver related death) in a subset of 460 patients from<br />

the cohort followed up for 7 years. RESULTS: Ruling-out fibrosis:<br />

Fewer than 15% of patients with liver fibrosis >S1/6 will have<br />

an ELF score 9.8 has a specificity to rule in S5,6 with only a<br />

10% false positive rate and a score >11.3 has 97% specificity<br />

for cirrhosis. An increase in the ELF score of 0.8 is associated<br />

with a clinically significant difference in histological severity<br />

of fibrosis. Prognosis: The risk of a liver related event within 5<br />

years was approximately 2%, 30% and 50% respectively for<br />

patients with ELF scores in the range 7.7-9.79, 9.8-11.29 and<br />

≥11.3 ELF respectively. CONCLUSIONS: The ELF thresholds<br />

7.7, and 9.8 stratify patients with CLD into clinically meaningful<br />

groups for disease severity and prognosis. The ELF test<br />

threshold of 11.3 improves specificity for cirrhosis and adds<br />

meaningful prognostic information. The use of a continuous<br />

variable score to assess liver fibrosis permits more refined stratification<br />

of fibrosis severity than histological staging.<br />

Disclosures:<br />

William M. Rosenberg - Advisory Committees or Review Panels: Janssen, Merk,<br />

Gilead, Merk, Gilead, GSK; Board Membership: iQur Limited, iQur Limited;<br />

Consulting: siemens; Speaking and Teaching: siemens, Roche<br />

Julie Parkes - Stock Shareholder: iQur (spouse is shareholder)<br />

The following authors have nothing to disclose: James W. Day<br />

777<br />

Changes in liver stiffness by transient elastography (TE)<br />

and serum lysyl oxidase-like-2 (sLOXL2) in patients with<br />

cirrhosis treated with ledipasvir/sofosbuvir (LDV/SOF)-<br />

based therapy<br />

Marc Bourlière 1 , Veronique Loustaud-Ratti 2 , Sophie Metivier 3 ,<br />

Vincent Leroy 4 , Armando Abergel 5 , Robert P. Myers 6 , Raul E.<br />

Aguilar Schall 6 , Robert H. Hyland 6 , Mani Subramanian 6 , John<br />

G. McHutchison 6 , Lawrence Serfaty 7 , Victor de Ledinghen 8 ;<br />

1 Hépato-Gastro-Entérologie, Hôpital Saint Joseph, Marseilles,<br />

France; 2 Recherche Clinique et de l’Innovation, CHU de Limoges<br />

and Inserm UMR 1092, Université de Limoges, Limoges, France;<br />

3 Service d’Hépato-Gastro-Entérologie, Hôpital Purpan, Toulouse,<br />

France; 4 Hépato-Gastro-Entérologie, CHU de Grenboble, Grenoble,<br />

France; 5 Médecine Digestive, CHU Estaing, Clermont-Ferrand,<br />

France; 6 Gilead Sciences, Inc., Foster City, CA; 7 Hépatologie,<br />

Hôpital Saint Antoine, Paris, France; 8 Service d’Hépato-Gastro-Entérologie<br />

et d’Oncologie Digestive, CHU de Bordeaux, Pessac,<br />

France<br />

Background: Liver stiffness measurement (LSM) by TE is routinely<br />

used to monitor HCV-related fibrosis. Our aim was to<br />

explore changes in liver stiffness and a novel fibrosis marker,<br />

sLOXL2, in cirrhotic patients with a sustained virologic response<br />

(SVR) to SOF/LDV-based therapy. Methods: The study included<br />

adults with HCV cirrhosis who achieved an SVR in the SIR-<br />

IUS trial comparing LDV/SOF+ribavirin for 12 weeks vs. LDV/<br />

SOF for 24 weeks. All patients had cirrhosis defined by liver<br />

biopsy, or LSM by TE >12.5 kPa (FibroScan®; Echosens; Paris,<br />

France), or Fibrotest >0.75 and APRI >2.0. TE was performed<br />

at baseline (BL) and at the SVR24 time point, and sLOXL2<br />

was measured by immunoassay (VIDAS® LOXL2; bioMérieux,<br />

Marcy L’Etoile, France). The estimated stage of fibrosis based<br />

on TE at BL and SVR24 was categorized as F0-2 (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 597A<br />

Estimated fibrosis stage by TE at baseline and SVR24<br />

Disclosures:<br />

Marc Bourlière - Advisory Committees or Review Panels: Schering-Plough,<br />

Bohringer inghelmein, Schering-Plough, Bohringer inghelmein, Transgene; Board<br />

Membership: Bristol-Myers Squibb, Gilead, Idenix; Consulting: Roche, Novartis,<br />

Tibotec, Abott, glaxo smith kline, Merck, Bristol-Myers Squibb, Novartis, Tibotec,<br />

Abott, glaxo smith kline; Speaking and Teaching: Gilead, Roche, Merck, Bristol-Myers<br />

Squibb<br />

Veronique Loustaud-Ratti - Board Membership: Gilead, Roche, Schering Plough<br />

MSD; Speaking and Teaching: Roche, Schering Plough MSD, Janssen, Bristol<br />

meyers squibb, gilead, Abbvie<br />

Sophie Metivier - Speaking and Teaching: Roche, BMS, Janssen, Merck, Gilead,<br />

abbvie<br />

Vincent Leroy - Board Membership: Abbvie, BMS, Gilead; Consulting: Janssen,<br />

MSD; Speaking and Teaching: Abbvie, BMS, Gilead, Janssen, MSD<br />

Armando Abergel - Consulting: gilead, msd, bms; Speaking and Teaching: abbvie<br />

Robert P. Myers - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Raul E. Aguilar Schall - Employment: Gilead Sciences, Inc.<br />

Robert H. Hyland - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Lawrence Serfaty - Board Membership: BMS, Gilead; Consulting: Merck; Speaking<br />

and Teaching: Roche, Janssen, Merck, Janssen, BMS, Gilead<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

778<br />

A Hepatic Stellate Cell Secreted Protein Signature that<br />

Has Potential for use in the Non-invasive Detection of<br />

Fibrosis and Fibrogenic Activity in Chronic Liver Diseases<br />

Daniel L. Motola 1 , Chan Zhou 1 , Samuel York 1 , Yujin Hoshida 2 ,<br />

Alan Mullen 1 , Raymond T. Chung 1 ; 1 MGH, Boston, MA; 2 Division<br />

of Liver Diseases,, Icahn School of Medicine at Mount Sinai, New<br />

York, NY<br />

Background: Emerging clinical data highlight that the degree<br />

of hepatic fibrosis correlates with clinical outcomes in patients<br />

with liver disease (Ekstedt et al Hep. 2015). However, there<br />

remains no reliable method for detection of fibrosis, other<br />

than liver biopsy, which carries potential morbidity. Furthermore,<br />

non-invasive methods such as Fibroscan do not predict<br />

early stages of fibrosis well. Accordingly, development<br />

of non-invasive techniques to detect hepatic fibrosis and its<br />

progression is critically needed. It is well-known that activated<br />

hepatic stellate cells (HSCs) deposit extracellular collagen<br />

matrix and directly contribute to fibrosis. Furthermore, <strong>studies</strong><br />

reveal direct correlation between HSC number and degree of<br />

hepatic fibrosis. Therefore, we hypothesize that HSC-derived<br />

secreted proteins can be used as functionally relevant, non-invasive<br />

biomarkers to characterize disease stage and identify<br />

high-risk patients. Methods: We performed RNA-sequencing<br />

of primary human activated HSCs to comprehensively characterize<br />

their secreted protein gene expression profile, as these<br />

proteins have potential to be detected systematically. Through<br />

bioinformatic analysis, we identified genes that are specifically<br />

expressed by activated HSCs compared to whole liver and<br />

36 other human tissues (dbGaP). Further, to identify candidates<br />

of greater biological relevance we focused on: 1. genes<br />

enriched in activated versus quiescent HSCs, 2. those directly<br />

regulated by TGFβ, defined by upregulaton of expression by<br />

TGFβ and proximity to SMAD3 sites, and 3. genes within<br />

super-enhancers (presence of H3K27ac marks 10kb upstream<br />

of transcription start site). We cross-referenced candidates to<br />

a previously published 186-gene set that predicts outcomes of<br />

patients with HCC or hepatitis-C related early-stage cirrhosis.<br />

Additionally, we compared candidates to recently published<br />

and unpublished data sets corresponding to stellate signatures.<br />

Results: We identified a total of 218 genes enriched in activated<br />

HSCs. Of these 90 are upregulated by TGFβ, of which<br />

62 have associated SMAD3 sites, suggesting direct regulation.<br />

Further, 10 candidates are found within super-enhancers, 5<br />

of which are TGFβ-regulated and contain SMAD3 sites. All 5<br />

genes are implicated in liver fibrosis and may serve as putative<br />

biomarkers. Finally, we found that our 218 HSC-specific gene<br />

signature is unique and associated with negative clinical outcomes<br />

in patients with HCC and hepatitis C-related cirrhosis.<br />

Conclusion: Taken together, this work provides a functionally<br />

relevant list of genes and proteins that have potential as serum<br />

biomarkers in the non-invasive assessment of hepatic fibrogenic<br />

activity.<br />

Disclosures:<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Daniel L. Motola, Chan Zhou,<br />

Samuel York, Yujin Hoshida, Alan Mullen<br />

779<br />

Diagnostic disparity of FIB4, APRI and AAR in predicting<br />

liver fibrosis in HDV/HBV co-infected patients<br />

Varun K. Takyar 1 , David E. Kleiner 2 , Kenneth J. Wilkins 3 , Pallavi<br />

Surana 1 , Ohad Etzion 1 , Jay H. Hoofnagle 1 , T. Jake Liang 1 , Theo<br />

Heller 1 , Christopher Koh 1 ; 1 Liver Diseases Branch, NIH, Bethesda,<br />

MD; 2 Laboratory of Pathology, NCI, Bethesda, MD; 3 Office of the<br />

Director, NIDDK, Bethesda, MD<br />

Introduction: Information on staging of liver fibrosis is essential<br />

in managing chronic hepatitis B (HBV), C (HCV) and D (HDV)<br />

virus infections. Simple, non-invasive methods for the assessing<br />

fibrosis originally designed for HCV have proven usefulness<br />

in HBV. However, their accuracy and utility have not been<br />

evaluated in HDV, which involves dual virulent pathogens and<br />

accelerated hepatic fibrosis. Aims: To evaluate the utility of<br />

well established non-invasive biomarkers of fibrosis in HDV<br />

infection. Methods: Non-invasive biomarkers of liver disease<br />

(aspartate aminotransferase (AST)/alanine aminotransferase<br />

(ALT) ratio (AAR), AST-to-platelet ratio index (APRI), Fibrosis-4<br />

(FIB-4) index) were correlated with liver histology in HCV, HBV<br />

and HDV infected subjects. Laboratory measurements were<br />

obtained within 24-hours of liver biopsy. Fibrosis scores (Ishak<br />

and Knodell) were mapped to a F0-4 equivalent scale. After<br />

normalization, Mann-Whitney U test used to assess statistical<br />

differences of area under the receiver operator curve (AUROC)<br />

between the 3 biomarkers. Results: Out of 2682 unique viral<br />

hepatitis liver biopsies,1043 index pre-treatment liver biopsies<br />

with biomarkers were analyzed. The mean age was 45.4+/-<br />

12.4 at the time of biopsy; 688 (66%) male and 646 (61.9%)<br />

Caucasian. Of these, 706 (68%) were HCV, 276 (26%) were<br />

HBV and 61 (6%) were HDV. Histologically, HDV had the<br />

greatest percentage of advanced fibrosis (29%) and necroinflammation<br />

(35%) compared to HCV (11% and 6%, respectively)<br />

and HBV (16% and 7%, respectively). In HCV, FIB-4 had<br />

the best area under the receiver operating characteristic curve


598A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(AUROC) distinguishing severe (F3-F4) from mild-to-moderate<br />

fibrosis (F0-F2); 0.89 (95% confidence interval (CI), 0.86-0.91)<br />

followed by APRI 0.86 (95%CI, 0.83-0.90) and AAR 0.63<br />

(95%CI, 0.58-0.68). Similarly, FIB-4 performed best in HBV;<br />

0.73 (95%CI, 0.66-0.80), followed by APRI 0.64 (95%CI,<br />

0.56-0.72) and AAR 0.63 (95%CI, 0.55-0.72). Biomarkers<br />

performed the worst in HDV; AAR 0.65 (95%CI, 0.49-0.81),<br />

followed by FIB-4 0.60 (95% CI, 0.42-0.78) and APRI 0.55<br />

(95% CI, 0.37-0.72). Conclusion: Similar to published data,<br />

analysis of our cohort shows that FIB-4 performs better than<br />

APRI and AAR in predicting clinically significant hepatic fibrosis<br />

in HBV and HCV infections. However, in HDV, these noninvasive<br />

fibrosis markers lack diagnostic accuracy probably due<br />

to the greater necroinflammation. Noninvasive markers should<br />

be used with caution in HDV infected patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Varun K. Takyar, David E.<br />

Kleiner, Kenneth J. Wilkins, Pallavi Surana, Ohad Etzion, Jay H. Hoofnagle, T.<br />

Jake Liang, Theo Heller, Christopher Koh<br />

780<br />

Collagen maturation measured by true formation of<br />

Type III collagen (PRO-C3) provides high diagnostic<br />

accuracy in alcoholic liver fibrosis: A prospective biopsy<br />

controlled study with 196 patients<br />

Aleksander Krag 1 , Bjørn S. Madsen 1 , Diana J. Leeming 2 , Maja<br />

Thiele 1 , Mette J. Nielsen 2 , Annette D. Fialla 1 , Ove B. Schaffalitzky<br />

de Muckadell 1 , Janne F. Hansen 1 , Linda S. Møller 1 , Sönke Detlefsen<br />

3 , Morten A. Karsdal 2 ; 1 Dept. Gastroenterology and Hepatology,<br />

Odense Universityhospital, Odense, Denmark; 2 Biomarkers &<br />

Research, Nordic Bioscience, Herlev, Denmark; 3 Dept. Pathology,<br />

Odense Universityhospital, Odense, Denmark<br />

Background In alcoholic fibrosis and early cirrhosis the majority<br />

of patients have normal liver function tests. Serological biomarkers<br />

for early detection of alcoholic fibrosis are needed to<br />

monitor and prevent development and complications of cirrhosis<br />

and to motivate persistent alcohol abstinence. During fibrogenesis,<br />

type III collagen is synthesized and the fibril formation<br />

in the extracellular matrix (ECM) is dependent on site-specific<br />

cleavage of the pro-peptide. Consequently, Type III collagen<br />

formation may represent altered collagen remodelling. PRO-C3<br />

is a new serum marker of type III collagen formation, targeted<br />

at the site where the pro-peptide of type III collagen is cleaved<br />

by ADAM proteases. In contrast, existing PIIINP assays detect<br />

total type III collagen pro-peptides. We assess the diagnostic<br />

accuracy of PRO-C3 for alcoholic liver fibrosis with liver biopsy<br />

as the gold standard. Methods: We recruited 196 at-risk<br />

patients (prior or current alcohol abuse, no known liver disease)<br />

from: (A) A high a-priori fibrosis risk group from primary<br />

referrals to outpatient liver clinics (126 patients); and (B) a low<br />

a-priori fibrosis risk group from a municipal alcohol rehabilitation<br />

centre (70 patients). Results: The spectrum of fibrosis was<br />

well covered among the 196 patients with METAVIR scores of<br />

F0/1/2/3/4 in 12/102/35/13/34 patients. Among the 34<br />

with cirrhosis, median Child-Pugh was 6 (IQR 2). Median age<br />

was 56 years (IQR 13). The prevalence of significant fibrosis<br />

(≥F2) and cirrhosis (F4) in group A were 56% and 25%, versus<br />

17% and 4% in group B. AUROC analyses of PRO-C3 to<br />

diagnose significant fibrosis (≥F2) and cirrhosis (F4) was: 1) all<br />

patients ≥F2 = 0.84 (95% CI 0.78-0.90), =F4 = 0.85 (0.77-<br />

0.92), 2) ≥F2 in high-risk group = 0.85 (0.78-0.91), =F4 in<br />

high risk group= 0.80 (0.71-0.90), 3) ≥F2 in low-risk group=<br />

0.71 (0.50-0.92), =F4 in low-risk group= 0.94 (0.84-1.00).<br />

A cut-off of 11.3 ng/ml ruled out cirrhosis with a sensitivity of<br />

97% and a NPV of 98%. Below 11.3 ng/ml only one patient<br />

had cirrhosis. Above 113.6 ng/ml only one patient did not<br />

have cirrhosis. Conclusion Collagen maturation, evidenced by<br />

specific type III collagen formation, is highly upregulated in<br />

alcoholic liver fibrosis. The diagnostic accuracy of PRO-C3 to<br />

diagnose significant fibrosis and cirrhosis in two independent<br />

populations of patients with alcohol overuse and no known<br />

liver disease is high.<br />

Diagnostic accuracy of ProC3<br />

Disclosures:<br />

Aleksander Krag - Advisory Committees or Review Panels: Norgine; Speaking<br />

and Teaching: Norgine<br />

Diana J. Leeming - Employment: Nordic Bioscience; Patent Held/Filed: Nordic<br />

Bioscience<br />

Mette J. Nielsen - Grant/Research Support: Nordic Bioscience A/S<br />

Morten A. Karsdal - Management Position: Nordic Bioscience; Stock Shareholder:<br />

Nordic Bioscience<br />

The following authors have nothing to disclose: Bjørn S. Madsen, Maja Thiele,<br />

Annette D. Fialla, Ove B. Schaffalitzky de Muckadell, Janne F. Hansen, Linda S.<br />

Møller, Sönke Detlefsen<br />

781<br />

Usefulness of ultrasound elastography in assessing liver<br />

fibrosis in patients with chronic hepatitis and the difference<br />

of liver stiffness between HBV-related chronic hepatitis<br />

and non-HBV-related chronic hepatitis.<br />

Takashi Nishimura 2,1 , Chikage Nakano 2,1 , Tomoko Aoki 3 , Kyohei<br />

Kishino 1 , Yoshihiro Shimono 1 , Kunihiro Hasegawa 1 , Ryo Takata 1 ,<br />

Kazunori Yoh 1 , Akio Ishii 1 , Tomoyuki Takashima 1 , Yoshiyuki<br />

Sakai 1 , Nobuhiro Aizawa 1 , Naoto Ikeda 1 , Hiroki Nishikawa 1 ,<br />

Yoshinori Iwata 1 , Seiichi Hirota 4 , Jiro Fujimoto 5 , Shuhei Nishiguchi<br />

1 , Hiroko Iijima 2,1 ; 1 Department of Internal Medicine, Division<br />

of Hepatobiliary and Pancreatic Disease, Hyogo College of Medicine,<br />

Nishinomiya, Japan; 2 Ultrasound Imaging Center, Hyogo<br />

College of Medicine, Nishinomiya, Japan; 3 Department of Internal<br />

Medicine, Yoka Hospital, Yabu-city, Japan; 4 Department of Surgical<br />

Pathology, Hyogo College of Medicine, Nishinomiya, Japan;<br />

5 Department of Surgery, Hepato-biliary-pancreas Surgery, Hyogo<br />

College of Medicine, Nishinomiya, Japan<br />

Background and aims: Liver biopsy remains the reference standard<br />

in assessing liver fibrosis though it is an invasive method.<br />

Many <strong>studies</strong> have shown that ultrasound elastography have<br />

good diagnostic accuracy for the assessment of liver fibrosis<br />

in various chronic hepatitis. We evaluated liver stiffness<br />

in patients with various chronic hepatitis using shear wave<br />

elastography, and found out that the difference of etiologies<br />

was associated with liver stiffness difference in patient with<br />

liver cirrhosis. We compared liver stiffness in patients with different<br />

types of chronic hepatitis using Virtual Touch Quantification<br />

(VTQ, Siemens, ACUSON S2000/3000®). Methods:<br />

1406 patients with chronic liver disease who had undergone<br />

VTQ measurements from October 2008 to November 2014,<br />

and subsequently underwent liver biopsy within 4 months after<br />

VTQ measurement were evaluated to find relationship between<br />

VTQ and other parameters including the difference of etiology.<br />

This study was approved by the hospital ethics committee in<br />

our institution. Results: Multivariate analysis for the association<br />

with VTQ demonstrated that F parameter, A parameter,<br />

etiology HBV, platelet counts, total bilirubin, serum albumin,<br />

γ-GTP, PT-INR and fasting plasma glucose were associated<br />

with VTQ(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 599A<br />

positive correlation with VTQ(P


600A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

784<br />

Predicting clinical outcomes in chronic liver disease: The<br />

ELF test is superior to histology and simple scores<br />

Katharine Irvine 1 , Leesa F. Wockner 2 , Mihir Shanker 1 , Kevin<br />

Fagan 1 , Leigh Horsfall 1 , Linda Fletcher 3 , Jacobus Ungerer 4 , Carel<br />

J. Pretorius 4 , Gregory Miller 1 , Andrew D. Clouston 1 , Guy Lampe 4 ,<br />

Elizabeth E. Powell 1,3 ; 1 Centre for Liver Disease Research, The<br />

University of Queensland, Brisbane, QLD, Australia; 2 Statistics<br />

Unit, QIMR Berghorfer Medical Research Institute, Brisbane, QLD,<br />

Australia; 3 Department of Gastroenterology and Hepatology, Princess<br />

Alexandra Hospital, Brisbane, QLD, Australia; 4 Pathology<br />

Queensland, Brisbane, QLD, Australia<br />

Background: Early identification of subjects at-risk of progressive<br />

liver disease is important because these people require<br />

specialist care and surveillance for liver cancer and decompensation.<br />

However current tools for risk stratification of chronic<br />

liver disease (CLD) subjects are limited. Aims: We aimed to<br />

determine whether the serum-based Enhanced Liver Fibrosis<br />

(ELF) test predicted liver-related clinical outcomes, or progression<br />

to advanced liver disease, and to compare the performance<br />

of ELF to liver biopsy and non-invasive algorithms in<br />

predicting clinical outcomes. Methods: 300 patients with ELF<br />

scores assayed at the time of liver biopsy were followed up<br />

for liver-related clinical outcomes (ascites, encephalopathy,<br />

variceal bleeding, liver cancer) by review of medical records<br />

and clinical data. Fibrosis was staged from liver biopsy using<br />

a modified METAVIR score, and patients were grouped into<br />

non-advanced (F0-2) and advanced fibrosis (F3-4) categories.<br />

Primary outcomes were liver related morbidity and mortality<br />

and, in the F0-2 cohort, clear evidence of progression to<br />

advanced fibrosis. Results: During a median follow-up period<br />

of 6.1 years, hepatic decompensation (n=8) or HCC (n=8)<br />

occurred in 14 of 73 (19.2%) patients with ELF score ≥9.8 (the<br />

manufacturer’s cut-off for advanced fibrosis, on average 10.2<br />

years after recruitment), and in 2 of 227 ( 3 was 0.930 [cutoff 9.2kPa, sensitivity<br />

(Se) 91.5%, specificity (Sp) 85.2%), 0.789 (Se 64.7%, Sp<br />

85.5%), 0.684 (Se 59.6%, Sp 68.2%) and 0.714 (Se 57.7%,<br />

Sp 81.7%), respectively. Similarly, the AUROCs for detecting<br />

F4 using SSI, FIB-4, AAR, APRI was 0.958 (Cutoff 11.4 kPa,<br />

Se 94.9%. Sp 85.5%), 0.864 (Se 69.7%, Sp 95.7%), 0.773<br />

(Se 72.0%, Sp 70.4%) and 0.771 (Se 56.6%, Sp 90.3%),<br />

respectively. Pairwise comparison of AUROCs showed that<br />

SSI demonstrated significantly better diagnostic performance<br />

for detecting F3-4 and F4 stage fibrosis when compared to<br />

all other clinical scoring systems presented here. Conclusions:<br />

Our data indicates that SSI is a valuable diagnostic tool for<br />

the diagnosis of severe fibrosis (F3-4) and liver cirrhosis with<br />

AUROCs greater than 0.9 that demonstrated higher accuracy<br />

than several routinely used clinical scoring systems.<br />

Disclosures:<br />

Eugene R. Schiff - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead, Merck, Janssen, Salix Pharmaceutical, Pfizer, Arrowhead; Consulting:<br />

Acorda; Grant/Research Support: Bristol Myers Squibb, Abbott / AbbVee, Gilead,<br />

Merck, Conatus, Medmira, Roche, Janssen, Orasure Technologies, Discovery<br />

Life Sciences, Siemens, Beckman Coulter, Siemens<br />

The following authors have nothing to disclose: Masato Yoneda, Emmanuel<br />

Thomas, Tiffannia Grant, Seth N. Sclair


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 601A<br />

786<br />

Comparative 10-Years Prognosis of Fibrotest (FT) and<br />

Liver Stiffness Measurement (LSM) by Transient Elastography<br />

(TE, Fibroscan) in 9364 Chronic Liver Diseases<br />

(CLD) Patients (Pts)<br />

Mona Munteanu 1,2 , Yen Ngo 2 , Hugo Perazzo 3 , Pascal Lebray 3 ,<br />

Helmi Mkada 4 , Olivier Deckmyn 2 , Denis Monneret 4 , Françoise<br />

Imbert-Bismut 4 , Dominique Bonnefont-Rousselot 4 , Vincent Thibault<br />

5 , Vlad Ratziu 3 , Chantal Housset 6 , Thierry Poynard 3 ; 1 University<br />

Pierre and Marie Curie – Univ Paris 06 UMR_S 938, Paris,<br />

France; 2 Biopredictive Hepatology Resarch Unit, Paris, France;<br />

3 UPMC Liver Center APHP, Paris, France; 4 Biochemistry Department,<br />

UPMC Liver Center, Paris, France; 5 Virology Department,<br />

Hôpital Pitié-Salpêtrière, AP-HP, Paris, France; 6 Centre de Recherche<br />

Saint-Antoine & Institute of Cardiometabolism and Nutrition<br />

(ICAN), INSERM & UPMC –Univ Paris 06 UMR_S 938, Paris,<br />

France<br />

FT and LSM were validated for fibrosis and cirrhosis staging:<br />

F4.1-no complications, F4.2-varices, F4.3-decompensated.<br />

(JHepatol2014) Aim.To assess comparatively FT and LSM<br />

10-years survivals, overall (OS) and without liver-related deaths<br />

(SLD), according to the severity of fibrosis/cirrhosis in a prospective<br />

CLD cohort. Methods. Cut-offs for F4.1/F4.2/F4.3<br />

were: for FT 0.74/0.85/0.95 and LSM(kPa)12.5/20/50.<br />

Death(D) and liver transplantation(LT) rates were collected<br />

from the national and hospital registries.Statistics used the<br />

time to event Cox proportional-hazard. Results. N=9364pts<br />

with FT were included between 2003-2010: 60%males,<br />

age48.9yrs; F0F1 68%,F2F3 17% and F4 15%;etiologies:<br />

alcohol 6%, virus B 24%, C 45%, non-alcoholic fatty<br />

liver 10%,other15%; mean(max) follow-up 51(121)months.<br />

N=2724pts had applicable LSM. The OS %(95%CI)/%D as<br />

per FT decreased significantly with fibrosis stage increase<br />

(all p


602A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 603A<br />

(NPV 99%). To rule in cirrhosis, the optimal cut off values were<br />

51.4 and 27.3 for TE and 2D-SWE. Above these limits, only<br />

four patients did not have cirrhosis on biopsy (PPV 86% for TE<br />

and 84% for 2D-SWE). Conclusion: TE and 2D-SWE can be<br />

used as screening tools for alcoholic cirrhosis, with higher cutoff<br />

values than for HCV cirrhosis. Elastography reliably predicts<br />

the absence of cirrhosis. However, the risk of having cirrhosis<br />

in case of a positive test varies substantially according to the<br />

population’s disease prevalence. The positive predictive value<br />

should be taken into account when interpreting liver stiffness<br />

results.<br />

Disclosures:<br />

Aleksander Krag - Advisory Committees or Review Panels: Norgine; Speaking<br />

and Teaching: Norgine<br />

The following authors have nothing to disclose: Maja Thiele, Bjørn S. Madsen,<br />

Janne F. Hansen, Sönke Detlefsen, Linda S. Møller, Annette D. Fialla<br />

790<br />

Comparison of noninvasive fibrosis scores and association<br />

with mortality in adults with moderate to severe<br />

hepatic steatosis and NAFLD<br />

Benjamin D. Renelus 1 , Fengxia Yan 2 , Michael C. Flood 3 ; 1 Internal<br />

Medicine, Morehouse School of Medicine, Atlanta, GA; 2 Clinical<br />

Research Center, Community Health and Preventive Medicine,<br />

Morehouse School of Medicine, Atlanta, GA; 3 Division of Gastroenterology,<br />

Morehouse School of Medicine, Atlanta, GA<br />

Background: Non-alcoholic fatty liver disease (NAFLD) is highly<br />

prevalent, causing the majority of elevated serum aminotransferases.<br />

NAFLD is a disease spectrum ranging from steatosis<br />

to steatohepatitis. Steatohepatitis, which requires liver biopsy<br />

for confirmation, may progress to fibrosis, cirrhosis and liver<br />

cancer, thus contributing to overall morbidity and mortality.<br />

Knowledge regarding non-invasive long-term prognostic factors<br />

of NAFLD is limited. Aim: The purpose of this study is to evaluate<br />

accuracy of non-invasive fibrosis markers and scores with<br />

mortality among adults with moderate to severe steatosis and<br />

NAFLD. Methods: We conducted a cross-sectional analysis of<br />

data derived from the Third National Health National Health<br />

and Nutrition Examination Survey 1988-1994 (NHANES III)<br />

and ensuing follow-up mortality data through December of<br />

2006. NAFLD was defined as ultrasonographic findings consistent<br />

with moderate to severe hepatic steatosis in the absence<br />

of viral hepatitis, or excessive alcohol use. Excessive alcohol<br />

use was defined as >/=2 drinks/day in males, and >/=1<br />

drinnk/day in females. Mortality was paired with noninvasive<br />

markers of inflammation and fibrosis. These included NAFLD<br />

fibrosis score (NFS), Fibrosis-4 score (FIB-4), aspartate aminotransferase<br />

to platelet ratio index (APRI), C-reactive protein<br />

(CRP) and serum vitamin D (VDP). Results: A total of 11,490<br />

adults between the ages of 20-74 who underwent abdominal<br />

ultrasound were found to have NAFLD. Of those, 2641 were<br />

found to have moderate to severe steatosis based on ultrasonographic<br />

findings. Among those with moderate to severe<br />

steatosis there were 549 documented deaths. The vast majority<br />

of secondary causes were related to cardiovascular disease<br />

(213), followed by non-cancer liver related deaths (8) and<br />

hepatocellular carcinoma (HCC) (4). The diagnostic threshold<br />

for NAFLD among the inflammatory and fibrosis markers were<br />

as follows: VDP level 3mg/L, APRI >0.98,<br />

NFS >0.676, and FIB-4 >3.25. The area under the receiver<br />

operating characteristic curve (ROC curve) for the noninvasive<br />

markers with all-cause mortality were NFS 0.722, FIB-4 0.733,<br />

CRP 0.559, APRI 0.515, VDP 0.517. P-value were significant<br />

for all markers except for APRI and VDP. Conclusion: FIB-4 has<br />

strongest association with mortality among adults with ultrasound<br />

based moderate to severe NAFLD. FIB-4 may have some<br />

prognostic role in those with advanced NAFLD however further<br />

<strong>studies</strong> are necessary to confirm.<br />

Disclosures:<br />

The following authors have nothing to disclose: Benjamin D. Renelus, Fengxia<br />

Yan, Michael C. Flood<br />

791<br />

Effect of inflammation on liver stiffness in active autoimmune<br />

liver disease: A simultaneous biopsy-controlled<br />

study using Acoustic Radiation Force Impulse (ARFI)<br />

elastography<br />

David I. Sherman 1 , Waleed Fateen 1 , Minal J Sangwaiya 2 , Paul<br />

Tadrous 3 , Philip J. Shorvon 2 ; 1 Gastroenterology, Central MIddlesex<br />

Hospital, London North West Healthcare NHS Trust, London,<br />

United Kingdom; 2 Radiology, Central Middlesex Hospital, London<br />

North West Healthcare NHS Trust, London, United Kingdom;<br />

3 Cellular Pathology, Northwick Park Hospital, London North West<br />

Healthcare NHS Trust, London, United Kingdom<br />

Introduction ARFI elastography (virtual touch quantification,<br />

VTq) is a well validated technique for non-invasive assessment<br />

of liver fibrosis in viral hepatitis. The interpretation of<br />

shear velocity readings in active autoimmune liver disease<br />

(AILD) may be affected by a number of factors. However, few<br />

<strong>studies</strong> have specifically examined the effect of inflammation<br />

on liver stiffness (LS) in this group. We report the results of a<br />

preliminary study in which LS and histology have been sampled<br />

simultaneously from the same region of liver tissue in a large<br />

cohort with active AILD. Patients and Methods Our local database<br />

of 101 patients with AILD (63 autoimmune hepatitis AICH<br />

+/- overlap, 38 cholestatic - PBC or PSC) was investigated. LS<br />

estimation by ARFI was performed using a standard validated<br />

protocol by a single operator. Biopsies were performed from<br />

the same region of liver using an 18G Biopince needle,<br />

immediately after LS measurement. Clinical, biochemical, ultrasonic<br />

and histopathological data were collated retrospectively.<br />

ARFI/histological variance (AHV) was defined as a difference<br />

of more than 1 Metavir or 2 Ishak stages from that predicted<br />

by ARFI, according to standard calibration. 1 Results Sixty one<br />

ARFI + liver biopsies performed at the same session were identified<br />

out of a total of 164 ARFI and 114 liver biopsies. Patients<br />

included group 1: 34 active AICH (diagnosis / flare on therapy);<br />

group 2: 7 AICH remission; and group 3: 17 cholestatic<br />

liver disease. Validation confirmed satisfactory ARFI quality:<br />

SD/mean > 0.3 in 4(6.9%), failure in 3(4.9%). Co-pathology<br />

was seen in 8(13%), mostly NAFLD. Mean ARFI shear velocities<br />

in groups 1, 2 and 3 were 2.41, 1.29 and 1.64 m/sec; AHV<br />

occurred in 44.1, 28.6 and 29.4%, respectively. AHV prevalence<br />

was 41.4% overall, with 100% in group 1 and 88%<br />

overall reflecting overestimation of fibrosis. Across all groups,<br />

Ishak necro-inflammatory grade was strongly correlated with<br />

both ARFI shear velocity (r=0.58,p


604A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: David I. Sherman, Waleed<br />

Fateen, Minal J Sangwaiya, Paul Tadrous, Philip J. Shorvon<br />

792<br />

2D Shear Wave Elastography is not Superior to Transient<br />

Elastography in Difficult-to-Scan Patients: A Comparative<br />

Feasibility Study<br />

Benjamin Staugaard 3 , Janne F. Hansen 3 , Belinda Mössner 3 ,<br />

Bjørn S. Madsen 1 , Aleksander Krag 1 , Peer B. Christensen 3 ,<br />

Maja Thiele 1,2 ; 1 Department of Gastroenterology and Hepatology,<br />

Odense University Hospital, Odense C, Denmark; 2 OPEN,<br />

Odense Patient data Exploratory Network, Odense University<br />

Hospital, Odense, Denmark; 3 Department of Infectious Diseases,<br />

Odense University Hospital, Odense, Denmark<br />

Introduction: 2D shear wave elastography (2D-SWE) is a promising<br />

new elastography technique. Some <strong>studies</strong> suggest that<br />

2D-SWE reduce the number of invalid measurements otherwise<br />

seen with transient elastography (TE), but a direct comparison<br />

between the two techniques in difficult-to-scan patients<br />

has not been performed. Methods: Single-center comparative<br />

study to evaluate the feasibility of TE (Echosens FibroScan) and<br />

2D-SWE (Supersonic Aixplorer). We included patients with<br />

hepatitic C and alcoholic liver disease with a invalid TE at<br />

their latest appointment. At inclusion, an experienced operator<br />

re-examined patients with same-day TE and 2D-SWE. TE was<br />

evaluated using standard quality criteria, while 2D-SWE quality<br />

criteria mimicked TE criteria: SD/mean should be ≥ 30%.<br />

Additionally, we judged 2D-SWE image quality subjectively:<br />

A uniform colour-coded region of interest should remain stable<br />

for 3 seconds and be at least 15 mm in diameter. Results:<br />

From 10,247 TE examinations performed between 2007-14,<br />

we identified 119 eligible patients and included 52 in the<br />

study. The median age was 57 years and 60% were male.<br />

The majority of patients were overweight (median BMI 31 kg/<br />

cm 2 ; range 19-55; BMI ≥ 30 in 54% of patients). None of the<br />

patients had ascites. The initial screening examination was a<br />

failure in 13 of patients, while 39 patients had a unreliable<br />

TE. At re-examination, TE and 2D-SWE were both feasible in<br />

33 patients (63%). Seventeen patients had a valid TE, but an<br />

invalid 2D-SWE (33%). In two patients, both TE and 2D-SWE<br />

failed (4%). Both these patients were investigated with the XL<br />

probe. The XL probe was used in a total of 31 patients. The<br />

applicability of TE was superior to 2D-SWE, as TE was valid<br />

in 96% of re-examined patients while 2D-SWE was valid in<br />

63% (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 605A<br />

William M. Rosenberg - Advisory Committees or Review Panels: Janssen, Merk,<br />

Gilead, Merk, Gilead, GSK; Board Membership: iQur Limited, iQur Limited;<br />

Consulting: siemens; Speaking and Teaching: siemens, Roche<br />

The following authors have nothing to disclose: Brian J. Hogan, James O’Beirne,<br />

David W. Patch, Dominic Yu, Ioanna Parisi, Andrew R. Hall, Ankur Srivastava,<br />

Paul M. Trembling, Sudeep Tanwar<br />

794<br />

Association of skin capsular distance with accuracy<br />

of liver stiffness measurements of the FibroScan XL<br />

probe—comparison with virtual touch quantification<br />

and the FibroScan M probe<br />

Erina Kumagai 1,3 , Keiko Korenaga 2 , Masaaki Korenaga 1,2 ,<br />

Misuzu Ueyama 1,3 , Yoshihiko Aoki 2 , Yoko Yamagiwa 1 , Masatoshi<br />

Imamura 2 , Kazumoto Murata 1,2 , Tatsuya Kanto 1,2 , Naohiko<br />

Masaki 1,2 , Sumio Watanabe 3 , Masashi Mizokami 1,2 ; 1 The<br />

Research Center for Hepatitis and Immunology, National Center<br />

for Global Health and Medicine at Kohnodai, Ichikawa, Japan;<br />

2 Department of Gastroenterology and Hepatology, Kohnodai Hospital,<br />

National Center for Global Health and Medicine, Ichikawa,<br />

Japan; 3 Department of Gastroenterology, Juntendo University<br />

School of Medicine, Bunkyo-ku, Japan<br />

Background and Aim: The FibroScan XL probe was specifically<br />

designed for obese patients. However, there is no data<br />

regarding its use in Japanese individuals, who tend to be less<br />

obese than Caucasians. This study evaluates the diagnostic<br />

performance of the XL probe, the FibroScan M probe, and the<br />

virtual touch quantification (VTQ) method in Japanese patients<br />

with chronic liver disease. Methods: A total of 664 consecutive<br />

patients with chronic liver disease who underwent ultrasonography<br />

were prospectively enrolled from June through October<br />

2014. The cause of liver disease was hepatitis C virus (HCV,<br />

n = 179), hepatitis B virus (HBV, n = 79), fatty liver (n = 264),<br />

or non-HBV non-HCV hepatitis (n = 142). Liver stiffness was<br />

measured by VTQ, the M probe, and the XL probe on the same<br />

day. Liver stiffness measurement (LSM) failure was defined as<br />

zero valid shots; unreliable examination was defined as


606A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

796<br />

Glyco-isomer of Serum Mac-2-Binding Protein is The<br />

Most Useful Biomarker for Liver Fibrosis in Patients with<br />

Chronic Hepatitis C<br />

Mina Hamano, Masafumi Naito, Toshifumi Ito; Japan Community<br />

Healthcare Organization Osaka Hospital, Osaka, Japan<br />

Background and Aim The degree of liver fibrosis is a major<br />

issue for making decision of therapeutic strategies on chronic<br />

liver diseases. Recently glyco-isomer of serum Mac-2-binding<br />

protein (M2BPGi) could be applied as a novel and useful biomarker<br />

to assess liver fibrosis. However little is known about<br />

its usefulness for diagnosis of chronic hepatitis C patients. In<br />

this study, we were trying to reveal the significance of serum<br />

M2BPGi for assessment of liver fibrosis. Method One hundred<br />

fifty patients with HCV infection were enrolled, one hundred<br />

nine patients were scheduled for the anti-viral therapy or eleven<br />

patients had received hepatectomy for hepatocellular carcinoma.<br />

Liver histology was assessed in all of the patients. We<br />

evaluated the relation between liver fibrotic degrees and the<br />

characteristics of the patients by Kruskal-Wallis test. And the<br />

correlation between the serum M2BPGi and examined liver<br />

function tests, i.e. platelet, T-Bil, AST, ALT, GGT, albumin, PT,<br />

AFP, γ-globulin and hyaluronic acid (HA). Additionally we evaluated<br />

the ability of M2BPGi by ROC analysis compared with<br />

HA and Fib4 index. The multiple logistic regression analysis<br />

was performed to evaluate the contribution of fibrosis (fibrosis<br />

stage ≥ F2). Results The patients were classified into four<br />

groups according to liver fibrosis stage, F0 (n=34), F1 (n=53),<br />

F2 (n=39), and F3-4 (n=24). The serum M2BPGi were significantly<br />

increased in relation to liver fibrosis, (F0/F1/F2/<br />

F3-4 median=0.85/1.30/2.13/2.75C.O.I., p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 607A<br />

798<br />

Machine-learned Image Analysis Models for Classifying<br />

Liver Fibrosis Stage from Magnetic Resonance Images<br />

Timothy G. St. Pierre 1 , Michael J. House 1 , Ajmal Mian 2 , Sander<br />

Bangma 3 , Gary Burgess 6 , Richard A. Standish 4 , Stephen Casey 5 ,<br />

Emma Hornsey 7 , Peter W. Angus 5 ; 1 School of Physics M013, University<br />

of Western Australia, Crawley, WA, Australia; 2 School of<br />

Computer Science and Software Engineering, The University of<br />

Western Australia, Crawley, WA, Australia; 3 Resonance Health<br />

Ltd, Claremont, WA, Australia; 4 School of Medicine, Deakin University,<br />

Waurn Ponds, VIC, Australia; 5 Liver Transplant Unit, Austin<br />

Hospital, Heidelberg, VIC, Australia; 6 Pfizer Inc, New York, NY;<br />

7 Department of Radiology, Austin Hospital, Heidelberg, VIC, Australia<br />

The aim of the study was to assess the potential of machinelearned<br />

image analysis models to classify liver fibrosis stage<br />

in patients with hepatitis C virus (HCV). Methods: Patients with<br />

liver fibrosis due to chronic HCV infection were recruited from<br />

the Victorian Liver Transplant Unit. All subjects had undergone<br />

a liver transplant for HCV cirrhosis at least 12 months prior<br />

to the screening visit, and each subject had also had a liver<br />

biopsy in the previous 6 months, with a fibrosis stage score<br />

of METAVIR F1 - F3. Twenty eight subjects were recruited and<br />

scanned with MRI at two visits (visit 1 and visit 2). Visit 2 was<br />

planned to occur between 35 and 45 days after visit 1. The<br />

distribution of Ishak stage scores on the re-examined pre-study<br />

biopsies was 0 (n=2); 1 (n=5); 2 (n=6); 3 (n=8); 4 (n=2); 5<br />

(n=3); 6 (n=2). A 1.5 T Siemens Avanto MR scanner was used<br />

to acquire axial images of the abdomen. A series of breathhold<br />

spoiled gradient echo images was acquired before and<br />

after a bolus injection of 10 mL or 0.1 mmol/kg (whichever<br />

was lower) of gadoterate (Dotarem®) contrast agent including<br />

acquisitions at 3 minutes and 10 minutes after injection.<br />

Machine learning algorithms were trained on the images from<br />

visit 1 to classify subjects into Ishak stage 2 or below and 3 or<br />

above. These algorithms extract rotation invariant features by<br />

quantizing gradients from local neighbourhoods of key locations<br />

of the images at different scales. The most relevant feature<br />

statistics were then automatically selected to train Support Vector<br />

Machine (SVM) classifiers. The potential of the algorithms to<br />

correctly classify patients was tested by training the algorithms<br />

on 27 of the 28 image sets from visit 1 and testing the resulting<br />

model on the 28th image set. This “leave-one-out” strategy was<br />

used to classify all 28 image sets from visit 1. A model trained<br />

on all 28 subjects from visit 1 was then used to classify all of the<br />

image sets obtained at visit 2 in order to assess repeatability of<br />

the test. Results: Using the leave-one-out models trained on 27<br />

of the 28 visit 1 images sets, 27, 28, and 28 out of 28 cases<br />

were correctly classified in visit 1 data for pre-, 3-mins post,<br />

and 10 mins post-contrast images. The model trained from all<br />

28 visit 1 datasets correctly classified 23 out of 27, and 26 out<br />

of 28, and 25 out of 28 cases for the pre-, 3-mins post, and<br />

10 mins post-contrast images from visit 2 (corresponding to<br />

sensitivities and specificities of 80% & 92%, 93% & 92%, and<br />

93% & 85% respectively. Conclusions: Machine-learned image<br />

analysis models have the potential to classify liver fibrosis stage<br />

from high contrast MR images of the liver in patients with HCV.<br />

Disclosures:<br />

Timothy G. St. Pierre - Consulting: Resonance Health Ltd; Patent Held/Filed:<br />

Resonance Health Ltd; Stock Shareholder: Resonance Health Ltd<br />

Michael J. House - Consulting: Resonance Health; Patent Held/Filed: Resonance<br />

Health<br />

Sander Bangma - Employment: Resonance Health Analysis Services Pty Ltd; Stock<br />

Shareholder: Resonance Health Ltd<br />

Gary Burgess - Employment: Pfizer, Conatus<br />

Peter W. Angus - Advisory Committees or Review Panels: Gilead Sciences, BMS;<br />

Grant/Research Support: Gilead sciences<br />

The following authors have nothing to disclose: Ajmal Mian, Richard A. Standish,<br />

Stephen Casey, Emma Hornsey<br />

799<br />

Real-time shear wave elastography is effective for<br />

evaluating liver fibrosis in patients with chronic liver<br />

diseases<br />

Kaori Muromachi 1 , Tsutomu Tamai 1 , Kohei Oda 1 , Sho Ijuin 1 ,<br />

Hiroka Onishi 1 , Haruka Sakae 1 , Akihiko Oshige 1 , Kotaro Kumagai<br />

1 , Seiichi Mawatari 1 , Akihiro Moriuchi 2 , Hirofumi Uto 1,3 , Akio<br />

Ido 1 ; 1 Digestive and Lifestyle Diseases, Department of Human and<br />

Environmental Sciences, Kagoshima University Graduate School of<br />

Medical and Dental Sciences, Kagoshima, Japan; 2 Department of<br />

HGF Tissue Repair and Regenerative Medicine, Kagoshima University<br />

Graduate School of Medical and Dental Sciences, Kagoshima,<br />

Japan; 3 Center of Digestive and Liver Disease, Miyazaki Medical<br />

Center, Miyazaki, Japan<br />

Objective: Since liver fibrosis is the most important risk factor<br />

for liver cancer and mortality, accurate assessment of its severity<br />

is crucial. Liver biopsy is the gold standard for evaluating<br />

liver fibrosis, but it is an invasive method. Non-invasive assessment<br />

is desirable. Real-time shear wave elastography (SWE)<br />

installed on Logiq E9 ultrasound imaging system developed<br />

by GE Healthcare (USA) is a new method for evaluating liver<br />

fibrosis, but its usefulness has not been fully elucidated. We<br />

examined the utility of SWE as a new method for evaluating<br />

liver fibrosis. Methods: Forty-two patients with biopsy-proven<br />

liver disease and eighteen patients diagnosed clinically with<br />

liver cirrhosis were enrolled in this study. We examined the<br />

diagnostic performance of SWE for identifying liver fibrosis,<br />

as well as liver fibrosis markers and scores. Results: Twenty<br />

patients were male, and the median age was 63.0 years.<br />

The etiology of liver disease was hepatitis B virus (HBV) in<br />

3 patients, hepatitis C virus (HCV) in 20, and non-B non-C<br />

(NBNC) in 37. The median platelet count was 15.0×10 4 /<br />

mL, serum ALT was 46 IU/L, hyaluronic acid was 145.9 ng/<br />

mL, type IV collagen 7S was 7.9 ng/mL, and procollagen III<br />

peptide was 0.9 U/mL. There were 30, 8, 3, and 19 patients<br />

with liver fibrosis scores of F0–1, F2, F3, and F4, respectively.<br />

Median SWE (m/s) in patients with F0–1/F2/F3/F4 fibrosis<br />

was 1.38/1.64/1.77/1.73. SWE increased with increasing<br />

severity of liver fibrosis and was significantly correlated with<br />

liver fibrosis markers (platelet count, hyaluronic acid, type IV<br />

collagen 7S, procollagen III peptide) and scores (FIB4 index,<br />

APRI). SWE was also significantly associated with liver histopathology<br />

findings (r=0.661, P


608A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

800<br />

Role of Acoustic Radiation Force Impulse (ARFI) elastography<br />

in non-invasive assessment of fibrosis in active<br />

autoimmune liver disease: A simultaneous biopsy-controlled<br />

study<br />

David I. Sherman 1 , Waleed Fateen 1 , Minal J Sangwaiya 2 , Paul<br />

Tadrous 3 , Philip J. Shorvon 2 ; 1 Gastroenterology, Central MIddlesex<br />

Hospital, London North West Healthcare NHS Trust, London,<br />

United Kingdom; 2 Radiology, Central Middlesex Hospital, London<br />

North West Healthcare NHS Trust, London, United Kingdom;<br />

3 Cellular Pathology, Northwick Park Hospital, London North West<br />

Healthcare NHS Trust, London, United Kingdom<br />

Introduction The predictive accuracy of ARFI elastography (virtual<br />

touch quantification, VTq) for the non-invasive assessment<br />

of liver fibrosis is well validated in viral hepatitis. Whilst there<br />

is evidence that inflammation can increase liver stiffness, the<br />

role of ARFI in interpreting active autoimmune liver disease is<br />

unclear. We report the results of a preliminary biopsy-controlled<br />

study in a large autoimmune cohort, in which liver stiffness<br />

(LS) and histology were sampled simultaneously from the same<br />

region of liver tissue. Patients and Methods Our local database<br />

of 101 patients with autoimmune liver disease (63 AICH +/-<br />

overlap, 38 PBC or PSC) was interrogated. LS estimation by<br />

ARFI was performed using a standard validated protocol by a<br />

single operator. Biopsies were performed from the same region<br />

of liver using an 18G Biopince needle, immediately after LS<br />

measurement. Clinical, biochemical, ultrasonic and histopathological<br />

data were collated retrospectively. Predictive accuracy<br />

variables were determined for fibrosis stage using both standard<br />

and local calibrations 1 . Results Sixty one ARFI + liver<br />

biopsies performed at the same session were identified out of a<br />

total of 164 ARFI examinations and 114 liver biopsies. Patients<br />

included group 1: 34 active AICH (diagnosis pre-therapy or<br />

flare on therapy); group 2: 7 AICH biochemical remission; and<br />

group 3: 17 cholestatic liver disease. Validation confirmed<br />

satisfactory ARFI quality: SD/mean > 0.3 in 4(6.9%), failure<br />

in 3(4.9%). Co-pathology was seen in 8(13%), mostly NAFLD.<br />

AUROC analysis showed overall predictive performance for<br />

all groups/group 1/group 3 were 54.8/41.7/56.3, respectively;<br />

for exclusion of ≥F2 69.5/44.9/81.8; for detection of<br />

F3/4 fibrosis 70.1/55.9/79.8. Using the chosen cut off point,<br />

sensitivity and NPV for F3/4 detection were 100% for both all<br />

groups and group 3. No difference was seen between performance<br />

of standard and local calibrations. Conclusion These<br />

simultaneous paired data demonstrate a reduced predictive<br />

performance for fibrosis stage using ARFI in active autoimmune<br />

liver disease, probably due to the inflammatory process. However,<br />

reasonable performance was demonstrated for both ≥F2<br />

exclusion and confirmation of F3/4 in PBC/PSC. The finding<br />

that ARFI can reliably exclude advanced fibrosis, particularly<br />

in cholestatic liver disease, is of potential practical importance.<br />

Further biopsy-controlled <strong>studies</strong> are needed to confirm the role<br />

of ARFI in this patient group. Reference D. Sherman et al. J<br />

Hepatol 2014;60(1):Suppl. p S413.<br />

Disclosures:<br />

The following authors have nothing to disclose: David I. Sherman, Waleed<br />

Fateen, Minal J Sangwaiya, Paul Tadrous, Philip J. Shorvon<br />

801<br />

A clustering based fully automated method for collagen<br />

proportional area extraction in liver biopsy images<br />

Nikos Giannakeas 2 , Zoi Tsianou 2 , Markos Tsipouras 2 , Alexandros<br />

T. Tzallas 3 , Pinelopi Manousou 1 , Epameinondas Tsianos 2 ;<br />

1 Institute for Liver and Digestive Health, London, Royal Free Hospital<br />

and UCL, London, United Kingdom; 2 1st Division of Internal<br />

Medicine and Division of Gastroenterology, Faculty of Medicine,<br />

School of Health Sciences, University of Ioannina, Ioannina,<br />

Greece; 3 School of Applied Technology, Department of Computer<br />

Engineering, Technological Educational Institute of Epirus, Arta,<br />

Greece<br />

Background and Aim: Collagen Proportional Area (CPA)<br />

extraction using digital image analysis (DIA) in liver biopsies<br />

provides an effective way to estimate liver disease staging.<br />

CPA represents accurately fibrosis expansion in liver tissue<br />

compared to semiquantitative staging scores. However, CPA<br />

has not reached everyday clinical practice. The lack of standardized<br />

and robust methods for computerized DIA for CPA<br />

assessment is a major limitation. We aim to create a fully<br />

automated methodology for CPA computation in liver biopsy<br />

images. Method: The proposed methodology is based in<br />

three stages. 1)The tissue detection stage: a k–means clustering<br />

approach is used to separate the tissue from the background.<br />

2)The tissue characterization stage: machine learning<br />

techniques are employed to characterize liver tissue, muscle,<br />

vessels, blood clots, structural collagen, dye stain and artifacts<br />

(scratches or contaminations of the substrate). Several shape<br />

and color features are extracted from each tissue area. This set<br />

of features is used for training and testing a classification algorithm,<br />

for automated area classification. 3) The CPA calculation<br />

stage: this stage also employs the k-means clustering, to separate<br />

fibrosis areas from normal liver tissue. CPA is computed as<br />

the number of fibrosis pixels divided by the number of pixels<br />

of the whole liver tissue. Results: The proposed methodology<br />

was evaluated using a dataset of 93 images of liver biopsies,<br />

obtained from 93 hepatitis C patients. The dataset was kindly<br />

provided by the Royal Free Hospital, London and our results<br />

were compared with the standardized method. Liver biopsies<br />

were formalin fixed, paraffin embedded and stained using<br />

Picrosirius red. High accuracy results were obtained for all<br />

stages of the proposed methodology (1 st stage: 98% accuracy<br />

in tissue detection, 2 nd stage: 85% accuracy in tissue area<br />

characterization, and 3 rd stage: less than 2% mean error). Conclusions:<br />

According to the results, CPA is extracted accurately<br />

in most of the cases (errors


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 609A<br />

802<br />

Feasibility of three ultrasound shear wave elastographic<br />

methods in overweight and obese patients<br />

Ioan Sporea; University of Medicine and Pharmacy Timisoara,<br />

Timisoara, Romania<br />

In the last decade, different types of ultrasound based elastographic<br />

methods that noninvasively quantify liver fibrosis<br />

have been developed. Even though, Transient Elastography<br />

is a validated method for liver fibrosis assessment in chronic<br />

B and C hepatitis, reliable elasticity measurements are difficult<br />

to obtain in obese (BMI>30kg/m 2 ) patients. The aim of this<br />

study was to compare the feasibility of three elastrographic<br />

methods used for liver fibrosis evaluation (Transient Elastography-TE,<br />

point Shear Wave Elastography (pSWE) using ARFI<br />

technique - VTQ and SuperSonic Shear Imaging-2D-SWE) in<br />

overweight and obese patients. Material and methods: The<br />

study included 183 consecutive subjects with or without chronic<br />

hepatopathies, in which liver stiffness (LS) was evaluated in<br />

the same session by means of 3 elastographic methods: TE<br />

(Fibroscan, Echosens), VTQ (Siemens Acuson S200 TM ) and<br />

2D-SWE (Aixplorer, SuperSonic Imagine S. A). Reliable LS<br />

measurements were defined as follows: for TE and VTQ – the<br />

median value of 10 LS measurements with a success rate≥60%<br />

and an interquartile range 25kg/m 2 ), 32.2 % with<br />

obesity grade 1 (BMI=30-34.9 kg/m 2 ), 4.3% with obesity<br />

grade 2 (BMI=35-39.9 kg/m 2 ) and 1.1% with obesity grade<br />

3 (BMI>40 kg/m 2 ). From the 114 overweight patients, reliable<br />

LS measurements (LSM) were obtained in 83.3% (95%)<br />

by means of TE, 85.1% (97) by means of VTQ and in 88.6%<br />

(101) by 2D-SWE. TE, 2D-SWE and VTQ had similar rates<br />

of reliable LSM in overweight patients: TE (83.3%) vs. VTQ<br />

(85.1%) p=0.84; TE (83.3%) vs. 2D-SWE (88.6%) p=0.33;<br />

2D-SWE (88.6%) vs. VTQ (85.1% ) p=0.55. The percentage of<br />

reliable LSM in obese patients was similar: VTQ (79.7%) vs TE<br />

(82.6%) (p=0.82), 2D SWE (85.5%) vs TE (82.6%) p=0.81,<br />

VTQ (79.7%) vs 2D SWE (85.5%) p=0.5 Conclusion: Feasibility<br />

of TE, VTQ and 2D-SWE in overweight and obese patients<br />

was similar.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ioan Sporea<br />

803<br />

Fast macromolecular proton fraction (MPF) mapping of<br />

the human liver in vivo for quantitative assessment of<br />

hepatic fibrosis<br />

Vasily L. Yarnykh 1,2 , Erica Tartaglione 3 , George N. Ioannou 3,4 ;<br />

1 Radiology, University of Washington, Seattle, WA; 2 Research<br />

Institute of Biology and Biophysics, Tomsk State University, Tomsk,<br />

Russian Federation; 3 Research and Development, Veterans Affairs<br />

Puget Sound Health Care System, Seattle, WA; 4 Medicine, Division<br />

of Gastroenterology, University of Washington, Seattle, WA<br />

Background: MPF is a quantitative MRI parameter determining<br />

the magnetization transfer (MT) effect and defined as a relative<br />

amount of immobile macromolecular protons involved into<br />

magnetization exchange with mobile water protons. MPF has<br />

a potential for quantitative assessment of fibrous tissue due to<br />

intrinsically high MPF in collagen. A recent fast single-point<br />

MPF mapping method (Magn Reson Med 2012;68:166)<br />

enables the capability of liver MPF assessment. The goal of this<br />

study was to prospectively investigate a relationship between<br />

MPF in the liver parenchyma and fibrosis stage. Methods:<br />

16 patients with chronic HCV infection underwent MRI at 3T<br />

using a three-dimensional MPF mapping protocol comprising<br />

four breath-hold scans obtained with 2x3x6 mm 3 voxel size<br />

and 10 sections:1) dynamic acquisition of MT-weighted (TR/<br />

TE=18.5/2.3 ms, flip angle (FA)=80, saturation frequency 2<br />

kHz and FA=2520) and reference (same parameters, no saturation)<br />

images; 2) dynamic acquisition of three images for variable<br />

FA T1 mapping (same TR/TE, FA=3,10,20°); 3) dual-TE<br />

B0 map; and 4) dual-TR B1 map. MPF maps were reconstructed<br />

using the single-point algorithm. 14 patients had liver<br />

biopsy within one year prior to study entry, and two patients<br />

had clinically diagnosed cirrhosis. Fibrosis was histologically<br />

staged according to METAVIR. Score F4 was assigned to clinical<br />

cirrhosis. The mode of liver MPF histograms was compared<br />

between groups with (F2-F4, n=6) and without (F0-F1, n=10)<br />

significant fibrosis by Mann-Whitney test. Association between<br />

MPF and fibrosis score tested using Spearman correlation coefficient<br />

(ρ). Results: MPF was increased in the significant fibrosis<br />

group compared to the no or mild fibrosis group (6.49±0.36%<br />

vs. 5.94±0.26%, P


610A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

7 days. Oxysterols and bile acids in the liver and brain were<br />

extracted, and analyzed by HPLC or LC/MS. Hepatic mRNA<br />

were determined by RT-PCR. Primary hepatocytes were also<br />

isolated from the same strains of mice. Cultures were incubated<br />

with [ 14 C]-cholesterol and formed [ 14 C]-oxysterols/bile acids<br />

were analyzed. Results: In wild type mice, hepatic CYP7B1<br />

mRNA expression was suppressed (>80%Ñ) after StARD1 overexpression.<br />

SHP mRNA expression was markedly increased<br />

(300%É), while CYP7A1(40%Ñ) and CYP8B1(80%Ñ) mRNAs<br />

were decreased. Interestingly, gallbladder bile acid concentration<br />

in these mice was 30% higher as compared to control<br />

mice. Primary hepatocyte culture experiments showed that the<br />

rate of bile acid synthesis increased 3-fold after StARD1 overexpression.<br />

These observations implied that an additional bile<br />

acid synthetic pathway is present which compensates for the<br />

CYP7A1 and the CYP27A1/CYP7B1 pathways. StARD1 overexpressed<br />

liver accumulated 27-Hydroxycholesterol(HC) and<br />

25HC. Also found was unexpected high level of 24HC(Ratio of<br />

24, 25, 27HC, 6:1:6). Level of 24HC in the brain was identical<br />

to control. When primary hepatocytes were incubated with<br />

[ 14 C]-cholesterol following StARD1 overexpression, wild type<br />

hepatocytes formed all three [ 14 C]-oxysterols, 24HC, 25HC<br />

and 27HC(ratio, 3:1:10). However, hepatocytes isolated from<br />

CYP27A1 -/- mice formed only [ 14 C]-25HC. These observations<br />

show that 24HC originates from liver mitochondrial CYP27A1.<br />

Furthermore, when StARD1 was overexpressed in CYP7B1 -/-<br />

mouse liver in vivo and in vitro, 27HC and 25HC accumulated<br />

but, not 24HC; suggesting 24HC is responsible for an alternative<br />

metabolic pathway to bile acids. Conclusion: The results<br />

indicate that hepatic CYP27A1 catalyzes not only 25- and<br />

27-hydroxylations, but also 24-hydroxylation. Since it has been<br />

reported that 24HC is metabolized by CYP7B1, CYP39A and<br />

CYP7A1, this oxysterol may be an important intermediate in a<br />

second alternative pathway to bile acids.<br />

Disclosures:<br />

William M. Pandak - Employment: Virginia Commonwealth University, Veterans<br />

Affairs Medical Center<br />

The following authors have nothing to disclose: Genta Kakiyama, Dalila M.<br />

Marques, Kuniko Mitamura, Hajime Takei, Hiroshi Nittono, Daniel Rodriguez-Agudo,<br />

Gregorio Gil, Huiping Zhou, Phillip B. Hylemon<br />

805<br />

Bile acid induced cholestatic liver injury involves TLR9<br />

initiated inflammatory signals in mouse hepatocytes<br />

Shi-Ying Cai, Xinshou Ouyang, Albert Mennone, Carol J. Soroka,<br />

Wajahat Z. Mehal, James L. Boyer; Yale Univ, New Haven, CT<br />

Background: The inflammatory response plays an important<br />

role in cholestatic liver injury where bile acid (BA) induction<br />

of proinflammatory cytokines in hepatocytes may initiate this<br />

event. However, the signaling pathways involving BA stimulation<br />

of cytokine production remain elusive, although BA injures<br />

mitochondria in hepatocytes. Toll-like receptors (TLR) respond<br />

to both endogenous sterile insults and pathogen recognitions<br />

and play a critical role in tissue damage initiated inflammatory<br />

responses. TLR9 has been characterized as an intracellular<br />

DNA receptor that, upon activation, stimulates cytokine production<br />

as part of the innate immune response. Aim: To determine<br />

Tlr9’s role in BA induced liver injury. Methods: Hepatocytes<br />

isolated from wild-type (WT) and Tlr9 knockout (KO) mice were<br />

treated with BAs. Mitochondrial damage was assessed using a<br />

JC-1 assay and Western blot. Q-PCR was used to detect inflammatory<br />

chemokine gene expression. WT and Tlr9 KO mice<br />

were subjected to bile duct ligation (BDL) or sham operation<br />

for 7 days. Plasma biochemistry, liver gene expression, and<br />

liver histology were examined. Results: At pathophysiological<br />

concentration (≥25 mM), taurocholic acid (TCA) caused mitochondrial<br />

membrane potential changes in mouse hepatocytes.<br />

Mitochondrial specific proteins cytochrome C, AIF and ER<br />

specific protein Grp78 were detected in the cytosolic fraction<br />

by Western blot analysis, indicating mitochondria damage<br />

and ER stress, whereas no ROS were detected. Cxcl2 induction<br />

was significantly less in Tlr9 KO hepatocytes than in WT<br />

hepatocytes (KO: 3.7-fold v/s WT: 14-fold) after treatment with<br />

100mM TCA for 24 hr. The synthetic Tlr9 ligand CpG ODNs<br />

and BAs demonstrated synergistic effects in stimulating Cxcl2<br />

expression in mouse hepatocytes, consistent with Tlr9’s involvement<br />

in BA induction of Cxcl2 expression. After BDL, plasma<br />

ALT, ALP and BA levels and liver BA levels were also significantly<br />

lower in Tlr9 KO mice compared to WT BDL mice, confirming<br />

a role for Tlr9 in mediating cholestatic liver injury. Bile<br />

duct proliferation, assessed by liver H&E histology and CK19<br />

labeling, and hepatic levels of Ccl2 and Cxcl2 mRNA were<br />

also significantly less in Tlr9 KO than in WT mice after BDL.<br />

Conclusion: BA injures mitochondria resulting in DNA release<br />

and triggers an innate immune response by activating Tlr9.<br />

Preventing mitochondrial damage or blocking Tlr9 activation<br />

may be new strategies for treating cholestasis.<br />

Disclosures:<br />

James L. Boyer - Advisory Committees or Review Panels: Pfizer; Consulting:<br />

abbvie<br />

The following authors have nothing to disclose: Shi-Ying Cai, Xinshou Ouyang,<br />

Albert Mennone, Carol J. Soroka, Wajahat Z. Mehal<br />

806<br />

Proteomics analysis of rat canalicular membrane<br />

reveals the expression of several P4-ATPases in liver<br />

Pururawa M. Chaubey, Bruno Stieger; Department of Clinical<br />

Pharmacology and Toxicology, University Hospital Zurich, Zurich,<br />

Switzerland<br />

Background: Transport processes in the canalicular membrane<br />

are key elements in bile formation and are the driving force<br />

for the enterohepatic circulation of bile salts. The canalicular<br />

membrane is constantly exposed to a hostile environment due<br />

to the detergent action of the high conentration of canalicular<br />

bile salts. Aims: To obtain more insights into the molecular entities<br />

rendering the canalicular membrane resistant to the detergent<br />

action of bile salts, the proteome of highly purified rat<br />

canalicular membrane vesicles was determined. Methods: Isolated<br />

canalicular membrane vesicles (JCB 98:991,1983) from<br />

Sprague Dawley rats were stripped from adherent proteins<br />

(JCB 93:97,1982), deglycosylated, protease digested and subjected<br />

to shot gun proteomic analysis. Expression of P4-ATPases<br />

was confirmed by RT-PCR and Western blotting. Localization of<br />

P4-ATPases was carried out with immunofluorescence methodology<br />

on liver sections. Results: We identified 1745 unique<br />

proteins in canalicular vesicles. A comparative analysis with a<br />

renal brush border membrane proteome and liver proteomes<br />

revealed a large set of membrane and membrane associated<br />

proteins (1287) specifically identified in our study, possibly<br />

because we enriched for membrane proteins. Assignment to<br />

functional categories of the newly identified proteins revealed<br />

an over-representation of transporter specific categories. We<br />

identified several transporters out of which 19 members are<br />

from SLC families and 11 members are from ABC families.<br />

Additionally, we also identified P4-ATPases (ATP8A1, ATP8B1,<br />

ATP9A, ATP11A, & ATP11C) and a P5-ATPase (ATP13A1).<br />

Four of 5 identified P4-ATPases are not known to be expressed<br />

in liver. Furthermore, expression analysis of P4-ATPases using<br />

RT-PCR and Western blotting could verify the expression of<br />

identified P4-ATPase with a rather high expression of ATP11C.<br />

Finally, ongoing immunofluorescence analysis demonstrates


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 611A<br />

localization of some of the identified P4-ATPase at the canalicular<br />

membrane. Conclusions: Our data suggest that there<br />

is not a large common proteome between the brush border<br />

membrane of kidney proximal tubule cells and the canalicular<br />

membrane. Furthermore, the identification of several P4-AT-<br />

Pases in rat liver together with the previous findings on mice<br />

with disrupted Atp11c displaying hyperbilirubinemia (PNAS<br />

108:7890,2011) suggest that important proteins involved in<br />

hepatocellular lipid homeostasis and canalicular lipid secretion<br />

may await characterization. Acknowledgement: The expert<br />

support by Dr. B Roschitzki from the Functional Genomics Center<br />

in Zurich is greatly appreciated.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pururawa M. Chaubey, Bruno<br />

Stieger<br />

807<br />

TGR5 agonists improve pancreatic beta cell mass and<br />

function by reprogramming alpha cells to produce GLP-<br />

1: a novel mechanism of action of bile acids in diabetes<br />

Divya P. Kumar 1 , Amon Asgharpour 2 , Faridoddin Mirshahi 2 , So<br />

Hyun Park 3 , Sichen Liu 3 , Yumi Imai 3 , Jerry L. Nadler 3 , Karnam S.<br />

Murthy 1 , Arun J. Sanyal 2 ; 1 Physiology and Biophysics, Virginia<br />

Commonwealth University, Richmond, VA; 2 Internal Medicine, Virginia<br />

Commonwealth University, Richmond, VA; 3 Internal Medicine,<br />

Eastern Virginia Medical School, Norfolk, VA<br />

BACKGROUND: Pancreatic α cells secrete glucagon (GLU)<br />

under euglycemic conditions and GLP-1 under hyperglycemic<br />

conditions. GLU and GLP-1 are derived from alternate splicing<br />

of a common precursor by prohormone convertase 2 (PC2) and<br />

PC1 respectively. Bile acids (BA) via their cognate receptors,<br />

TGR5, stimulate GLP-1 release in intestinal L cells. Recently,<br />

TGR5 have been identified on pancreatic α cells. The potential<br />

effects of BA on pancreatic α cells and their physiological<br />

relevance are unknown. HYPOTHESIS: BA, acting via TGR5<br />

on pancreatic α cells, enhances hyperglycemia-induced PC1<br />

expression thereby releasing GLP-1 which, in turn, acts as a<br />

paracrine to increase pancreatic β cell mass and function.<br />

AIMS: To examine (1) the in vivo effect (intraperitoneal) of<br />

TGR5 specific agonist, INT-777 on glucose homeostasis, PC1<br />

expression, GLP-1 release, β cell proliferation in control and<br />

db/db mice (2) the effect of GLP-1 receptor antagonist on<br />

TGR5-mediated insulin secretion in human islets, and (3) the in<br />

vitro effect of INT-777 on β cell proliferation. METHODS: The<br />

in vivo effect of INT-777 on body weight, glucose homeostasis<br />

(glucose and insulin tolerance test), PC1 expression (immunofluorescence<br />

and RT-PCR), GLP-1 release (ELISA) and β cell<br />

mass and proliferation (morphometric analysis of pancreatic<br />

sections) was examined in control and db/db mice. The in vitro<br />

effect of INT-777 on β cell proliferation was measured by MTT<br />

assay in MIN6 cells. Insulin secretion in response to INT-777<br />

from human islets was measured in the presence or absence of<br />

GLP-1 receptor antagonist, exendin (9-39) by ELISA. RESULTS:<br />

Treatment with INT-777 for 7 weeks attenuated the increase<br />

in body weight, and improved glucose tolerance and insulin<br />

sensitivity in db/db mice compared to control. INT-777 augmented<br />

PC1 expression in α cells and stimulated GLP-1 release<br />

from islets of db/db mice compared to control. INT-777 also<br />

increased pancreatic β cell proliferation and insulin synthesis<br />

in vivo, but had no effect on β cell proliferation in vitro. Insulin<br />

secretion by INT-777 in human islets cultured under high<br />

(25mM) glucose was inhibited by exendin (9-39). These results<br />

suggest that in diabetes, GLP-1 released from α cells augments<br />

β cell proliferation and function. CONCLUSION: In diabetes,<br />

BA reprogram α cells to switch from GLU synthesis to GLP-1<br />

synthesis by increasing PC1 expression with associated trophic<br />

effects on adjacent β cells and improvement in insulin sensitivity<br />

and hyperglycemia. Thus, these results identify a novel mechanism<br />

of action of TGR5 agonist INT-777 in glucose homeostasis<br />

and its potential use in the treatment of insulin resistance and<br />

type 2 diabetes.<br />

Disclosures:<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: Divya P. Kumar, Amon Asgharpour,<br />

Faridoddin Mirshahi, So Hyun Park, Sichen Liu, Yumi Imai, Jerry L. Nadler,<br />

Karnam S. Murthy<br />

808<br />

Higher order metabolites of heme increase under conditions<br />

of hyperbilirubinemia, trigger cholestasis and<br />

impair hepatocellular integrity<br />

Raphael A. Seidel 1,5 , Franziska Schleser 1 , Christoph Sponholz 1 ,<br />

Frans Cuperus 2,3 , Sandor Nietzsche 4 , Georg Pohnert 5 , Michael<br />

Bauer 1 ; 1 Department of Anaesthesiology and Intensive Care<br />

Medicine / Center for Sepsis Control and Care, Jena University<br />

Hospital, Jena, Germany; 2 Hans Popper Laboratory of Molecular<br />

Hepatology, Division of Gastroenterology and Hepatology,<br />

Department of Internal Medicine III, Medical University of Vienna,<br />

Vienna, Austria; 3 Pediatric Gastroenterology and Hepatology,<br />

Department of Pediatrics, Center for Liver, Digestive, and Metabolic<br />

Diseases, Beatrix Children’s Hospital - University Medical<br />

Center Groningen, University of Groningen, Groningen, Netherlands;<br />

4 Electron Microscopy Center, Jena University Hospital, Jena,<br />

Germany; 5 Institute of Inorganic and Analytical Chemistry, Friedrich<br />

Schiller University Jena, Jena, Germany<br />

Background & Aim: Heme catabolism physiologically produces<br />

large amounts of bilirubin that can function as potent and regenerative<br />

scavenger of reactive oxygen species (ROS). Nonetheless,<br />

ROS attacking heme, biliverdin or bilirubin can lead to the<br />

formation of higher order metabolites of heme, among them the<br />

regio-isomeric bilirubin oxidation products (BOXes) Z-BOX A<br />

and Z-BOX B, with unknown implications for hepatic function.<br />

The aim of this study was to clarify the appearance of Z-BOX<br />

A and Z-BOX B under conditions of health and hyperbilirubinemia<br />

as well as their fate and functional role in the liver as<br />

a key-player in heme catabolism. Methods: Determination of<br />

Z-BOX A and Z-BOX B was carried out via HPLC-MS/MS procedures<br />

in patients suffering from liver failure, healthy human<br />

controls, Gunn rats as animal model of hereditary bilirubin<br />

conjugation failure, and Wistar rats. Pharmacokinetics, pharmacodynamics<br />

and hepatic hemodynamics were studied in<br />

Wistar rats. Cytoskeletal remodeling, cytotoxicity, and cellular<br />

metabolism were investigated in cultured HepG2 and HepaRG<br />

cells by means of fluorescence and scanning electron microscopy<br />

in addition to biochemical assays. Results: The higher<br />

order heme metabolites Z-BOX A and Z-BOX B were formed<br />

under physiological conditions across species (combined levels<br />

in blood: 25.4 ± 3.4 nmolL -1 , n = 10, healthy human;<br />

19.1 ± 4.6 nmolL -1 , n = 12, Wistar rat) and were significantly<br />

elevated under conditions of hyperbilirubinemia, i.e. human<br />

liver failure (504.3 ± 30.3 nmolL -1 , n = 28; p < 0.0001) and<br />

animal model of Crigler-Najjar syndrome type I (136.9 ± 15.1<br />

nmolL -1 , Gunn rat, n = 14; p < 0.0001). Z-BOX A and Z-BOX<br />

B appeared in an almost fixed ratio (~1.45:1) and highly correlated<br />

with serum bilirubin (r > 0.86, p = 4.110 19 ). After<br />

equimolar i.v. bolus application, the pharmacokinetic profiles<br />

of these regio-isomers exhibited significantly shorter serum half-


612A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

life and stronger biliary enrichment (approx. 30x) of Z-BOX A<br />

compared to Z-BOX B. In parallel, Z-BOX A and Z-BOX B triggered<br />

an instantaneous and significant reduction of bile flow<br />

(1 min vs. 3.5 min after bolus injection; n = 5; p = 0.0002).<br />

Regardless their reported vasoconstrictive properties in cerebral<br />

arterioles, Z-BOX A and Z-BOX B did not affect hepatic<br />

hemodynamics. However, Z-BOX A and Z-BOX B showed a<br />

dose-dependent decrease in cellular metabolic rate (IC 50<br />

=<br />

383 mmolL -1 and 141 mmolL -1 , respectively), while only Z-BOX<br />

A induced a cytoskeletal remodeling of cultured hepatocytes.<br />

Conclusions: Higher order metabolites of heme are elevated<br />

under conditions of hyperbilirubinemia and are able to impair<br />

bile production and cellular integrity.<br />

Disclosures:<br />

The following authors have nothing to disclose: Raphael A. Seidel, Franziska<br />

Schleser, Christoph Sponholz, Frans Cuperus, Sandor Nietzsche, Georg Pohnert,<br />

Michael Bauer<br />

809<br />

NF-kB is critically important for the up-regulation of<br />

organic solute transporter OSTα/OSTβ in cholestasis<br />

Natarajan Balasubramaniyan, Frederick J. Suchy; Pediatrics, University<br />

of Colorado Denver, Aurora, CO<br />

The up-regulation of the organic solute transporter OSTα/<br />

OSTβ in cholestasis provides an alterative pathway for excretion<br />

of bile acids and other cholephiles across the hepatocyte<br />

basolateral membrane.However, the mechanisms underlying<br />

this adaptation have been enigmatic. The mouse and human<br />

Ostα/β(OSTα/β) promoters contain functional FXR and LRH<br />

elements, which mediate, respectively, positive and negative<br />

feedback regulation by bile acids in normal liver, but these<br />

binding sites don’t explain how this transport system is up-regulated<br />

in cholestasis when FXR-dependent pathways are compromised.We<br />

identified previously unknown,well-conserved NF-kB<br />

binding sites in the Ostα(OSTα) and Ostβ(OSTβ)promoters.<br />

On chromatin immunoprecipitation(ChIP) analysis of Huh-7<br />

cells,over-expressed NF-kB was recruited to these sites. In vivo<br />

ChIP assays done with liver nuclei obtained from mice after 3<br />

days of CBDL or 6 hours post LPS injection, showed markedly<br />

increased recruitment of NF-kB p65 to the NF-kB binding sites<br />

of the Ostα and Ostβ promoters and a diminution of FXR at its<br />

binding sites (FXRE). Next Huh-7 cells were transiently transfected<br />

with plasmid vectors containing the OSTα and OSTβ<br />

promoter sequences that included the NFkB response element<br />

linked to luciferase as reporter. In contrast to the inhibitory<br />

effect of NFkB p65 on activity of the BSEP promoter, expression<br />

of NFkB p65 and p50 subunits together synergistically<br />

activated the OSTα and OSTβ promoters in a dose-dependent<br />

fashion. The inhibitors IkBα and IkBSR blocked induction by<br />

p65 and p50. Previous <strong>studies</strong> have shown that NF-kB, acting<br />

as a transcriptional activator, recruits a coactivator complex<br />

that has striking similarities to that recruited by nuclear receptors<br />

which usually includes the cyclic AMP response element<br />

binding protein (CREB)-binding protein (CBP) and p300.CBP/<br />

p300 is particularly important in possessing intrinsic histone<br />

acetyltransferase activity and providing a platform for recruitment<br />

of a variety of coactivator proteins required for NF-kB-dependent<br />

gene expression. In vivo ChIP assays using liver nuclei<br />

obtained from both cholestatic mouse models and an antibody<br />

that recognizes the homologous CBP and p300 proteins<br />

showed a markedly increased recruitment of CBP/p300 to the<br />

NF-kB sites in the Ostα and Ostβ promoters compared with<br />

controls. In contrast, there was significant depletion of CBP/<br />

p300 at the FXRE of Ostα and Ostβ in these models compared<br />

with controls. These <strong>studies</strong> reveal a novel, NF-kB-dependent<br />

mechanism for the up-regulation OSTα/OSTβ in cholestasis,<br />

providing an alternative export pathway to prevent accumulation<br />

of hydrophobic bile salts and other toxic products.<br />

Disclosures:<br />

The following authors have nothing to disclose: Natarajan Balasubramaniyan,<br />

Frederick J. Suchy<br />

810<br />

The Ileal Bile Acid Transporter Inhibitor A4250 Modulates<br />

Bile Acid Synthesis and Decreases Serum Bile<br />

Acids.<br />

Hanns-Ulrich Marschall 1 , Per-Göran Gillberg 2 , Hans Graffner 2 , Leif<br />

Rikner 2 ; 1 Department of Molecular and Clinical Medicine, Sahlgrenska<br />

Academy, Institute of Medicine, University of Gothenburg,<br />

Gothenburg, Sweden; 2 Albireo, Gothenburg, Sweden<br />

Background and Aim: Reabsorption of BAs from the intestine<br />

by ileal bile acid transporter (IBAT) is pivotal for the enterohepatic<br />

circulation of BAs and sterol homeostasis. We aimed to<br />

study BA metabolism in a phase 1 trial with the selective IBAT<br />

inhibitor A4250. Methods: In a double-blind, multiple ascending<br />

dose design, A4250 capsules or matching placebo was<br />

administered for 7 days; 1 mg once daily; 3 mg once daily<br />

and 1.5 mg twice daily, respectively. Each group consisted of<br />

8 healthy volunteers whereof 6 received active compound and<br />

2 placebo. BAs were measured by HPLC-MS in plasma and<br />

feces, FGF19 and C4 in plasma. Results: No serious adverse<br />

events occurred and all participants finished the trial per protocol.<br />

Plasma total BAs, and FGF19 decreased, whereas plasma<br />

C4 and fecal BAs increased. The majority of fecal BAs in the<br />

A4250 groups were the primary BAs cholic and chenodeoxycholic<br />

acids. Conclusion: A4250 is a highly efficient and<br />

well-tolerated compound that by blocking the IBAT in the terminal<br />

ileum interrupts the enterohepatic circulation of BAs, which<br />

should be of benefit in patients with cholestatic liver diseases.<br />

Disclosures:<br />

Hanns-Ulrich Marschall - Consulting: Albireo<br />

Per-Göran Gillberg - Employment: Albireo<br />

Hans Graffner - Employment: Albireo<br />

Leif Rikner - Consulting: Albireo AB<br />

811<br />

Inhibition of Adenylyl Cyclase 5 reduces cyst growth in<br />

mice with polycystic liver disease (ADPKD) caused by<br />

defective Polycystin-2<br />

Carlo Spirli 1 , Ambra Villani 1 , Valeria Mariotti 1,2 , Luca Fabris 3 ,<br />

Romina Fiorotto 1 , Mario Strazzabosco 1,2 ; 1 Yale University, Section<br />

of Digestive Diseases, New Haven, CT; 2 Department of Surgery<br />

and Interdisciplinary Medicine, University of Milano-Bicocca,<br />

Milan, Italy; 3 Department of Molecular Medicine, University of<br />

Padova, Padua, Italy<br />

Genetic defects in Polycystin-1 (PC1) or Polycystin-2 (PC2)<br />

cause polycystic liver disease associated with ADPKD (PLD). In<br />

PC-2 defective cystic cholangiocytes, the interaction of STIM1<br />

(the molecular sensor that mediates Ca 2+ entry following<br />

ER-Ca 2+ decrease) with ORAI channels is uncoupled accounting<br />

for a reduced Store Operated Calcium Entry (SOCE) and<br />

low intracellular and Endoplasmic Reticulum (ER) Ca 2+ levels.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 613A<br />

On the other hand, in PC2-defective cells STIM-1 couples with<br />

a Ca 2+ -inhibitable Adenylyl Cyclase (ACs) and stimulates the<br />

store-operated production of cAMP (SOcAMP), PKA-dependent<br />

activation of the ERK1/2 and increase in VEGF production<br />

by cystic cells. This mechanism plays a key role in promoting<br />

liver cyst growth in PLC. Two Ca 2+ -inhibitable ACs (AC6 and<br />

AC5) are expressed in cholangiocytes and have similar properties,<br />

however the molecular identity of the AC responsible for<br />

inappropriate cAMP production in PLC is unknown. Results:<br />

PC2-conditional KO mouse (Pkd2 flox/- ;pCxreER TM ) (Pkd2cKO)<br />

were crossed with AC6 -/- mouse to generate double Pkd2/<br />

AC6KO mice. Successful deletion was confirmed by RT/PCR.<br />

Mice were sacrificed eight weeks after induction of PC2 excision<br />

and cystic area was assessed by K19 staining and computer-assisted<br />

morphometric analysis. However, Pkd2/AC6KO<br />

mice did not show a reduction of the liver cystic area as respect<br />

to Pkd2cKO mice. Moreover, in cystic cholangiocytes isolated<br />

from Pkd2/AC6KO mice, intracellular levels of cAMP generated<br />

after an acute ER [Ca 2+ ] depletion (TPEN 1mM) were not<br />

different from those measured in cholangiocytes from Pkd2cKO<br />

mice, suggesting that in absence of AC6, the AC5 isoform, may<br />

be activated. In fact, inhibition of AC5 by two different inhibitors<br />

(NKY90 and SQ 22,536) significantly reduced cAMP<br />

following ER [Ca 2+ ] depletion. Consistent with these data, in<br />

vivo treatment of Pkd2/AC6KO with SQ 22,536 ((300 mg/Kg<br />

day, I.P) caused a significant reduction in liver cystic area was<br />

significantly reduced, along with the liver/body weight ratio.<br />

In conclusion, these data indicate that AC5 is the Ca2+ inhibitable<br />

AC responsible for cAMP production by PC2-defective<br />

cells in condition of low cellular and ER [Ca 2+ ] and the mechanisms<br />

is of pathophysiological relevance. Inhibition of cAMP<br />

production by somatostatin analogs has been shown to be of<br />

therapeutic value in patients with PLC. Our study indicate that<br />

pharmacologic inhibition of AC5 may represent a more specific<br />

target to reduce of cyst growth in polycystic liver diseases.<br />

Disclosures:<br />

The following authors have nothing to disclose: Carlo Spirli, Ambra Villani, Valeria<br />

Mariotti, Luca Fabris, Romina Fiorotto, Mario Strazzabosco<br />

812<br />

Cholestatic-induced biliary proliferation and fibrosis<br />

are decreased in mice lacking mast cells: an innovative<br />

study using C-kit knockout mice<br />

Laura Hargrove 1 , Lindsey Kennedy 2 , Jennifer Owens 2 , Heather<br />

L. Francis 2,1 ; 1 Scott and White Memorial Hospital, Temple, TX;<br />

2 Research, Central Texas Veterans Health Care System, Temple, TX<br />

Background: Cholestatic liver injury is marked by increased<br />

biliary proliferation. Bile duct ligation (BDL) induces large, but<br />

not small, cholangiocyte proliferation coupled with increased<br />

fibrosis. Mast cells (MCs) are inflammatory cells that, once<br />

activated, release histamine (HA). Our previous <strong>studies</strong> have<br />

demonstrated that HA increases vascular endothelial growth<br />

factor (VEGF) expression and knockdown of histidine decarboxylase<br />

(HDC (synthesizes histamine)) decreases biliary VEGF<br />

expression/secretion. We have demonstrated that MCs infiltrate<br />

the liver following BDL increasing intrahepatic bile duct<br />

mass (IBDM) and fibrosis. Inhibition of MC-derived HA using<br />

cromolyn sodium decreases these parameters suggesting that<br />

MCs contribute to biliary proliferation and fibrosis. Our aim<br />

was to evaluate the effects of BDL on biliary proliferation and<br />

fibrosis in MC deficient mice. Methods: WT and C-kit -/- mice<br />

(MC deficient) were subjected to sham or BDL for 3 days before<br />

collecting serum, liver blocks and pure cholangiocytes. IBDM<br />

and proliferation were evaluated in sections by CK-19 and<br />

PCNA immunohistochemistry, respectively. In total liver (TL)<br />

and cholangiocytes we evaluated PCNA expression by qPCR<br />

and western blotting. Apoptosis was measured by gene expression<br />

for Bax and BCL-2 and TUNEL staining in tissues. Fibrosis<br />

was detected by Sirius Red/Fast Green staining in tissues and<br />

by qPCR for α-SMA, fibronectin, collagen type-1a and TGFβ1<br />

in TL. In cholangiocytes and TL we measured HDC and<br />

VEGF-A/C, gene expression and secretion by EIA in serum<br />

and cholangiocyte supernatants. To determine if MC-derived<br />

HA regulates biliary proliferation and fibrosis in vitro, cultured<br />

MCs were transfected with HDC shRNA prior to co-culture with<br />

murine cholangiocytes. Biliary proliferation was measured by<br />

MTS, and VEGF expression and fibrosis marker expression by<br />

qPCR. Results: In BDL C-kit -/- mice there was enhanced ductular<br />

reaction with increased small IBDM and decreased large IBDM<br />

compared to BDL WT. Proliferation and liver fibrosis were<br />

decreased in BDL C-kit -/- mice compared to BDL WT coupled<br />

with enhanced apoptosis. HDC and VEGF-A/C expression was<br />

decreased in BDL C-kit -/- mice as was HA and VEGF secretion<br />

in serum and supernatants compared to BDL WT. In vitro,<br />

knockdown of MC-HDC expression decreased biliary proliferation,<br />

VEGF expression and fibrosis suggesting MC-derived<br />

HA contributes to biliary proliferation and fibrosis. Conclusion:<br />

BDL-induced biliary proliferation and fibrosis are decreased in<br />

mice lacking mast cells via the HDC/HA/VEGF axis. Our <strong>studies</strong><br />

support the novel concept that hepatic mast cells are critical<br />

to biliary response following injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Laura Hargrove, Lindsey Kennedy,<br />

Jennifer Owens, Heather L. Francis<br />

813<br />

YAP Reduces Liver Parenchymal Damage by Promoting<br />

Bile Duct Capacity in Cholestatic Injury<br />

Quy P. Nguyen 1 , Nan Wu 2 , Tianhao Zhou 2 , Haibo Bai 1 ; 1 BaylorScott&White<br />

Hospital, Temple, TX; 2 Texas A&M University College<br />

of Medicine, Temple, TX<br />

Background&Aims: YAP is highly expressed in cholangiocytes.<br />

YAP protein levels are further increased in cholangiocytes after<br />

bile duct ligation (BDL). However, the physiological role of<br />

biliary YAP in maintaining biliary homeostasis and in response<br />

to cholestatic injury is not clear due to lacking efficient methods<br />

to ablate YAP in cholangiocytes. Methods: We found that overexpression<br />

of YAP in hepatocytes induces a compensatory loss<br />

of YAP in cholangiocytes. Moreover, even suppressing YAP<br />

activity with transgenic Vgl4, a YAP competitor, is not able to<br />

rescue the YAP deficiency in cholangiocytes. Because Yap/<br />

Vgl4 double transgenic mice (Yap/Vgl4 DTg) have normal<br />

liver function, we use it as model mice to study the function of<br />

YAP in cholangiocytes. Sham or BDL was performed with these<br />

groups of mice. Two weeks later, liver histology, double immunofluorescence<br />

staining (E-cadherin and CK19), immunohistochemistry<br />

staining (CK19), Sirius red staining were compared<br />

between control and Yap/Vgl4 DTg groups. Primary cholangiocytes<br />

were isolated from CK19-CreER; Yap flox/flox mice and<br />

cultured in vitro. Yap deletion was induced through addion of<br />

4-OH tamoxifen. Immunofluorescence staining of E-cadherin<br />

was compared between control and Yap KO cholangiocytes.<br />

YAP expression in bile ducts of the Mdr2 -/- mouse model of<br />

primary sclerosing cholangities was also studied by immunohistochemistry.<br />

Results: Bile infarcts and fibrosis were more<br />

severe in Yap/Vgl4 DTg mice compared to their WT littermate<br />

controls. CK19 staining revealed bile ducts are more dilated<br />

in WT controls than in Yap/Vgl4 DTg livers. E-cadherin staining<br />

revealed more well formed E-cadherin junction between<br />

cholangiocytes in Yap/Vgl4 DTg mice. E-cadherin junctions


614A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

are also more mature in primary cholangiocytes with Yap deficiency.<br />

YAP activity correlates with Mdr2-/- mice disease progress.<br />

Conclusions: YAP functions to increase bile duct dilation<br />

upon bile duct obstruction through promoting cholangiocyte<br />

E-cadherin junction dynamics. The dilated bile ducts are able<br />

to hold more bile in bile ducts of WT controls, which in turn<br />

reduces hepatocyte damage caused by bile overflow.<br />

Disclosures:<br />

The following authors have nothing to disclose: Quy P. Nguyen, Nan Wu, Tianhao<br />

Zhou, Haibo Bai<br />

814<br />

Glutathione depletion is critical for the early pathogenesis<br />

of biliary atresia (BA) in the toxin-induced BA model<br />

Xiao Zhao 1 , Benjamin J. Wilkins 2 , Kristin Lorent 1 , John R. Porter 3 ,<br />

Rebecca G. Wells 1 , Michael Pack 1 ; 1 Medicine/Division of Gastroenterology,<br />

University of Pennsylvania, Philadelphia, PA; 2 Division<br />

of Anatomic Pathology, Department of Pathology and Laboratory<br />

Medicine, The Children’s Hospital of Philadelphia, Philadelphia,<br />

PA; 3 University of the Sciences, Philadelphia, PA<br />

Biliary atresia (BA) is a complex neonatal cholangiopathy that<br />

is the leading indication for liver transplantation in the pediatric<br />

population. Using an in vivo zebrafish biliary assay, we have<br />

isolated biliatresone, a novel plant isoflavonoid with selective<br />

extrahepatic biliary toxicity that is responsible for BA outbreaks<br />

in newborn Australian livestock. The phenotype induced in<br />

this new BA model is conserved between zebrafish and mammals,<br />

and recapitulates the cardinal features of human BA.<br />

While extrahepatic cholangiocytes are exquisitely sensitive to<br />

biliatresone, hepatocytes and intrahepatic cholangiocytes are<br />

resistant to its deleterious effect. This difference in susceptibility<br />

can potentially be explained by the intrinsic differential capability<br />

of specific liver cell types in mitigating stress. Global<br />

expression profiling of mRNA recovered from larval zebrafish<br />

cholangiocytes and total liver revealed that biliatresone<br />

leads to early up-regulation of the glutathione (GSH) redox<br />

response, suggesting that GSH depletion is an inciting event<br />

in biliatresone-mediated biliary injury. GSH, as a ubiquitous<br />

tripeptide thiol, is a vital cellular antioxidant. These data are<br />

consistent with in vitro assays that confirm the reactivity of biliatresone<br />

with glutathione. Using a specific genetically encoded<br />

GSH redox probe (redox-sensitive GFP, roGFP), we were able<br />

to establish quantitative in vivo mapping of the GSH redox<br />

potential (E GSH<br />

) in hepatocytes and extrahepatic cholangiocyte<br />

(EHC) of the biosensor-transgenic larvae. Interestingly, we<br />

detected endogenous differences in the redox status between<br />

hepatocytes and EHC. Pharmacological manipulation of GSH<br />

redox homeostasis further supported the importance of GSH in<br />

modulating biliatresone-induced injury. This is evidenced by the<br />

temporary attenuation of its biliary toxicity with N-acetylcysteine<br />

(NAC), a GSH precursor. Altogether, these data strongly<br />

suggest redox stress as a contributing factor in biliatresone-induced<br />

injury and differences in intrinsic stress responses can<br />

explain its cell-type specificity. Insufficient antioxidant capacity<br />

of EHC may potentially be critical to the early pathogenesis of<br />

BA.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xiao Zhao, Benjamin J. Wilkins,<br />

Kristin Lorent, John R. Porter, Rebecca G. Wells, Michael Pack<br />

815<br />

The bile acid Ursodeoxycholic acid activates cholangiocyte<br />

TMEM16A Cl- channels<br />

Qin Li, Amal K. Dutta, Charles Kresge, Andrew P. Feranchak;<br />

Department of Pediatric Gastronenterology, Hepatology, University<br />

of Texas Southwestern Medical Center, Dallas, TX<br />

Ursodeoxycholic acid (UDCA) stimulates a bicarbonate-rich<br />

choleresis in part through effects on cholangiocytes. During<br />

cholehepatic shunting, uptake of bile acids at the apical cholangiocyte<br />

membrane is associated with an increase in [Ca2+]<br />

I, Cl- efflux and Cl-/HCO3- exchange, though the cellular<br />

mechanism is unknown. As TMEM16A is a Ca2+-activated Clchannel<br />

in the apical membrane of cholangiocytes (JBC 2011;<br />

286:766-776) the aim of the present study was to determine if<br />

TMEM16A is the target of UDCA-stimulated Cl- secretion and<br />

to identify the regulatory pathway involved. Methods: Studies<br />

were performed in SV40-immortalized mouse cholangiocytes<br />

isolated from the large intrahepatic ducts (MLC) and in vivo<br />

<strong>studies</strong> were performed in a novel bile duct-cannulated murine<br />

model which we recently developed. This model involves placement<br />

of an indwelling catheter into the bile duct of a live mouse<br />

for measurement of bile flow rate and composition in response<br />

to systemic administration of agonists. [Ca2+]I was measured<br />

by Fura-2, ATP release by luciferase based assay and reported<br />

as arbitrary light units (ALUs), and Cl- currents by patch clamp<br />

techniques. Results: Exposure of MLC to UDCA (100 mM) rapidly<br />

increased [Ca2+]I (increase in Fura-2 fluorescence from<br />

0.65 to 1.05, p ≤0.01), stimulated ATP release (from 331 ± 22<br />

ALUs to 780 ± 106 ALUs, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 615A<br />

816<br />

The bile-acid receptor TGR5 regulates biliary epithelium<br />

permeability<br />

José Ursic-Bedoya 1,2 , Hayat Simerabet 1,2 , Isabelle Doignon 1,2 ,<br />

Noémie Péan 1,2 , Zahra Tanfin 1,2 , Christoph Ullmer 3 , Lydie Humbert<br />

4,5 , Dominique Rainteau 4,5 , Doris Cassio 1,2 , Thierry Tordjmann<br />

1,2 ; 1 U1174, INSERM, Orsay, France; 2 Université Paris sud,<br />

Orsay, France; 3 Pharma Research and Early Development, Roche<br />

Innovation Center Basel, Basel, Switzerland; 4 U1057, INSERM,<br />

Paris, France; 5 UPMC, Paris, France<br />

Background: TGR5, the bile acid (BA) G-protein-coupled receptor<br />

protects the liver against BA overload during regeneration.<br />

After partial hepatectomy, peribiliary hepatic necrosis occurs<br />

in TGR5 KO but not in wild type (WT) mice, suggesting that<br />

TGR5, highly expressed in biliary epithelial cells, regulates<br />

biliary epithelium permeability. Methods: Trans-epithelial resistance<br />

(TER) and FITC-dextran transfer were measured in the<br />

NRC (Normal Rat Cholangiocyte) cell line cultured on transwell<br />

inserts. To study biliary epithelial permeability in vivo in mice,<br />

fluorescent dextran or a modified BA (Glycocholic acid) were<br />

injected gallbladder (GB) lumen and traced (spectrofluorimetry<br />

& Mass Spectrometry) in plasma and liver. Cells and mice<br />

were stimulated with RO5527239, a TGR5 specific agonist,<br />

and with taurolitocholic acid. TGR5-induced signaling pathways<br />

were studied by western blot (WB), using selective inhibitors.<br />

Tight junction (TJ) proteins expression was investigated by<br />

qPCR, WB and immunofluorescence in NRC, in livers and GB<br />

from WT and TGR5-KO mice. Results: In NRC, TGR5 agonists<br />

significantly increased TER, reduced dextran passage, induced<br />

ERK phosphorylation and EGFR transactivation. Inhibition of<br />

cAMP production or MAPK pathways suppressed TGR5 agonists<br />

effect on NRC permeability. In TGR5-KO as compared<br />

with WT mice, although TJ proteins (ZO-1, occludin and JAM-<br />

A) mRNA expression in GB was significantly reduced, only<br />

JAM-A expression and localization at TJ were altered in bile<br />

ducts and GB. TGR5 agonists induced JAM-A phosphorylation<br />

and TJ localization in vitro in NRC, and in vivo after injection<br />

in WT but not in TGR5-KO GB lumen. In vivo, BA and dextran<br />

transepithelial transfer after GB injection was increased in<br />

TGR5-KO as compared with WT mice (Fig.1). BA transporters<br />

mRNA expression (ASBT, OSTb, MRP3) was similar in WT<br />

and TGR5-KO GB, suggesting that TGR5 controls paracellular<br />

rather than transcellular permeability. Conclusion: The BA<br />

receptor TGR5 reinforces biliary epithelial sealing in vitro and<br />

in vivo, likely through modulation of TJ protein expression and<br />

phosphorylation, protecting liver parenchyma against bile leakage.<br />

Disclosures:<br />

Christoph Ullmer - Employment: F. Hoffmann-La Roche AG<br />

The following authors have nothing to disclose: José Ursic-Bedoya, Hayat Simerabet,<br />

Isabelle Doignon, Noémie Péan, Zahra Tanfin, Lydie Humbert, Dominique<br />

Rainteau, Doris Cassio, Thierry Tordjmann<br />

817<br />

Knockout of the Secretin/Secretin Receptor Axis Delays<br />

Regeneration of small and large cholangiocytes by<br />

enhanced senescence following 70% partial hepatectomy<br />

(PH)<br />

Ying Wan 3 , Shannon S. Glaser 1 , Nan Wu 4 , Fanyin Meng 2 ,<br />

Julie Venter 4 , Romina Mancinelli 5 , Heather L. Francis 2 , Antonio<br />

Franchitto 5 , Paolo Onori 5 , Tina Kyritsi 4 , Shanika Avila 4 , Guido<br />

Carpino 7 , Holly A. Standeford 6 , Gianfranco Alpini 2 , Eugenio<br />

Gaudio 5 ; 1 Central Texas Veterans Health Care System and Texas<br />

A&M HSC College of Medicine, Temple, TX; 2 Research, S&W<br />

DDRC and Medicine, Veterans Health Care System and Texas<br />

A&M HSC and BaylorScott&White, Temple, TX; 3 Operational<br />

Funds, BaylorScott&White, Temple, TX; 4 Medicine, Texas A&M<br />

University HSC, Temple, TX; 5 Department of Anatomical, Histological,<br />

Forensic Medicine and Orthopedics Sciences, University La<br />

Sapienza, Rome, Italy; 6 Research, Central Texas Veterans Health<br />

Care System, Temple, TX; 7 Dept Health Sciences, University of<br />

Rome “Foro Italico, Rome, Italy<br />

Limited data exists regarding the neuroendocrine factors that<br />

regulate the renewal of the biliary tree after PH. The secretin<br />

(SCT)/secretin receptor (SR) axis is normally expressed only by<br />

large cholangiocytes. However, during the damage of large<br />

bile ducts small cholangiocytes (more resistant to injury) replenish<br />

the large damaged ducts by de novo acquisition of large<br />

biliary phenotypes. Cellular senescence is a state of irreversible<br />

cell cycle arrest involved in biliary diseases including primary<br />

sclerosing cholangitis and primary biliary cirrhosis. Large,<br />

senescent cholangiocytes are the target cells in the cholestatic<br />

models of bile duct ligation and MDR2 KO (PSC). Hypothesis:<br />

in hepatectomized SCT/SR double knockout (DKO) mice there<br />

is reduced regeneration of small and large bile ducts due to<br />

enhanced senescence. Methods: The <strong>studies</strong> were performed<br />

in normal wild type (WT) and SCT/SR DKO mice at 0, 3, 6<br />

hours and 1, 3, 7, and 14 days after sham or 70% PH. Liver<br />

injury was evaluated by: (i) H&E staining in liver sections; and<br />

(ii) measurement of serum levels of transaminases. We evaluated<br />

biliary proliferation by: (i) immunohistochemistry (IHC) for<br />

PCNA and intrahepatic bile duct mass (IBDM) by CK-19 IHC in<br />

liver sections; and (ii) qPCR for PCNA and CK-19 in total liver<br />

samples and small and large cholangiocytes. Cellular senescence<br />

was evaluated by SA-β-Gal staining in liver sections<br />

and qPCR for P18, PAI-1 and CCl2 in total liver samples and<br />

small and large cholangiocytes. Results: Small cholangiocytes<br />

from PH WT mice displayed less senescent features compared<br />

to large cholangiocytes. In PH WT mice, small less senescent<br />

cholangiocytes regenerate as early as 1 hr and rebuild IBDM<br />

within 1 day of PH, whereas large regenerate as early as 6 hr<br />

and rebuild IBDM at a lower rate (after 3-7 days of PH). Large<br />

cholangiocytes from DKO PH animals displayed increased<br />

expression of senescent markers relative to PH WT mice. In<br />

DKO PH mice, there was reduced regeneration time (by PCNA)<br />

and regrowth (by CK-19) of small and large bile ducts concomitant<br />

with enhanced expression of senescence markers (p18,<br />

PAI-1 and CCL2) in large cholangiocytes as well as in small<br />

cholangiocytes that display normally low levels of senescence.<br />

In SCT/SR double KO mice, at day 14 of PH IBDM returned at<br />

only 50% of the original biliary mass (0.238 ± 0.014, sham,<br />

vs. 0.128 ± 0.007, SCT/SR KO, p


616A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Ying Wan, Shannon S. Glaser,<br />

Nan Wu, Fanyin Meng, Julie Venter, Romina Mancinelli, Heather L. Francis,<br />

Antonio Franchitto, Paolo Onori, Tina Kyritsi, Shanika Avila, Guido Carpino,<br />

Holly A. Standeford, Gianfranco Alpini, Eugenio Gaudio<br />

818<br />

Pre-treatment with TIMP-3 prevents the development of<br />

biliary injury in an LPS enhanced ischaemia-reperfusion<br />

animal model<br />

Janske Reiling 1,2 , Kim Bridle 1,3 , Catherine M. Campbell 4 , Nishreen<br />

Santrampurwala 1,3 , Laurence J. Britton 1,3 , Ari J. Cohen 5 , Darrell H.<br />

Crawford 1,3 , Cornelis H. Dejong 2,6 , Jonathan Fawcett 1,7 ; 1 school<br />

of medicine, University of Queensland, Brisbane, QLD, Australia;<br />

2 NUTRIM School of Nutrition and Translational Research in Metabolism,<br />

Maastricht University, Maastricht, Netherlands; 3 Gallipoli<br />

Medical Reserach Foundation, Brisbane, QLD, Australia; 4 Envoi<br />

Specialist Pathologists, Brisbane, QLD, Australia; 5 Laboratory of<br />

Transplant Reseach, Institute of Translational Research, Ochsner<br />

Clinical Foundation, New Orleans, LA; 6 Department of Surgery,<br />

Maastricht University Medical Centre, Maastricht, Netherlands;<br />

7 Queensland Liver Transplant Service, Princess Alexandra Hospital,<br />

Brisbane, QLD, Australia<br />

Introduction Transplantation of donor livers retrieved after circulatory<br />

death is often complicated by the development of biliary<br />

injury and subsequent stricture formation. Our recent <strong>studies</strong><br />

identified a role for lipopolysaccharides (LPS) in the development<br />

of biliary injury. This study aimed to assess whether<br />

inhibition of TNF-α cleavage into its biologically active form,<br />

by pre-treatment with tissue inhibitor of metalloproteinases-3<br />

(TIMP-3), would prevent the development of biliary injury.<br />

Methods Male Sprague Dawley rats underwent either sham<br />

surgery (sham), 1000ng/kg human recombinant TIMP-3 one<br />

hour prior to sham surgery (TIMP-3), 70% partial liver ischaemia<br />

for thirty minutes with 1mg/kg LPS (combi) or TIMP-3<br />

followed by ischaemia and LPS (combi+TIMP). Following 6<br />

hours of reperfusion, blood, bile, liver and bile duct tissue was<br />

collected for analysis of liver function, biliary injury and expression<br />

of genes encoding for bile acid transporters and cytokines.<br />

Results TIMP-3 treatment prior to liver ischaemia and LPS administration<br />

resulted in a decreased bile duct injury severity score<br />

compared to non-pre-treated animals (sham: 0.00 (0.00-3.00);<br />

TIMP: 0.00 (0.00–4.00); combi: 4.50 (1.00-6.00); combi+TIMP:<br />

2.00 (0.00–5.00)). This was further supported by a<br />

significant decrease in lactate dehydrogenase in bile, a marker<br />

of biliary injury (combi: 15.78±9.46; combi+TIMP: 6.54±2.33<br />

U/L/ gram wet liver weight p=0.007). Despite improvement<br />

of biliary injury, hepatocellular injury remained unaltered as<br />

evidenced by increased serum alanine transaminase in both<br />

the combi and combi+TIMP groups (sham: 40.90±9.63;<br />

combi: 109.64±40.96; combi+TIMP: 114.63±84.23, p:<br />

0.03 and 0.024 respectively). Bile flow was normalised following<br />

TIMP pre-treatment (combi: 64.20±14.57; combi+LPS:<br />

85.37±10.99, p=0.012), as were gene expression levels of<br />

CYP27a1 and CYP7b1 (2 fold and 1,5 fold reduction, p=0.01<br />

and 0.02 respectively). This may reflect a normalisation of total<br />

bile acid synthetic capacity. Furthermore, gene expression of<br />

MIP-2, osteopontin and LBP were significantly down regulated<br />

following TIMP pre-treatment (8 fold, 6 fold and 2 fold respectively,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 617A<br />

oncocytic type were common. In contrast, group C and D were<br />

located in perihilar and distal bile ducts, and almost all cases<br />

were of high grade with invasion, and a majority of them were<br />

of intestinal and pancreatobiliary phenotypes. Group B was<br />

found along the biliary tree, and a majority of cases were of<br />

intestinal phenotype and all were of high grade with or without<br />

invasion. Conclusion: PNB with histopathologic similarities to<br />

IPMN and those without similarities to IPMN were different in<br />

their location and pathological behaviors. IPNB may be a heterogeneous<br />

group showing different biological behaviors and<br />

undergoing different tumorigenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yasuni Nakanuma, Yuko Kakuda,<br />

Katsuhiko Uesaka, Takashi Miyata, Yuki Fukumura, Masaru Takase<br />

Disclosures:<br />

The following authors have nothing to disclose: Natalie L. Berntsen, Bjarte Fosby,<br />

Corey Tan, Elisabeth Schrumpf, Tonje Bjoernetroe, Aksel Foss, Pål-Dag Line, Tom<br />

H. Karlsen, Richard S. Blumberg, Espen Melum<br />

820<br />

Intraductal papillary neoplasm of bile duct is a heterogeneous<br />

disease with respect to its histopathologic<br />

similarities to pancreatic intraductal papillary mucinous<br />

neoplasm<br />

Yasuni Nakanuma 1 , Yuko Kakuda 1 , Katsuhiko Uesaka 1 , Takashi<br />

Miyata 1 , Yuki Fukumura 2 , Masaru Takase 3 ; 1 Shizuoka Cancer<br />

Center, Sunto-Nagaizumi, Japan; 2 Department of Pathology, Juntendo<br />

University School of Medicine, Tokyo, Japan; 3 Department<br />

of Clinical Laboratory, Koshigawa City Hospital, Koshigaya, Japan<br />

Background and Aim: Intraductal papillary neoplasm of bile<br />

duct (IPNB) is characterized by papillary tumor covered by<br />

well-differentiated neoplastic epithelium with fine fibrovascular<br />

cores in the dilated bile ducts. IPNB reportedly resembles<br />

intraductal papillary mucinous neoplasm (IPMN) of pancreas<br />

to a varied degree, though the exact comparison remain to be<br />

clarified. Herein, we examined IPNB with respect to its histopathologic<br />

similarities to IPMN. Materials and Methods: A total<br />

of surgically resected 52 cases of IPNB and 42 cases of IPMN<br />

(with mural nodule) which met the WHO 2010 criteria of IPNB<br />

and IPMN were histologically, immunohistochemically and<br />

semiquantitatively compared with respect to the size, grade<br />

of intraepithelial neoplasm (low/intermediate, and high grade<br />

without invasion and with invasion), mucus hypersecretion,<br />

and phenotypes (intestinal, gastric, pancreatobiliary and oncocytic<br />

type) with a help of immunostaining of MUC1, MUC2,<br />

MUC5AC, MC6, CK7, CK20, CDX2 and CD10. This comparative<br />

study was done by two experts of biliary pathology and<br />

those of pancreatic pathology. Results: There were no differences<br />

in gender and age distributionof IPNB and IPMN (F:M<br />

35:17 in IPNB and 26:16 in IPMN, mean age 70.1 yrs in IPNB<br />

and 67.1 in IPMN). Mucus hypersecretion was more freuqent<br />

in IPMN (83%) than in IPNB (39%). Grades were different in<br />

IPNB and IPMN; low/intermediate (9.6% and 38.1%), high<br />

grade (25% and 23.8%) and high grade with invasion (65.4%<br />

and 38.1%). IPNB cases were classifiable into four groups with<br />

their histopathological similarities to IPMN: group A (identical<br />

to IPMN, 19 cases), group B (similar but a little different from<br />

IPMN, 18 cases), group C (focally or vaguely similar to IPMN,<br />

5 cases), and group D (different from IPMN, 10 cases). Group<br />

A was mainly found in the intrahepatic and perihilar bile ducts,<br />

and a majority of them were low/intermediate grade or high<br />

grade intraepithelial neoplasm without invasion. Gastric and<br />

821<br />

Wnt/β-catenin cooperates with Yap signaling to promote<br />

hepatobiliary repair in a model of chronic liver<br />

injury<br />

Kari Nejak-Bowen, Hirohisa Okabe, Satdarshan (Paul) S. Monga;<br />

University of Pittsburgh, Pittsburgh, PA<br />

Sustained cholestasis results in injury to the biliary epithelium<br />

with subsequent atypical ductular proliferation (ADP), which<br />

is thought to be a reparative mechanism. ADP is observed in<br />

animal models where biliary injury is the primary insult to the<br />

liver, such as 0.1% 3,5-diethoxycarbonyl-1,4-dihydro-collidine<br />

(DDC). Intriguingly, during cholestatic injury, small subsets of<br />

hepatocytes also begin to acquire a cholangiocyte-like phenotype.<br />

This phenomenon, known as transdifferentiation, or<br />

cellular reprogramming, may contribute to biliary repair by<br />

creating de novo channels for bile export. Because transgenic<br />

mice expressing a serine 45-mutated stable form of β-catenin<br />

(S45-TG) in liver have a significant number of hepatocytes<br />

expressing A6 (a biliary marker), after long-term exposure to<br />

DDC compared to wild-type (WT) littermates, we hypothesized<br />

that Wnt signaling plays a role in this process. To test this,<br />

WT animals on control and DDC diet were analyzed for Wnt<br />

expression. We found upregulation of Wnt 7a, 7b, and 10a<br />

– in the portal triad of DDC-fed mice, as determined by tissue<br />

microdissection. Cell sorting showed expression of these Wnts<br />

exclusively in the EpCAM+ cell compartment. Direct co-culture<br />

of Hek293 cells expressing Wnt7a significantly increased<br />

Sox-9 gene expression in AML12 cells and primary hepatocytes.<br />

Treatment with recombinant Wnt7A protein also induced<br />

Sox9 and EpCAM expression in AML12 cells. We further<br />

analyzed livers of WT and S45-TG mice on DDC diet and<br />

observed increased expression of biliary markers EpCAM and<br />

cytokeratin 19 (CK19) in S45-TG mice, which occurs as early<br />

as 28d after DDC. Concurrently, S45-TG hepatocytes in the<br />

periportal region also begin to express Sox9, a transcription<br />

factor normally restricted to bile ducts. Additionally, mice that<br />

lack Wnt signaling in liver have fewer A6-positive hepatocytes<br />

after bile duct ligation, further implicating Wnt signaling in this<br />

process. Finally, Yap activation, which regulates liver cell fate,<br />

was notably increased in hepatocyte nuclei of S45-TG livers,<br />

concomitant with increased expression of Yap target genes.<br />

Thus, Wnt/β-catenin signaling cooperates with the Yap pathway<br />

to regulate liver repair after biliary injuries.<br />

Disclosures:<br />

Satdarshan (Paul) S. Monga - Advisory Committees or Review Panels: Abbvie;<br />

Grant/Research Support: Dicerna Pharmaceuticals<br />

The following authors have nothing to disclose: Kari Nejak-Bowen, Hirohisa<br />

Okabe


618A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

822<br />

Prolonged usage of H1 or H2 histamine receptor antagonists<br />

decreases fibrosis and liver damage in MDR2 -<br />

/-<br />

mice: novel evidence of the therapeutic benefits of<br />

anti-histamines<br />

Hannah Jones 1 , Laura Hargrove 1 , Lindsey Kennedy 2 , Jennifer<br />

Owens 2 , Heather L. Francis 2,1 ; 1 Scott and White Memorial Hospital,<br />

Temple, TX; 2 Research, Central Texas Veterans Health Care<br />

System, Temple, TX<br />

Background: Primary Sclerosing Cholangitis (PSC) arises from<br />

damaged cholangiocytes, epithelial cells lining the biliary<br />

epithelium. The multi-drug resistant 2-gene knockout mouse<br />

(MDR2 -/- ) is similar to human PSC. Histamine increases biliary<br />

proliferation and intrahepatic ductal mass (IBDM) by<br />

interaction with histamine receptors (HRs). H1HR signals via<br />

increased PKC/Ca 2+ signaling, whereas H2HR induces effects<br />

via cAMP-dependent signaling. H1/H2 HR blockers like Claritin®<br />

and Zantac® are used to alleviate allergic reactions and<br />

heartburn. Our current study aimed to evaluate the effects of<br />

prolonged usage of H1 and H2 blockers on biliary proliferation,<br />

hepatic damage and fibrosis in MDR2 -/- mice. Methods:<br />

Wild-type (WT) and MDR2 -/- mice were treated by osmotic<br />

minipumps with 0.9% NaCl, fexofenadine (10mg/kg BW/<br />

day) or ranitidine (10mg/kg BW/day) for 4 weeks to recapitulate<br />

long-term usage. Liver blocks, serum and isolated cholangiocytes<br />

were obtained. Cholangiocyte proliferation and IBDM<br />

was evaluated by qPCR for PCNA and immunohistochemistry<br />

for CK-19 (bile duct marker). Fibrosis was detected by Masson’s<br />

Trichrome staining and by qPCR for the fibrotic markers,<br />

α-SMA, fibronectin, collagen type-1a and TGF-β1. In serum,<br />

ALT and AST levels were measured and histamine and TGF-β1<br />

by EIA. Necrosis, inflammation, and lobular damage were<br />

assessed by H&E. Biliary function was measured in purified<br />

cholangiocytes by evaluating intracellular IP 3<br />

and cAMP levels<br />

by EIA. In vitro, cholangiocytes were treated with fexofenadine<br />

or ranitidine (25 μM) for up to 72 hours before measuring<br />

biliary proliferation by MTS assay and PCNA immunoblots<br />

and the expression of the aforementioned fibrotic markers by<br />

qPCR. Results: In MDR2 -/- mice treated with H1/H2 HR blockers<br />

there was a significant decrease in liver damage, IBDM,<br />

proliferation and fibrosis compared to saline treated MDR2 -<br />

/-<br />

mice. AST and ALT levels were decreased in mice treated<br />

with HR blockers and histamine and TGF-β1 secretion was<br />

also reduced compared to saline treated MDR2 -/- mice. Also,<br />

IP 3<br />

and cAMP levels were decreased in cholangiocytes from<br />

MDR2 -/- mice treated with H1 or H2 blockers, respectively. No<br />

adverse effects on biliary proliferation, liver damage or fibrosis<br />

were found in WT mice treated with HR blockers. In vitro, stimulation<br />

with H1 or H2 blockers reduced biliary proliferation via<br />

IP 3<br />

/Ca 2+ or cAMP signaling, respectively. Conclusion: Chronic<br />

usage of H1 or H2 blockers has no adverse effect on WT mice,<br />

but reduces liver damage, biliary proliferation and fibrosis in<br />

MDR2 -/- mice. These drugs should be considered as a promising<br />

therapy for patients suffering from cholangiopathies such<br />

as PSC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hannah Jones, Laura Hargrove,<br />

Lindsey Kennedy, Jennifer Owens, Heather L. Francis<br />

823<br />

Hepatic CD4+ lymphocyte responses and initiation of<br />

biliary injury in mdr2 knockout mice depend on dendritic<br />

cell costimulation<br />

Celine S. Lages 1 , Tiffany Shi 1 , Paige Bolcas 1 , Julia Simmons 1 ,<br />

Avery Maddox 1 , Shiva K. Shanmukhappa 2 , Alexander G.<br />

Miethke 1 ; 1 Department of Gastroenterology, Cincinnati Children’s<br />

Hospital Medical Center, Cincinnati, OH; 2 Pathology and Laboratory<br />

Medicine, Cincinnati Children’s Hospital Medical Center,<br />

Cincinnati, OH<br />

T lymphocytes are implicated in initiation of cholangiopathies<br />

like biliary atresia or primary sclerosing cholangitis (PSC).<br />

CD11c+ dendritic cells (DCs) modulate expansion and activation<br />

of T lymphocytes through CD80/86-mediated costimulation.<br />

Blockade of CD86/CD28 at the immunological synapse<br />

with the CTLA4-IgG fusion protein was shown to improve autoimmune<br />

diseases. We hypothesize that DC costimulation is<br />

critical for T lymphocyte activation during the initiation of bile<br />

duct epithelial injury in mdr2-/- mice, a model for “toxic bile”<br />

induced sclerosing cholangitis. Methods: Kinetics of hepatic<br />

DC and T cell responses in male double-transgenic mdr2 -/-<br />

and CD11c-GFP mice were determined by flow cytometry. For<br />

costimulatory blockade, mice received i.p. injections of 20<br />

mcg/g of CTLA4-Ig every other day between day of life 7<br />

and 15. Plasma ALT levels and ductal proliferation on CK19-<br />

stained liver sections served as surrogate markers for hepatocellular<br />

and biliary injury, respectively. Results: Frequency of<br />

hepatic CD11c (GFP)+ DCs was increased in mdr2-/- mice<br />

at d8 and d14, but decreased thereafter (mean % GFP+MH-<br />

CII+B220-/Lin-7AAD- at d8: 0.46 vs 0.27, p=0.08; at d14:<br />

0.96 vs 0.46, p=0.02; at d30: 1.30 vs 2.95, p=0.03 in<br />

mdr2-/- vs mdr2+/+). The immunogenic subset of CD11b+<br />

myeloid mDCs followed a similar kinetic (d8: 0.22 vs 0.09,<br />

p=0.02; d14: 0.77 vs 0.12, p=0.001; d30: 1.12 vs 2.52,<br />

p=0.04). Frequency of DCs expressing the costimulatory molecules<br />

CD80 or CD86 were increased at d8 and d14, but<br />

decreased at d30. Treatment of mdr2-/- mice with CTLA4-Ig<br />

was associated with significantly decreased plasma ALT levels<br />

at d15 (mean 162.1 vs 465.1 IU/L in CTLA4-Ig vs non-treated,<br />

p=0.02) and reduced number of bile duct profiles per portal<br />

tract (mean 1.51 vs 2.11 in CTLA4-Ig vs hIgG, p=0.002).<br />

Administration of CTLA4-Ig was associated with reduced<br />

hepatic CD4+ T cell responses (mean % CD4+CD3+/7AAD-:<br />

2.37 vs 3.94, p=0.02), but did not affect expansion of CD8+<br />

T, NK or NKT cells. Typical for costimulatory blockade, differentiation<br />

of naïve to effector CD4+ T cells was attenuated<br />

(mean % of hepatic effector memory CD62L-CD44+/CD4:<br />

10.82 vs 60.90, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 619A<br />

824<br />

Tetrahydroxylated Bile Acids Reduce Cholangiocyte<br />

Damage in Abcb4 -/- mice<br />

Renxue Wang 1 , Jonathan Sheps 1 , Lin Liu 1 , Victor Ling 1,2 ; 1 BC Cancer<br />

Research Centre, Vancouver, BC, Canada; 2 Pathology, University<br />

of British Columbia, Vancouver, BC, Canada<br />

Background and Aim: Primary sclerosing cholangitis (PSC)<br />

is a chronic cholestatic disease of the liver and bile ducts,<br />

with nearly 50% of patients requiring liver transplantation and<br />

patients having a greatly increased risk of liver and GI cancers.<br />

There are no proven treatments for PSC and the synthetic<br />

bile acid, ursodeoxycholic acid (UDCA), is toxic and is not<br />

recommended. The best preclinical model of PSC is the Abcb4 -<br />

/-<br />

mouse which develops liver damage due to bile acid toxicity<br />

- cholangitis at 2 weeks after birth, gallstone formation and<br />

progressive fibrosis after 8 weeks, and extensive liver damage<br />

leading to hepatocellular carcinoma after 12 months. Tetrahydroxylated<br />

bile acids (THBAs) are a novel species of bile<br />

acids with properties that suggest they may be significantly<br />

less toxic and more effective than UDCA in treating cholestasis.<br />

The purpose of this study was to compare the ability of<br />

THBAs and UDCA to prevent liver damage in Abcb4 -/- mice.<br />

Methods: Three isoforms of THBA were synthesized and their<br />

choleretic ability was measured following bile duct cannulation<br />

experiments in mice. Progressive liver damage in Abcb4 -/- mice<br />

(assessment of liver enzymes, morphology and histology) was<br />

determined following dietary supplementation of either THBA<br />

or UDCA. Results: The three isoforms of THBA studied were<br />

shown to be significantly more hydrophilic than UDCA and<br />

other known bile acids, as demonstrated by the retention time<br />

on a reversed-phase UPLC column. However, in mice, THBAs<br />

and UDCA were equally as efficient at stimulating bile flow in<br />

a dose-dependent fashion. Dietary supplementation of either<br />

THBA or UDCA in Abcb4 -/- mice showed that THBA-fed mice<br />

had greatly reduced gallstone formation compared to UDCAfed<br />

mice. Assessment of liver chemistry and histology showed<br />

that THBA-fed Abcb4 -/- mice had less cholestatic pressure, cholangiocyte<br />

damage, and fibrosis than UDCA-fed and non-fed<br />

controls. Conclusion: THBAs are novel bile acids that function<br />

as efficient choleretic agents and are more hydrophilic (and<br />

presumably less toxic) than UDCA. Dietary supplementation<br />

<strong>studies</strong> in Abcb4 -/- mice showed that THBAs can prevent progressive<br />

liver damage under conditions where UDCA is ineffective.<br />

Thus, THBAs have the potential to be used in treating<br />

cholestatic and other liver diseases where a therapeutic agent<br />

of high choleretic efficiency and low toxicity may be advantageous.<br />

Disclosures: Victor Ling, Renxue Wang, Jonathan Sheps<br />

- Patents Held/Filed: Polyhydroxylated Bile Acids for the Treatment<br />

of Biliary Disorders.<br />

Disclosures:<br />

Renxue Wang - Board Membership: Qing Bile Therapeutics Inc<br />

Jonathan Sheps - Advisory Committees or Review Panels: Qing Bile Therapeutics<br />

Inc.; Grant/Research Support: Qing Bile Therapeutics Inc.; Patent Held/Filed:<br />

Qing Bile Therapeutics Inc.; Stock Shareholder: Qing Bile Therapeutics Inc.<br />

Lin Liu - Stock Shareholder: Qing Bile Therapeutic Inc.<br />

Victor Ling - Management Position: Qing Bile Therapeutics Inc<br />

825<br />

Role of Cellular Inhibitor of Apoptosis Proteins (cIAP) in<br />

Primary Sclerosing Cholangitis (PSC).<br />

Maria Eugenia Guicciardi, Anuradha Krishnan, Steven Bronk,<br />

Petra Hirsova, Gregory J. Gores; Mayo Clinic, Rochester, MN<br />

Primary sclerosing cholangitis (PSC) is a progressive cholestatic<br />

liver disease of unknown etiopathogenesis. Cellular inhibitor<br />

of apoptosis proteins (cIAP-1 and cIAP-2) are regulators<br />

of death receptor signaling, and death receptor signaling<br />

has been implicated in the pathogenesis of PSC. We have<br />

recently observed that cIAP protein levels are reduced in the<br />

interlobular bile ducts of human PSC livers. The AIM of this<br />

study was to investigate whether downregulation of cIAPs is<br />

linked to generation of a PSC-like phenotype. METHODS: Normal<br />

human cholangiocyte cell lines (NHC, H69), Receptor-Interacting<br />

Protein 1 (RIP1)- and NF-κB2-knock-down H69 cells,<br />

were treated with the IAP antagonist TL-32711 (1 mM). The<br />

resulting samples were analyzed by apoptosis assays, immunoblot<br />

and qPCR. In order to verify whether inhibition of cIAPs<br />

in the bile ducts is sufficient to produce a PSC-like phenotype,<br />

IAP antagonist BV6 (100 ml/mouse) was directly instilled into<br />

the biliary tree. Livers and serum was analyzed by standard<br />

methods. Cholangiograms were obtained using a biliary cast<br />

and imaging by microCT with 3-D reconstruction. RESULTS:<br />

TL-32711 treatment induced apoptosis in 12-15% of cholangiocytes<br />

after 24 hr. As previously reported, the transcription<br />

factor NF-κB was activated following TL-32711 treatment via<br />

the non-canonical, RIP1-independent pathway. Notably, IAP<br />

antagonism led to proinflammatory cytokine (IL-8, IL-6, IL-1β)<br />

production in cholangiocytes, which was significantly reduced<br />

in cells deficient in NF-κB2 (a mediator of the non-canonical<br />

pathway), but not RIP1, suggesting it was largely mediated by<br />

the non-canonical activation of NF-κB. Five days after biliary<br />

injection, PSC-like changes were observed in the BV6-treated<br />

mice, including a fibrous cholangiopathy of the interlobular<br />

bile ducts, portal inflammation, significant elevation of serum<br />

alkaline phosphatase, total bile acids and bilirubin, and cholangiographic<br />

evidence of intrahepatic biliary strictures and<br />

dilations, and damage of the small bile ducts. CONCLUSIONS:<br />

We have demonstrated that decrease in cIAP levels in cholangiocytes<br />

induces modest cholangiocyte death, and triggers a<br />

significant proinflammatory response mediated by the activation<br />

of NF-κB in vitro. More importantly, we have generated<br />

an acute murine model duplicating the fibrous cholangiopathy<br />

of PSC by directly injecting an IAP antagonist into the biliary<br />

tree, suggesting that downregulation of cIAPs in cholangiocytes<br />

may actively contribute to the development of PSC. These data<br />

suggest that new therapeutic strategies aimed to prevent the<br />

non-canonical activation of NF-κB may be beneficial in the<br />

treatment of PSC.<br />

Disclosures:<br />

Gregory J. Gores - Advisory Committees or Review Panels: Conatus<br />

The following authors have nothing to disclose: Maria Eugenia Guicciardi, Anuradha<br />

Krishnan, Steven Bronk, Petra Hirsova<br />

826<br />

CD39/ENTPD1 is protective in a model of biliary fibrosis<br />

and modulates CD8+ T cell gut-homing via α4β7 integrin<br />

Sonja Rothweiler, Zhen-Wei Peng, Susan B. Liu, Naoki Ikenaga,<br />

Yury Popov, Simon C. Robson; Gastroeneterology, Beth Israel Deaconess<br />

Medical Center, Boston, MA<br />

BACKGROUND: Primary sclerosing cholangitis (PSC) is a<br />

chronic cholestatic liver disease of unknown etiology characterized<br />

by inflammation and fibrosis of both intra- and<br />

extra-hepatic bile ducts. The ectonucleotidase CD39 (ENTPD1)<br />

regulates purinergic signaling by hydrolizing proinflammatory<br />

extracellular nucleotides, such as ATP and ADP, to ultimately<br />

generate immunosuppressive adenosine. There are strong associations<br />

between PSC and inflammatory bowel disease (IBD).<br />

Genetic polymorphisms of CD39 were found to be linked to<br />

IBD, indicating CD39 to be involved in the gut immune balance.<br />

We studied the impact of CD39-controlled purinergic signaling


620A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

on biliary fibrosis in the Mdr2-/- mouse model, which develops<br />

periportal fibrosis similar to human PSC. METHODS: Histology,<br />

inflammatory cell infiltration, gene expression, and collagen<br />

architecture and content were studied in livers from Mdr2-/-<br />

;CD39-/- and Mdr2-/- mice. Liver infiltrating lymphocytes were<br />

analyzed by flow cytometry. CD8-depleting antibody was<br />

administered to cholic acid challenged Mdr2-/-;CD39-/- mice.<br />

In vitro retinoic acid (RA)-induced gut-homing α4β7 expression<br />

was assessed in the presence of extracellular nucleotides or<br />

adenosine. RESULTS: Mdr2-/-;CD39-/- mice exhibited severe<br />

liver injury and significantly worse portal fibrosis when compared<br />

to Mdr2-/- mice. The Mdr2-/-;CD39-/- mice showed<br />

more pronounced elevations in serum ALT levels, increase in<br />

total collagen levels, as determined by hydroxyproline quantification,<br />

and upregulated pro-fibrogenic mRNA levels in the<br />

liver. Flow cytometry analysis of liver infiltrating lymphocytes<br />

revealed a selective increase in CD8+ T cells in CD39 deficient<br />

Mdr2-/- mice. Further, hepatic transcription of the b7 integrin, a<br />

T cell gut tropism marker, was significantly increased in Mdr2-<br />

/-;CD39-/- mice. Antibody-mediated CD8+ T cell depletion<br />

in Mdr2-/-;CD39-/- mice improved outcomes with decreased<br />

serum alkaline phosphatase and lower expression of liver fibrosis-related<br />

genes, supporting the pathogenetic role of CD8+ T<br />

cells in PSC. RA alone was able to induce the expression of the<br />

α4β7 gut-homing integrin in naïve CD8+ T cells. Interestingly,<br />

stimulation with ATP further enhanced the RA-mediated expression<br />

of α4β7, whereas adenosine demonstrated the opposing<br />

effect by downregulating integrin expression levels. CONCLU-<br />

SION: Genetic deletion of CD39 exacerbates biliary injury<br />

and fibrosis in Mdr2-/- mice. Disease progression occurs in a<br />

CD8+ T cell-dependent manner and extracellular nucleotides<br />

increase the expression of the gut-homing recpetor α4β7. Thus,<br />

the purinergic system provides possible pharmacological targets<br />

for the treatment of PSC.<br />

Disclosures:<br />

Yury Popov - Grant/Research Support: Gilead Sciences, Inc, Takeda<br />

Simon C. Robson - Grant/Research Support: Pfizer, NIH, Dainippon; Independent<br />

Contractor: Biolegend, EMD Millipore, Mersana; Management Position:<br />

eBioscience; Speaking and Teaching: ACP, Elsevier, ATC; Stock Shareholder:<br />

Nanopharma, Puretech<br />

The following authors have nothing to disclose: Sonja Rothweiler, Zhen-Wei<br />

Peng, Susan B. Liu, Naoki Ikenaga<br />

827<br />

Feedback regulation of autotaxin in mice: why cholestatic<br />

mice don’t scratch<br />

Ruth Bolier 1 , Dagmar Tolenaars 1 , Jacqueline Langedijk 1 , Piter J.<br />

Bosma 1 , Ulrich Beuers 1,2 , Ronald Oude Elferink 1 ; 1 Tytgat Institute<br />

for Liver and Intestinal Research, Academic Medical Center,<br />

Amsterdam, Netherlands; 2 Department of Gastroenterology and<br />

Hepatology, Academic Medical Center, Amsterdam, Netherlands<br />

Introduction: Serum activity of autotaxin (ATX, which produces<br />

lysophosphatidic acid, LPA) is increased in humans during<br />

pregnancy and cholestasis and correlates with itch intensity<br />

(Gastroenterology 2010;139:1008). In contrast, serum ATX in<br />

mice is only marginally increased under these conditions and,<br />

consistently, there is no increase in scratch behavior. Aim: To<br />

study the difference in regulation of ATX expression between<br />

humans and mice. Methods: Expression of LPA receptors<br />

(LPAR 1-6<br />

) and feedback repression of ATX expression trough<br />

LPA receptor signaling was studied in human and mouse fibroblasts<br />

by qPCR. LPA signaling was stimulated with the stable<br />

LPA-analogue XY17 (1uM) or via endogenous ATX production<br />

with lipid phosphate phosphatase 1 inhibitor XY14 (10uM), to<br />

prevent the rapid breakdown of LPA. LPA production by endogenous<br />

ATX was inhibited by ATX inhibitor HA115 (10uM) and<br />

LPAR 1-3<br />

signaling was blocked with the receptor antagonist<br />

Ki16425 (25uM). Involvement of G-proteins and their downstream<br />

pathways was investigated using the cAMP stimulator<br />

forskolin (25uM), protein kinase A inhibitor RP-adenosine<br />

(100uM), G α<br />

i/o-inhibitor pertussis toxin (500ng/mL), inositosol<br />

1,4,5-triphophate inhibitor U73122 (1uM) and protein<br />

kinase C inhibitor Gö6983 (10uM). In addition, the established<br />

induction of ATX by TNFα (10ng/mL) was compared<br />

in fibroblasts of the two species. Results: TNFα stimulated<br />

ATX expression 3- to 4-fold in fibroblasts from both species. In<br />

murine cells, stimulation of LPAR-signaling by XY14 and XY17<br />

repressed ATX mRNA expression by 70 and 82%, respectively<br />

(both p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 621A<br />

bated cleaved caspase-3 enzyme activity and overexpression<br />

of pJNK and RIP3 was characteristic of liver explants from<br />

PBC patients compared with healthy controls, accompanied<br />

by elevated protein levels of TNF and IL-6. Casp8 Δhepa after<br />

BDL displayed decreased number and size of necrotic foci<br />

compared with Casp8 f/f mice. Significantly decreased serum<br />

alanine (ALT) and aspartate (AST) transaminases, TUNEL<br />

staining, and cleaved Caspase-3 and cytochrome C protein<br />

overexpression were found in Casp8 Δhepa mice 28 days after<br />

BDL, along with diminished compensatory proliferation (Ki-67,<br />

PCNA) and ductular reaction (CK-19). These results were translated<br />

into a decreased inflammatory profile elicited by lower<br />

IL-6, TNF-α protein levels and reduced infiltration of F4/80 + ,<br />

Cd11b + and CD45 + populations. Overall, liver fibrogenesis<br />

(Sirius red, Collagen IA1, ASMA) was significantly ameliorated<br />

in Casp8 Δhepa . Mechanistically, decreased activation of<br />

JNK, RIP1 and RIP3 correlated with the protection shown after<br />

chronic BDL. Conclusions: CASP8 in hepatocytes is an essential<br />

orchestrator of cholestatic liver injury and its specific modulation<br />

might be an interesting pharmacologic target that can be<br />

used in the clinic as treatment for obstructive liver disease.<br />

Disclosures:<br />

Christian Trautwein - Grant/Research Support: BMS, Novartis, BMS, Novartis;<br />

Speaking and Teaching: Roche, BMS, Roche, BMS<br />

The following authors have nothing to disclose: Francisco Javier Cubero, Jin<br />

Peng, Maximilian Hatting, Gang Zhao, Miguel Eugenio Zoubek, Ricardo<br />

Macías-Rodríguez, Astrid Ruiz-Margáin, Johanna Reissing, Henning W. Zimmermann,<br />

Nikolaus Gassler, Tom Luedde, Christian Liedtke<br />

829<br />

NHERF-1 assembles macromolecular complexes in the<br />

liver that are essential for the inflammatory response in<br />

cholestasis<br />

Man Li 1 , Albert Mennone 1 , Carol J. Soroka 1 , Lee R. Hagey 3 ,<br />

Kathy M. Harry 1 , Edward J. Weinman 2 , James L. Boyer 1 ; 1 Internal<br />

Medicine, Yale University School of Medicine, New Haven, CT;<br />

2 University of Maryland School of Medicine, Baltimore, MD; 3 University<br />

of California, San Diego, La Jolla, CA<br />

The Na+/H+ exchanger regulatory factor 1 (NHERF-1/EBP50)<br />

is a PDZ protein that has been shown to participate in signal<br />

transduction by interacting with signaling molecules such as G<br />

protein-coupled receptors as well as regulatory proteins such<br />

as protein kinases. Our previous <strong>studies</strong> showed that deletion<br />

of NHERF-1 in mice leads to reduced hepatic expression of<br />

cytoskeletal ezrin-radixin-moesin (ERM) proteins and intercellular<br />

adhesion molecule-1 (ICAM-1), a molecule that plays a key<br />

role in neutrophil-mediated liver injury in mice after bile duct<br />

ligation (BDL). NHERF-1-/- BDL mice have significantly lower<br />

scores of hepatic necrosis, serum ALT, as well as reduced neutrophil<br />

accumulation in the liver compared with the wild-type<br />

(WT) BDL group. Aims: To further investigate the mechanisms<br />

of protection of liver injury induced by BDL in NHERF-1-/- mice.<br />

METHODS: Co-Immunoprecipitation was performed to identify<br />

protein complexes that are formed with NHERF-1 in the liver<br />

of sham and 7 day BDL WT mice. Immunoblotting was utilized<br />

to compare protein expression in sham and BDL WT and<br />

NHERF-1-/- mouse liver. Hepatic bile acids were analyzed<br />

by mass spectrometry as well. RESULTS: NHERF-1 co-immunoprecipitated<br />

with ICAM-1 in mouse liver, assembling ERM<br />

proteins, ICAM-1 and F-actin into a macromolecule complex<br />

that is increased in mouse liver and participates in neutrophil<br />

infiltration after BDL. In the liver, NHERF-1 also interacts with<br />

PKCζ, an atypical PKC that has been shown to phosphorylate<br />

and activate NF-κB p65. Compared with the WT control,<br />

NHERF-1-/- mice have reduced expression of hepatic PKCζ.<br />

Expression of total NF-κB p65 as well as phosphorylated NF-κB<br />

p65 was significantly reduced in the liver of NHERF-1-/- BDL<br />

mice compared with WT BDL mice. While total bile acid concentrations<br />

in the serum and liver of sham and BDL NHERF-1-<br />

/- mice were not significantly different from the WT controls,<br />

hepatic tetrahydroxylated bile acids and Cyp3a11 mRNA levels<br />

were significantly higher in NHERF-1-/- BDL mice, indicating<br />

a more profound adaptive response of the hepatocytes to<br />

BDL in these mice. CONCLUSIONS: NHERF-1 participates in the<br />

inflammatory and adaptive responses that is associated with<br />

BDL induced liver injury. Deletion of NHERF-1 in mice leads<br />

to disruption of the formation of ICAM-1-ERM-NHERF-1 and<br />

PKCζ/NHERF-1 complexes, causing reduction of hepatic ERM<br />

proteins, ICAM-1 and PKCζ, molecules that are essential for<br />

the inflammatory response in cholestasis. Further study of the<br />

role of NHERF-1 in the inflammatory response in cholestasis<br />

and other forms of liver injury should lead to discovery of new<br />

therapeutic targets in hepatic inflammatory diseases.<br />

Disclosures:<br />

James L. Boyer - Advisory Committees or Review Panels: Pfizer; Consulting:<br />

abbvie<br />

The following authors have nothing to disclose: Man Li, Albert Mennone, Carol J.<br />

Soroka, Lee R. Hagey, Kathy M. Harry, Edward J. Weinman<br />

830<br />

Knockdown of α7-Nicotinic Receptor Inhibits Biliary<br />

Proliferation and Hepatic Fibrosis during Extrahepatic<br />

Cholestasis<br />

Allyson Martinez 1 , April O’Brien 1 , Laurent Ehrlich 1 , David E. Dostal<br />

1 , Shannon S. Glaser 1,2 ; 1 Department of Medicine, Texas A&M<br />

Health Science Center and Central Texas Veterans Health Care<br />

System, Temple, TX; 2 Scott & White Digestive Disease Research<br />

Center, Temple, TX<br />

In cholestatic liver diseases, cholangiocytes, through the products<br />

of their cellular activation, are implicated as the key link<br />

between bile duct injury and the subepithelial fibrosis that<br />

characterizes chronic hepatobiliary injury. We have previously<br />

demonstrated that activation of α7-nicotine receptor (nAChR)<br />

with nicotine stimulates cholangiocyte proliferation that was<br />

associated with increased profibrotic gene expression by cholangiocytes<br />

and deposition of collagen in portal areas in normal<br />

wild-type mice. We have also demonstrated that mechanical<br />

stress, which occurs due to increased biliary pressure during<br />

extrahepatic cholestasis, stimulates cholangiocyte proliferation.<br />

The role of the α7-nAChR during extrahepatic cholestasis or<br />

mechanical stress has not been thoroughly explored. Thus, the<br />

AIM of our study was to determine the role of α7-nAChR in the<br />

regulation of biliary proliferation and hepatic fibrosis during<br />

extrahepatic cholestasis induced by bile duct ligation (BDL).<br />

Methods: Studies were performed in normal and BDL (1 week)<br />

wild-type (WT) and α7-nAChR knockout (KO) mice. Intrahepatic<br />

bile duct mass (IBDM) and biliary proliferation was evaluated<br />

by immunohistochemistry for CK-19 and PCNA in liver<br />

sections and by immunoblots in isolated cholangiocytes. Liver<br />

fibrosis was evaluated by Sirius red staining in liver sections<br />

and by qPCR for the profibrotic markers [collagen 1 (COL1A1),<br />

fibronectin (FN-1), and αSMA] in isolated cholangiocytes and<br />

total liver samples. In vitro, mouse primary cholangiocytes were<br />

plated on BioFlex culture plates and static equiaxial strain was<br />

applied to the cells for 24 hrs in the presence and absence of<br />

shRNA to knockdown α7-nAChR and the α7-nAChR antagonist<br />

methyllycaconitine (1.4 nM) . Proliferation was evaluated by<br />

immunoblots for PCNA and the expression of COL1A1, FN-1<br />

and αSMA by qPCR. Results: In vivo, there was a significant<br />

reduction in biliary proliferation (PCNA) and IBDM (CK-19)<br />

in BDL α7-nAChR KO mice compared to BDL WT. In addi-


622A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tion to a reduction in proliferation, there was also a reduction<br />

in hepatic fibrosis by Sirius red staining and a significant<br />

reduction in COL1A1, FN-1 and αSMA gene expression. In<br />

vitro, knockdown or inhibition of α7-nAChR in mouse cholangiocytes<br />

prevented mechanical stress-induced proliferation.<br />

Knockdown or inhibition of α7-nAChR also inhibited mechanical<br />

stress-induced COL1A1, FN-1 and αSMA gene expression.<br />

Conclusion: The knockout of α7-nAChR inhibits BDL induced<br />

proliferation and hepatic fibrosis. Modulation of the α7-nAChR<br />

axis during extrahepatic cholestasis may represent a novel therapeutic<br />

approach for cholestatic liver diseases.<br />

Disclosures:<br />

The following authors have nothing to disclose: Allyson Martinez, April O’Brien,<br />

Laurent Ehrlich, David E. Dostal, Shannon S. Glaser<br />

831<br />

Keratin 23 represents a novel liver injury marker reflecting<br />

the severity of ductular reaction<br />

Nurdan Guldiken 1 , Gokce Kobazi Ensari 1 , Pooja Lahiri 2 , Christian<br />

Liedtke 1 , Henning W. Zimmermann 1 , Christian Trautwein 1 ,<br />

Marianne Ziol 3 , Pavel Strnad 1 ; 1 Internal Medicine III, Aachen,<br />

Germany; 2 Institut für Pathologie, Graz, Austria; 3 Pathology<br />

Department, Paris, France<br />

Background: Keratins (K) are the intermediate filaments of epithelial<br />

cells and constitute established diagnostic tools. In the<br />

liver, K7/K19 expression is restricted to hepatic progenitor<br />

cells (HPCs), intermediate (i.e. not fully differentiated) hepatocytes<br />

and biliary epithelial cells (BECs). Consequently, K7/<br />

K19 represent a widely used marker of the regenerative liver<br />

response termed ductular reaction (DR) that consists of activated<br />

BECs and HPCs. Methods: Given that K23 is a largely<br />

unknown keratin family member, we analysed its expression<br />

and localization in selected human liver disorders (8 controls,<br />

15 patients with simple alcoholic steatosis [ALD], 9 with<br />

chronic hepatitis B/ 13 with chronic hepatitis C (HCV), 12 with<br />

non-alcoholic steatohepatitis; 7 with acute liver failure [ALF];<br />

7 with end-stage primary biliary cirrhosis [PBC]) and mouse<br />

liver injury models using custom-made antibodies. Results: In<br />

untreated mice, K23 was found in biliary epithelia but not<br />

hepatocytes. It was (together with K7/K19) markedly up-regulated<br />

in three different DR/cholestatic injury models, i.e. four<br />

months old MDR2 knockouts, animals treated with 3,5-diethoxycarbonyl-1,4-dihydrocollidine<br />

for four weeks or subjected to<br />

bile duct ligation for five days. No changes in K23 levels were<br />

seen in hepatocellular injury models such as partial hepatectomy<br />

or carbon tetrachloride-induced fibrogenesis. K23 levels<br />

correlated with the DR marker Fn14 and immunofluorescence<br />

staining showed a distinct but not perfect co-localization with<br />

K7/K19. In humans, K23 levels were moderately up-regulated<br />

in active HCV (~3 times) and ALD (~10 times). K23 expression<br />

was higher in patients with more prominent inflammation/fibrosis.<br />

A dramatic increase (>200 times) was observed in patients<br />

with ALF and end-stage PBC. In ALF/PBC, K7/K23 positive,<br />

K19 negative cells corresponding to intermediate hepatocytes<br />

were frequently found. In conclusion, K23 represents a novel<br />

DR marker and its levels correlate with the severity of the underlying<br />

human disease.<br />

Disclosures:<br />

Christian Trautwein - Grant/Research Support: BMS, Novartis, BMS, Novartis;<br />

Speaking and Teaching: Roche, BMS, Roche, BMS<br />

The following authors have nothing to disclose: Nurdan Guldiken, Gokce Kobazi<br />

Ensari, Pooja Lahiri, Christian Liedtke, Henning W. Zimmermann, Marianne Ziol,<br />

Pavel Strnad<br />

832<br />

RIP3-dependent signalling contributes to necroinflammation<br />

in cholestatic liver injury<br />

Marta B. Afonso 2 , Pedro M. Rodrigues 2 , André L. Simão 2 , Helena<br />

Cortez-Pinto 3,4 , Dimitry Ofengeim 5 , Joana D. Amaral 1 , Rui E.<br />

Castro 1 , Junying Yuan 5 , Cecília M. Rodrigues 1 ; 1 iMed.ULisboa &<br />

Dep. of Biochemistry and Human Biology, Faculty of Pharmacy,<br />

University of Lisbon, Lisboa, Portugal; 2 iMed.ULisboa, Faculty of<br />

Pharmacy, University of Lisbon, Lisbon, Portugal; 3 Gastrenterology,<br />

Hospital de Santa Maria, Lisbon, Portugal; 4 Faculty of Medicine,<br />

University of Lisbon, Lisbon, Portugal; 5 Department of Cell Biology,<br />

Harvard Medical School, Boston, MA<br />

Cholestasis is a common pathological condition characterized<br />

by retention of bile acids in liver and serum, with concomitant<br />

liver inflammation and hepatocyte damage. Recent evidence<br />

suggests that regulated necrosis or necroptosis, which depends<br />

on the kinase activity of receptor interacting protein 3 (RIP3),<br />

may contribute to the pathogenesis of inflammation-driven<br />

liver diseases. Here we aimed to evaluate the role of necroptosis<br />

after common bile duct ligation (BDL) in mice, a classic<br />

experimental model for acute cholestasis and secondary<br />

biliary fibrosis, and in patients with primary biliary cirrhosis<br />

(PBC), a cholestatic chronic liver disease. RIP3 and its target<br />

mixed lineage kinase domain-like protein (MLKL), were evaluated<br />

by immunohistochemistry in liver specimens of patients<br />

with PBC and healthy volunteers. BDL or sham surgery was<br />

performed in C57BL/6 wild-type (WT) or RIP3-deficient (RIP3 -<br />

/- ) mice. Serum and livers were collected 3 and 14 days after<br />

BDL. Liver histology and serum liver enzymes were evaluated.<br />

Liver RIP3, proinflammatory and fibrotic markers were determined<br />

by qRT-PCR. Total and soluble/insoluble liver proteins<br />

were analysed by Western blot. RIP3 kinase activity was determined<br />

in vitro. Finally, the functional crosstalk between RIP3<br />

and antioxidant response was investigated. RIP3 and phosphorylated<br />

MLKL protein levels were upregulated in the liver of<br />

PBC patients (p < 0.05). In WT mouse livers, BDL resulted in<br />

progressive bile duct hyperplasia, multifocal necrosis, fibrosis<br />

and inflammatory cell infiltration. Concomitantly, necroptosis<br />

was activated as assessed by: 1) increased RIP3 mRNA and<br />

protein expression; 2) augmented RIP3 kinase activity, and 3)<br />

sequestration of RIP3 and MLKL in the insoluble protein fraction<br />

of the liver (at least p < 0.05). Remarkably, histological analysis<br />

revealed that absence of RIP3 significantly decreased liver<br />

necrosis and inflammation induced by BDL at both time-points<br />

(p < 0.05). Accordingly, at day 3, BDL RIP3 -/- mice showed<br />

lower circulating levels of hepatic enzymes and decreased<br />

inflammatory and fibrogenic liver gene expression, together<br />

with enhanced antioxidant response (at least, p < 0.05). Protection<br />

was no longer evident at 14 days after BDL in RIP3 -/- .<br />

In conclusion, RIP3-dependent signaling is triggered in human<br />

PBC and mediates hepatic necroinflammation in BDL-induced<br />

acute cholestasis. Targeting necroptosis may be a promising<br />

therapeutic approach to treat cholestatic liver injury, although<br />

complementary approaches may be required to control progression<br />

to fibrosis.<br />

Disclosures:<br />

Helena Cortez-Pinto - Advisory Committees or Review Panels: Norgine, Lundbeck;<br />

Speaking and Teaching: Janssen, Gilead Janssen<br />

The following authors have nothing to disclose: Marta B. Afonso, Pedro M.<br />

Rodrigues, André L. Simão, Dimitry Ofengeim, Joana D. Amaral, Rui E. Castro,<br />

Junying Yuan, Cecília M. Rodrigues


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 623A<br />

833<br />

Gut microbiota contributes to disease in a spontaneous<br />

mouse model of chronic cholangitis<br />

Elisabeth Schrumpf 1,2 , Martin Kummen 1,2 , Kristian Holm 1,2 , Tom<br />

H. Karlsen 1,2 , Johannes R. Hov 1,2 , Richard S. Blumberg 3 , Espen<br />

Melum 1,2 ; 1 Norwegian PSC Research Center, Division of Cancer,<br />

Surgery and Transplantation, Oslo University Hospital, Rikshospitalet,<br />

Oslo, Norway; 2 K.G. Jebsen Inflammation Research Centre,<br />

Research Institute of Internal Medicine, Oslo University Hospital,<br />

Rikshospitalet, Oslo, Norway; 3 Division of Gastroenterology,<br />

Hepatology and Endoscopy, Department of Medicine, Brigham<br />

and Women’s Hospital, Harvard Medical School, Boston, MA<br />

Purpose: NOD.c3c4 mice are on a NOD background and<br />

spontaneously develop biliary inflammation both in extra- and<br />

intra-hepatic bile ducts. Gut microbes and lymphocytes activated<br />

in the intestines are hypothesized to play a role in the<br />

human biliary disease primary sclerosing cholangitis. We<br />

aimed to clarify the role of the gut microbiota in NOD.c3c4<br />

mice. Methods: We sampled cecal content and mucosa from<br />

10 week old NOD.c3c4 (n=5) and NOD mice (n=5) housed<br />

in a minimal disease unit (MDU) facility. The NOD.c3c4 and<br />

NOD strains were rederived into a new MDU facility, and<br />

after three generations the experiment was repeated. DNA<br />

was extracted and the V4 region of the 16S rRNA gene was<br />

amplified and sequenced on the Illumina MiSeq. Data were<br />

analysed with Quantitative Insights Into Microbial Ecology<br />

(QIIME). NOD.c3c4 mice were also rederived into a germ free<br />

(GF) facility. Conventionally housed (n=9) and GF NOD.c3c4<br />

mice (n=9) were harvested at 9 weeks of age, with sampling of<br />

the liver and blood. Results: The gut microbial profiles of mice<br />

with and without biliary disease were different both before and<br />

after rederivation (beta-diversity, unweighted UniFrac-distance,<br />

anosim: p


624A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

835<br />

Knockdown of the melatonin receptor, MT1, reduces<br />

biliary hyperplasia and liver fibrosis in cholestatic bile<br />

duct ligated (BDL) mice<br />

Nan Wu 2 , Fanyin Meng 1 , Ying Wan 3 , Julie Venter 2 , Holly A. Standeford<br />

4 , Heather L. Francis 1 , Lindsey Kennedy 4 , Yuyan Han 1,2 ,<br />

Kelly McDaniel 3 , Antonio Franchitto 5 , Paolo Onori 5 , Sharon<br />

DeMorrow 1 , Tami Annable 3 , Eugenio Gaudio 5 , Shannon S. Glaser<br />

1 , Gianfranco Alpini 1 ; 1 Research, S&W DDRC and Medicine,<br />

Central Texas Veterans Health Care System, Scott & White and<br />

Texas A & M Health Science Center College of Medicine and<br />

Scott White, Temple, TX; 2 Medicine, Texas A&M University HSC,<br />

Temple, TX; 3 Operational Funds, BaylorScott&White, Temple, TX;<br />

4 Research, Central Texas Veterans Health Care System, Temple,<br />

TX; 5 Department of Anatomical, Histological, Forensic Medicine<br />

and Orthopedics Sciences, University La Sapienza, Rome, Italy<br />

Cholestatic liver diseases are characterized by increased biliary<br />

proliferation and intrahepatic bile duct mass (IBDM), liver<br />

damage and fibrosis. Melatonin, which is synthesized in the<br />

pineal gland as well as cholangiocytes, exerts its effects by<br />

interaction with MT1 and MT2 receptors. Knockdown of MT2<br />

increases biliary proliferation and liver fibrosis in cholestatic<br />

mice. Also, melatonin inhibits biliary hyperplasia by interacting<br />

with MT1 receptors. However, since these <strong>studies</strong> were<br />

performed only in vitro, we evaluated the role of MT1 in regulating<br />

biliary proliferation and liver fibrosis in vivo in MT1<br />

knockout (KO) mice, and wild-type (WT) mice in which the<br />

hepatic expression of MT1 was reduced by Vivo-Morpholinos<br />

for MT1. Methods: The <strong>studies</strong> were performed in normal (WT),<br />

MT1 Morpholino-treated or MT1 KO mice with or without BDL<br />

surgery. We evaluated the expression of MT1 in liver sections<br />

by immunohistochemistry (IHC) and immunoblots in cholangiocytes.<br />

Intrahepatic biliary ductal mass (IBDM) and biliary proliferation<br />

was evaluated by IHC for CK-19 and PCNA in liver<br />

sections, and qPCR for PCNA in cholangiocytes. Liver injury<br />

was determined by H&E staining in liver sections and serum<br />

levels of transaminases and bilirubin. Fibrosis was evaluated<br />

by Sirius red staining in liver sections and qPCR for TGFbeta1,<br />

alpha-SMA and fibronectin in cholangiocytes and total liver. In<br />

vitro, shRNA-MT1 transfected murine cholangiocytes (MCCs)<br />

and controls were used to measure proliferation and TGFbeta1,<br />

alpha-SMA and fibronectin (FN-1) mRNA expression by qPCR.<br />

Results: Normal cholangiocytes express low levels of MT1,<br />

which markedly increased after BDL. We demonstrated that in<br />

MT1 KO BDL and MT1 Morpholino-treated BDL mice, along<br />

with reduced MT1 expression, there was decreased biliary<br />

proliferation and IBDM. In MT1 KO BDL and MT1 Morpholino-treated<br />

BDL mice, there was decreased lobular damage,<br />

serum levels of transaminases and bilirubin, and liver fibrosis in<br />

liver sections and reduced expression of TGFbeta1, alpha-SMA<br />

and FN-1 compared to BDL WT mice. In shRNA-MT1 transfected<br />

MCCs, there was decreased proliferation and reduced<br />

expression of TGFb1, alpha-SMA and FN-1. The findings indicate<br />

that shifting the melatonin axis to signal via MT2 during<br />

the knockdown of MT1 expression decreases biliary proliferation<br />

and liver fibrosis. Conclusion: Our findings indicate that<br />

the balance of MT1/MT2 receptor expression for the melatonin<br />

signaling axis may play an important role in the regulation of<br />

biliary proliferation and hepatic fibrosis. Inhibition of MT1 may<br />

be a key approach for ameliorating biliary hyperplasia and<br />

liver fibrosis in cholestatic liver diseases.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nan Wu, Fanyin Meng, Ying<br />

Wan, Julie Venter, Holly A. Standeford, Heather L. Francis, Lindsey Kennedy,<br />

Yuyan Han, Kelly McDaniel, Antonio Franchitto, Paolo Onori, Sharon DeMorrow,<br />

Tami Annable, Eugenio Gaudio, Shannon S. Glaser, Gianfranco Alpini<br />

836<br />

Longer Term Efficacy of Simultaneous Liver Transplantation<br />

plus Sleeve Gastrectomy versus non-invasive<br />

Weight Loss followed by Liver Transplantation<br />

Daniel Zamora-Valdes 1 , Kymberly D. Watt 2 , John J. Poterucha 2 ,<br />

Charles B. Rosen 1 , Todd A. Kellogg 3 , Sara R. Di Cecco 2 , Nicki<br />

M. Francisco Ziller 2 , Julie Heimbach 1 ; 1 Transplant Surgery, Mayo<br />

Clinic, Rochester, MN; 2 Liver Transplant, Mayo Clinic, Rochester,<br />

MN; 3 General Surgery, Mayo Clinic, Rochester, MN<br />

Obesity is increasingly common before and after liver transplantation<br />

(LT), yet optimal management remains unclear. Our<br />

AIM was to analyze the long-term effectiveness of a multi-disciplinary<br />

protocol for obese patients requiring LT, including<br />

a non-invasive pre-transplant weight loss program, and combined<br />

LT plus sleeve gastrectomy (SG) for obese patients who<br />

failed to lose weight prior to LT. METHODS: since 2006, all<br />

patients referred for LT with a BMI >35 were enrolled. There<br />

are 46 patients who achieved weight loss and underwent LT<br />

alone, and 17 who underwent LT combined with SG. Outcomes<br />

for patients with > 2 years since LT alone or LT+SG<br />

were analyzed using Student T-test for continuos variables and<br />

c 2 for categorical variables. RESULTS: In those who received<br />

LT alone, 34/46 were successful on achieving >10% loss in<br />

total body weight (TBW) prior to LT. Of those with > 2 years<br />

since LT, weight gain to BMI>35 was seen in 21/34 (61.7%)<br />

at 2 years (Figure 1), though 10/34 (30%) who maintained<br />

>10% loss in TBW at 2 years (p=0.005; Figure 1). There were<br />

8/34 (23.5%) with diabetes post LT and 10/34 (18.9%) with<br />

evidence of steatosis on year 2 ultrasound. In patients undergoing<br />

the LT+SG, all patients achieved significant weight loss<br />

(mean BMI at listing 49.8±1.5 vs. 2 year post LT 29.3±6.4;<br />

p=0.001). No LT+SG patient developed post-LT DM or steatosis<br />

at 2 years post LT. Conclusion: Non-invasive pre-transplant<br />

weight loss was achieved by a majority, though weight gain<br />

post LT was common. Combined LT plus SG resulted in effective<br />

weight loss and was associated with fewer post-LT metabolic<br />

complications at 2 years post LT.<br />

Figure 1: post transplant BMI in morbidly obese patients who<br />

underwent medical management (n=34) or sleeve gastrectomy<br />

(n=5) combined with liver transplantation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Daniel Zamora-Valdes, Kymberly<br />

D. Watt, John J. Poterucha, Charles B. Rosen, Todd A. Kellogg, Sara R. Di<br />

Cecco, Nicki M. Francisco Ziller, Julie Heimbach


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 625A<br />

837<br />

Preliminary Data Analysis from the Improving DCD Outcomes<br />

in Liver Transplantation (IDOL) Consortium<br />

David S. Goldberg 1 , Seth J. Karp 3 , Julie Heimbach 5 , James F.<br />

Markmann 6 , Richard Gilroy 4 , Roberto Hernandez-Alejandro 7 ,<br />

Peter L. Abt 2 ; 1 Division of Gastroenterology, Hospital of the University<br />

of Pennsylvania, Philadelphia, PA; 2 Department of Surgery,<br />

University of Pennsylvania, Philadelphia, PA; 3 Department<br />

of Surgery, Vanderbilt University, Nashville, TN; 4 Department of<br />

Medicine, University of Kansas, Kansas City, KS; 5 Department of<br />

Surgery, Mayo Clinic, Rochest, MN; 6 Department of Surgery, Massachusetts<br />

General Hospital, Boston, MA; 7 Department of Surgery,<br />

London Health Sciences Center, London, ON, Canada<br />

Introduction: Increased utilization of donation after cardiac<br />

death (DCD) liver allografts could help meet the demand for<br />

transplantable livers. Current understanding of post-liver transplant<br />

(LT) outcomes in DCD recipients is limited to single-center<br />

reports, or national analyses of UNOS data that do not capture<br />

detailed biliary complication data. We created the IDOL<br />

Consortium to collect multi-center patient-level data to: 1) overcome<br />

limitations of prior DCD analyses; 2) validate previous<br />

single-center <strong>studies</strong>; and 3) identify novel factors associated<br />

with DCD outcomes. Methods: The consortium includes 10 US<br />

centers and 2 centers in Canada. Detailed patient-level data<br />

were entered into a Redcap database. Donor and recipient<br />

age, and donor warm ischemic time (WIT; extubation to crossclamp)<br />

were evaluated as primary exposures, with primary<br />

outcomes of biliary complications (BC; excluding bile leak)<br />

and graft failure within 6 months post-DCD transplantation.<br />

Results: To date, 171 DCD recipients have been entered into<br />

the IDOL Consortium database. The median recipient age<br />

was 57 years (IQR: 50-61), 117 (68.4%) were male, and the<br />

median lab MELD was 19 (IQR: 13-24). The median donor<br />

WIT was 25 minutes (IQR: 20-30). 41 (24.0%) patients had<br />

a BC within the first 6 months post-LT: 18 (43.9%) with diffuse<br />

intrahepatic strictures (ischemic cholangiopathy), 14 (34.1%)<br />

anastomotic strictures, 7 (17.1%) with ischemic intrahepatic<br />

strictures and an anastomotic stricture, and 2 (4.9%) had ischemic<br />

extra-hepatic non-anastomotic strictures. There were 21<br />

(12.3%) graft failures, and 18 (11.2%) deaths within the first<br />

6 months post-LT. Of these 21 graft failures, 4 (19.0%) had a<br />

history of a BC. Donor age >40 years of age was numerically,<br />

but not statistically, associated with increased risk of BCs (OR:<br />

1.86, 95% CI: 0.91-3.77, p=0.09), while a donor WIT>30<br />

minutes was associated with significantly increased risk of BCs<br />

(OR: 2.41, 95% CIL 1.06-5.47, p=0.04). Increase recipient<br />

age was associated with increased risk of graft failure (OR<br />

for every 5-year increase: 1.49, 95% CI: 1.06-2.10), while<br />

donor WIT>30 minutes was numerically, but not statistically<br />

associated with graft failure (OR: 2.48, 95% CI: 0.93-6.64,<br />

p=0.07). Laboratory MELD score was not associated with BC<br />

or graft failure. Discussion: Preliminary data from the IDOL<br />

Consortium provide estimates of BCs among a diverse cohort<br />

of DCD recipients, and identify potential donor and recipient<br />

factors associated with BCs and graft failure. Completion of<br />

the IDOL Consortium data collection is anticipated in the next 3<br />

months, which will allow for sufficient sample sizes to estimate<br />

all key exposures and outcomes.<br />

Disclosures:<br />

Richard Gilroy - Speaking and Teaching: Salix, NPS, Gilead, AbbVie, Novartis<br />

The following authors have nothing to disclose: David S. Goldberg, Seth J. Karp,<br />

Julie Heimbach, James F. Markmann, Roberto Hernandez-Alejandro, Peter L. Abt<br />

838<br />

A shift to the left: A comparison of the A2ALL experience<br />

using left lobe vs. right lobe grafts<br />

Benjamin Samstein 1 , Abigail Smith 9 , Talia B. Baker 8 , Chris<br />

Freise 12 , Jean C. Emond 2 , Brenda W. Gillespie 9 , David Grant 4 ,<br />

Elizabeth A. Pomfret 11 , Adrian Cotterell 3 , Trevor L. Nydam 6 ,<br />

Kim M. Olthoff 7 , Abhinav Humar 10 , Robert M. Merion 5 ; 1 Department<br />

of Surgery, Weill Cornell Medical Center, New York, NY;<br />

2 Columbia University, New York, NY; 3 Virginia Commonwealth<br />

University, Richmond, VA; 4 University of Toronto, Toronto, ON,<br />

Canada; 5 Arbor Research Collaborative for Health and University<br />

of Michigan Department of Surgery, Ann Arbor, MI; 6 Department<br />

of Surgery, University of Colorado School of Medicine, Aurora,<br />

CO; 7 University of Pennsylvania, Philadelphia, PA; 8 Comprehensive<br />

Transplant Center, Northwestern University Feinberg School<br />

of Medicine, Chicago, IL; 9 University of Michigan, Ann Arbor, MI;<br />

10 University of Pittsburgh, Pittsburgh, PA; 11 Lahey Hospital and<br />

Medical Center, Burlington, ME; 12 Transplant Division, University<br />

of California, San Francisco, San Francisco, CA<br />

Exploration of the use of left lobe grafts was a principal aim<br />

of A2ALL. Potential living donor liver transplant (LDLT) recipients<br />

and their donors were enrolled at 12 North American<br />

centers between 2004 and 2014 with transplants occurring<br />

between 1998 and 2014. Of the 1141 donors and 1136<br />

recipients enrolled 99 were left lobe donors and 88 were left<br />

lobe recipients, respectively. Preoperative characteristics, intraoperative<br />

and postoperative data including patient and graft<br />

survival were compared by lobe using t-tests, chi-square tests,<br />

Wilcoxon tests, and Cox models. Characteristics of graft type<br />

and outcomes were analyzed over time and between centers.<br />

There was a marked center effect in the use of left lobes, with<br />

two centers accounting for 73% and 71% of left lobe recipients<br />

and donors, respectively. Right and left lobe donors did not<br />

differ significantly with respect to clinical characteristics. Left<br />

lobe donor surgeries were more likely to be laparoscopic (55%<br />

vs. 31%, p=0.003). Postoperatively, left lobe donors had lower<br />

bilirubin (0.9±0.6 vs. 1.6±1.3, p


626A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

fact that all obese patients remained obese post-OLT without<br />

a significant difference in BMI pre- and post-OLT regardless<br />

of their pre-OLT category. In conclusion, our findings point to<br />

obesity being a major risk of NAFLD recurrence in liver graft<br />

post-OLT.<br />

Disclosures:<br />

Roberto J. Firpi - Advisory Committees or Review Panels: Quest Diagnostics,<br />

Gilead; Grant/Research Support: Bayer, Vertex, BMS, Janssen, Gilead, Merck<br />

The following authors have nothing to disclose: Angela Dolganiuc, Vikas Khullar,<br />

Virginia C. Clark<br />

Disclosures:<br />

The following authors have nothing to disclose: Benjamin Samstein, Abigail<br />

Smith, Talia B. Baker, Chris Freise, Jean C. Emond, Brenda W. Gillespie, David<br />

Grant, Elizabeth A. Pomfret, Adrian Cotterell, Trevor L. Nydam, Kim M. Olthoff,<br />

Abhinav Humar, Robert M. Merion<br />

839<br />

Obesity is predictive of NAFLD recurrence in the transplanted<br />

liver graft.<br />

Angela Dolganiuc, Vikas Khullar, Virginia C. Clark, Roberto J.<br />

Firpi; Medicine/GI, University of Florida, Gainesville, FL<br />

Non-Alcoholic Fatty Liver Disease (NAFLD) is an emerging<br />

forerunning cause of liver transplant (OLT) in USA and worldwide.<br />

We have previously reported that cryptogenic cirrhosis<br />

category masks three quarters of NASH patients getting OLT<br />

prior to 2013. We aimed to identify clinical and routine biochemical<br />

features predictive of NAFLD recurrence post-OLT.<br />

Our cohort includes patients from the University of Florida<br />

Liver Transplant database from 1990-2013. 174 out of 1,646<br />

patients were listed for cryptogenic cirrhosis (CC, 156 patients)<br />

or NASH cirrhosis (18 patient, pre-OLT known) as cause of<br />

OLT. In the CC group, 116 patients had electronic records;<br />

40 were excluded from analysis based on explant histology<br />

non-consistent with CC or NAFLD. The remaining 76 patients<br />

were further distributed, based on explant liver histology, into<br />

True CC (24 patients which did not have inflammation or steatosis,<br />

used as control), and NASH (consisting of a subgroup<br />

of n=24 active NASH with steatosis and inflammation, and<br />

n=28 late-stage NASH with inflammation but no steatosis) categories.<br />

There were no differences between the NASH cohort,<br />

CC group-derived NASH and true-CC cohort in terms of age,<br />

gender, or ethnicity or the frequency and characteristics of<br />

metabolic syndrome. NAFLD (all histological stages combined)<br />

recurrence was observed in 42% of the per-OLT known NAFLD<br />

cohort (follow-up interval 3.2±2.8 years), 32% in CC group-derived<br />

NASH with steatosis (follow-up interval 3.6±2.4 years)<br />

and 16% of CC group-derived NASH without steatosis (follow-up<br />

interval 1.2±1.1 years). In contrast, none of the true<br />

CC patients developed NAFLD post-OLT within the follow-up<br />

interval 3.2±2.8 years. Analysis of metabolic syndrome (blood<br />

pressure, lipid profile and fasting glucose) profile revealed<br />

that pre-transplant lipid profile (HDL, LDL, Cholesterol, Tg) was<br />

abnormally low in all groups pre-OLT, likely reflective of pre-<br />

OLT malnutrition, and was not predictive of NAFLD recurrence.<br />

Similarly, systolic blood pressure and presence of DM pre-OLT<br />

were not predictive of NAFLD recurrence in the liver graft.<br />

There were significantly more obese patients in the pre-OLT<br />

known NASH cohort (98%) and in CC group-derived NASH<br />

with (86%) or without (83%) steatosis compared to true-CC<br />

(69%) cohort. More importantly, all patients who had NAFLD<br />

recurrence in their liver graft were obese pre-OLT, despite the<br />

840<br />

3D Printed Liver Models for Precise Resection of Complex<br />

Hepatic Tumors<br />

Maggie H. Samaan, Bijan Eghtesad, Cristiano Quintini, Juliana<br />

Kissiedu, Ibrahim A. Hanouneh, Ryan Klatte, Charles M. Miller,<br />

Nizar N. Zein; Gastroenterology and Hepatology, Digestive Disease<br />

Institute, C.C.F, Cleveland, OH<br />

We recently described a novel method to obtain a patient specific<br />

3D printed liver model s and demonstrated applicability<br />

in living donor liver transplantation [1]. Hepatic resection for<br />

the treatment of liver tumors is an accepted therapy although<br />

associated with a number of limitations due to incomplete characterization<br />

of anatomy and proximity of tumor to vascular and<br />

biliary structures. 3D printed liver models offer better representation<br />

of spatial relationships and improve surgical outcome.<br />

The purpose of this study is to compare 2D images (CT, MRI) to<br />

3D printed liver models for preoperative surgical planning and<br />

intraoperative guidance in complex hepatic tumor resection.<br />

Methods: Prospective study of 6 consecutive patients with complex<br />

liver tumors (located centrally, infiltrating/abutting main<br />

HVs or requiring non-anatomical resections/intrahepatic vasculature<br />

reconstruction). Median lesion size was 7.1cm (range<br />

4.7–11cm). Preoperative evaluation included standard CT &<br />

MRI. A patient-specific 3D printed liver was developed for each<br />

case. For assessment of preoperative planning, a surgical plan<br />

was developed for each case using standard imaging (Plan<br />

A) and was revised after the surgical team was provided with<br />

the 3D printed liver specific to the case (plan B). 3D models<br />

were available during the operation as a reference to guide<br />

resection as deemed necessary by the surgical team in addition<br />

to standard intraoperative ultrasound. 3D printing process<br />

included 3D image reconstruction, digital preparation, 3D<br />

printing and post-printing processing. RESULTS: In 3 of the 6<br />

cases, the pre-operative plan was modified after review of anatomical<br />

spatial relationship of tumor to nearby structures in the<br />

3D model compared to initial plan based on standard imaging<br />

alone. Changes included resection modification, extension and<br />

intrahepatic vascular reconstruction. Intraopertively, surgeons<br />

reported greater confidence with use of 3D model for identification<br />

of intra and extrahepatic structure, segmentation and<br />

tumor specific extent. Surgeons agreed that 3D model offered<br />

a realistic representation that allowed interactive manipulation<br />

simulating intraoperative mobilization. 3D printed liver models<br />

offer a realistic patient-specific representation that may improve<br />

surgical planning and intraoperative process in patients with<br />

complex hepatic tumors. [1] Zein NN, et al. Liver Transpl<br />

2013;19:1304-10. Picture 1: to be submitted upon request<br />

(couldn’t be submitted for it exceeded maximum allowed characters,<br />

needs lower resolution)<br />

Disclosures:<br />

The following authors have nothing to disclose: Maggie H. Samaan, Bijan Eghtesad,<br />

Cristiano Quintini, Juliana Kissiedu, Ibrahim A. Hanouneh, Ryan Klatte,<br />

Charles M. Miller, Nizar N. Zein


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 627A<br />

841<br />

Physical function predicts length of stay (LOS) after liver<br />

transplant (LT): From the Functional Assessment in Liver<br />

Transplantation (FrAILT) Study<br />

Jennifer C. Lai, Kenneth E. Covinsky, Iryna Lobach, Sandy Feng;<br />

UCSF, San Francisco, CA<br />

Background: Cirrhosis causes sarcopenia and undernutrition<br />

resulting in progressive decline in physical function(fxn). The<br />

impact of physical fxn on outcomes after LT is unknown. Methods:<br />

Adult LT recipients with ≥90 days(d) post-LT follow up<br />

underwent outpatient pre-LT tests of physical fxn: gait speed+balance+chair<br />

stands=composite Short Physical Performance<br />

Battery (SPPB; range 12=high fxn to 0=low fxn). Cut-offs of<br />

SPPB≤9(low fxn) and lab LT-MELD≥30 were selected for clinical<br />

relevance. The 1° outcome was LT length of stay (LOS);<br />

the 2° outcome was home discharge (vs. acute care or skilled<br />

nursing facility). Adjusted negative binomial regression evaluated<br />

LOS with physical fxn; Weibull time-to-event analyses<br />

associated physical fxn with time to discharge home. Results:<br />

Included were 171 LT recipients: median age 60y. Median<br />

[interquartile range(IQR)] lab MELD at testing was 16(13-21)<br />

and time from physical fxn testing to LT 1.9mos (1.0-3.8). Low<br />

physical fxn (SPPB≤9) was observed in 36%. Median(IQR) lab<br />

MELD at LT was 22 (17-31); 30% had lab LT-MELD≥30. Median(IQR)<br />

LOS for SPPB≤9 and >9 were 9(7-20) and 8(6-14)<br />

[p9 (81vs.89%; p=0.18). In time-to-event analyses adjusted for<br />

MELD and age, SPPB≤9 was associated with 38% decreased<br />

hazard of home discharge (95%CI=9-58%; p=0.02). Conclusions:<br />

Pre-LT physical fxn is a critical determinant of post-LT LOS<br />

and discharge home. Our data support the implementation<br />

of objective measures of physical fxn into clinical practice to<br />

identify those who may benefit from pre- and post-LT physical<br />

activity interventions.<br />

842<br />

Low serum testosterone predicts outcome in men with<br />

cirrhosis independent of the MELD score<br />

Marie Sinclair 1,2 , Adam Shannon 2 , Mathis Grossmann 2 , Peter W.<br />

Angus 1,2 , Rudolf Hoermann 2 , Paul Gow 1,2 ; 1 Gastroenterology,<br />

Austin Hospital, Heidelberg, VIC, Australia; 2 Medicine, The University<br />

of Melbourne, Melbourne, VIC, Australia<br />

Background: Low serum testosterone has been independently<br />

associated with mortality in a retrospective study of men<br />

with advanced liver disease. This study aims to prospectively<br />

explore the relationship between low testosterone and outcome<br />

in a large cohort of men with cirrhosis Methods: We conducted<br />

a prospective observational study at a single centre. All men<br />

with cirrhosis were invited to undergo baseline sex hormone<br />

profile screening during outpatient consultations. Standard<br />

parameters including MELD score and Child Pugh Score were<br />

also documented at baseline. Patients were then followed for<br />

12 months to assess the composite end-point of mortality or<br />

liver transplantation. Secondary outcomes include mortality,<br />

transplantation, hospitalisation and fracture. Results: 203 men<br />

with cirrhosis were enrolled between April 2013 and March<br />

2014. The median age was 55 years [50, 64], Child Pugh<br />

score was 6 [5, 9] MELD score was 10 [8, 15]. The median<br />

total testosterone level was 17.6nmol/L [8.55, 25.35] (normal<br />

10-27.6nmol/L) and 28.1% of men had low total testosterone.<br />

During the 12 month follow-up period 27 patients died and 12<br />

patients received a liver transplant. Median serum testosterone<br />

was lower in men who died than those who did not (3.20nmol/L<br />

[1.05; 12.6] versus 18.1nmol/L [10.6; 26.0], p


628A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

safety improvements for these patients. Historically, in-depth<br />

reviews of medical records were needed to identify AEs, usually<br />

prompted by a sentinel event. This is very time-consuming,<br />

requiring a high level of skill and resources. The IHI Global<br />

Trigger Tool (GTT) helps identify AEs more efficiently, but lacks<br />

clarity of definitions of complications and guidance for meaningful<br />

interpretation. We developed a modified GTT to improve<br />

precision and is simple to use for review of every LD at LDLT<br />

programs. METHODOLOGY: Medical records were collected<br />

from 4 Adult to Adult Living Donor Liver Transplantation Cohort<br />

Study (A2ALL) centers from 2008-2010. Records included all<br />

documents from admission to discharge and any readmission<br />

within 30 days after LDLT. A multidisciplinary team (transplant<br />

surgeon, physician scientist, nurses, and medical student) modified<br />

the GTT to include transplant specific AEs and created evidence-based<br />

precise clinical definitions of outcomes, compiled<br />

into a codebook. Records were analyzed by two reviewers.<br />

RESULTS: 91 living donor medical records were reviewed. The<br />

mean age was 39. 57% were male, 77% were Caucasian,<br />

19% were Hispanic or black, 93% were English speaking.<br />

722 AEs were identified in 91 records (mean 7.8/patient<br />

(2-19)). 29% had 5-7 AEs and 14% had 12-14 AEs. The ten<br />

most frequent AEs were: nausea (68%), constipation (67%),<br />

hypoxia (64%), anemia (46%), respiratory depression (46%),<br />

tachypnea (46%), hypotension (42%), atelectasis (41%), pleural<br />

effusion (41%), fever (37%). With the enhanced GTT, the<br />

time per for each review was 60 min. Inter-rater reliability was<br />

> 0.9. DISCUSSION: AEs are far more prevalent than reported<br />

to date and can highlight areas requiring improvement before<br />

a sentinel event occurs. The enhanced GTT is a simple and<br />

efficient tool to identify AE in all living liver donors performed<br />

a LDLT center and it can be reliably used by anybody (e.g.<br />

coordinator) after minimal training.<br />

Disclosures:<br />

James V. Guarrera - Grant/Research Support: Organ Recovery Systems<br />

The following authors have nothing to disclose: Blake Platt, Donna Woods, Amna<br />

Daud, Elizabeth A. Pomfret, Mary Ann Simpson, Robert A. Fisher, Anton I. Skaro,<br />

Timothy Curtis, Ella F. Reyes, Erin Wymore, Daniela Ladner<br />

844<br />

Older Liver Transplant (LT) Candidates with Prolonged<br />

Wait-Times Have a Low Probability of LT<br />

Connie W. Wang, Jennifer C. Lai; University of California, San<br />

Francisco, San Francisco, CA<br />

Background: The growing wave of older adults with cirrhosis<br />

has increased demand for LT in ≥65 year(y) olds. While the<br />

decision to list a patient(pt) for LT may initially be straightforward,<br />

this decision is particularly dynamic for pts ≥65y who<br />

are vulnerable to the effects of aging, comorbidities, and even<br />

decreasing motivation for surgery over time. We aimed to evaluate<br />

the probability of LT over time among ≥65y pts. Methods:<br />

All US adult LT candidates listed without MELD exception<br />

points from 2/02-2/12 were categorized by wait-times


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 629A<br />

were more likely to die from technical/operative complications<br />

or multi-organ failure/hemorrhage than “Normal.” Pre-transplant<br />

functional status is a reliable predictor of postoperative<br />

mortality for US liver transplant recipients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Natasha Dolgin, Paulo N. Martins,<br />

Adel Bozorgzadeh<br />

846<br />

Recanalization of complete anastomotic biliary obstruction<br />

after liver transplantation by using a through the<br />

scope magnet: A case series<br />

Erkan Parlak 1 , Fahrettin Küçükay 2 , Aydin Koksal 1 , Ahmet T. Eminler<br />

1 , Mustafa I. Uslan 1 ; 1 Gastroenterology, Sakarya University,<br />

Faculty of Medicine, Sakarya, Turkey; 2 Radiology, Yüksek Ihtisas<br />

Hospital, Ankara, Turkey<br />

Aim: It is necessary to traverse a guidewire through the anastomotic<br />

biliary obstructions after liver transplantation in order to<br />

perform an endoscopic and/or percutaneous treatment. Otherwise<br />

the patient either undergoes a surgical revision or lives<br />

with an external biliary drainage catheter. Herein we evaluated<br />

the efficacy of a novel ‘’through the scope (TTS)’’ magnet<br />

in patients with anastomotic biliary obstructions. Methods:<br />

The magnetic compression anastomotic (MCA) technique: A<br />

‘’Ni coated cylindirical Neodymium-Iron-Boron rare earth magnet’’with<br />

a 5 mm length and 2.4 mm diameter was advanced<br />

over a guidewire to the proximal site of the obstruction via percutaneous<br />

route and to the distal site of the obstruction through<br />

a duodenoscope (Figure). Magnets were removed after recanalization<br />

was achieved. The patients were put on a stent<br />

exchange programme. Results: One hundred and ten patients<br />

(93 LRLT, 17 OLT) with biliary anastomotic stricture underwent<br />

ERCP. TTS MCA was performed in 7 of them (5 LRLT, 2 OLT).<br />

The procedure was unsuccesfull in 2 LRLT patients with dual<br />

anastomosis because of long strictures. The procedure was<br />

successfull in the remaining 5 patients. All of them had a priorly<br />

placed external biliary drainage catheter. Recanalization was<br />

achieved after a mean duration of 9 days (range:5-14). No<br />

magnet related complications were observed. Conclusions: TTS<br />

MCA is effective and safe in the treatment of complete anastomotic<br />

biliary obstructions after liver transplantation. Advantages<br />

and disadvantages with respect to previously defined<br />

larger magnets should be evaluated in future <strong>studies</strong> in order to<br />

determine its indications.<br />

Disclosures:<br />

The following authors have nothing to disclose: Erkan Parlak, Fahrettin Küçükay,<br />

Aydin Koksal, Ahmet T. Eminler, Mustafa I. Uslan<br />

847<br />

Patient Safety in Living Donor Liver Transplantation –<br />

We have a lot to learn<br />

Daniela Ladner 1 , Elizabeth A. Pomfret 2 , Mary Ann Simpson 2 ,<br />

James V. Guarrera 3 , Robert A. Fisher 4 , Anton I. Skaro 1 , Rebeca<br />

Khorzad 1 , Ella F. Reyes 1 , Amna Daud 1 , Erin Wymore 1 , Andy<br />

Hung-Yi Lee 1 , Donna Woods 1 ; 1 CHS, Northwestern University,<br />

Chicago, IL; 2 Lahey Medical Center, Boston, MA; 3 Columbia Medical<br />

Center, New York, NY; 4 Virginia Commonwealth University,<br />

Richmond, VA<br />

Background: Living donor hepatectomies (LDH) are amongst<br />

the most scrutinized surgeries by clinicians and patients alike.<br />

Hence, LDH are probably amongst the most streamlined surgeries<br />

performed today. To examine the streamlining we focused<br />

on equipment, an essential element to LDH. Equipment malfunction<br />

(EM), has been reported in other settings to lead to patient<br />

injury and in fact, 6% of deaths in hospitalized patients were<br />

found to be related to equipment errors. Methods: Direct or<br />

video observation were conducted at four large LDLT centers<br />

(Feb 2012 - March 2014). Web-based Safety (WBS) Debriefings<br />

were sent to every clinician (e.g. nursing, physician, technician)<br />

who participated in an LDH. Results: 110 LDHs were<br />

debriefed and 22% of all debriefings, reported issues related<br />

to equipment. Of those 21% related to incorrect, incomplete or<br />

unavailable instruments, 17% problems related to disposable<br />

accessories, 7% with surgical lights/monitors and 7% with surgical<br />

bed (missing/not working). 34 LDH which were directly<br />

or video observed, demonstrated equipment related issues in<br />

100% of the cases. Similar issues were observed as reported<br />

through WBS, but in addition provider’s frustration, lack of


630A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tidiness, and lack of standardization were observed in 9% of<br />

cases leading to patient injury. Conclusions: EMs happen in<br />

100% of observed LDHs, one of the most scrutinized surgeries<br />

performed. While LDH is extremely complex, the flawless<br />

function of equipment is not. We are far behind other industry<br />

standards in terms of equipment safety (and other safety), and<br />

as a medical community have exhibited unparalleled tolerance<br />

for safety concern even in our most scrutinized and treasured<br />

LDH. To improve patient safety we need to rethink how we<br />

organize and execute surgical procedures and reorganize the<br />

environment to optimally support the execution of the surgical<br />

procedure.<br />

Disclosures:<br />

James V. Guarrera - Grant/Research Support: Organ Recovery Systems<br />

The following authors have nothing to disclose: Daniela Ladner, Elizabeth A.<br />

Pomfret, Mary Ann Simpson, Robert A. Fisher, Anton I. Skaro, Rebeca Khorzad,<br />

Ella F. Reyes, Amna Daud, Erin Wymore, Andy Hung-Yi Lee, Donna Woods<br />

848<br />

The effect of recipient and donor diabetes on long-term<br />

outcomes following liver transplantation, analyzed<br />

according to underlying disease etiology<br />

Oliver J. Duncan, Leon A. Adams, Gerry C. MacQuillan, Gary P.<br />

Jeffrey, George Garas; Hepatology, Sir Charles Gairdner Hospital,<br />

Perth, WA, Australia<br />

Background and aims Pre-transplant recipient diabetes and<br />

donor diabetes have been shown to be independent predictors<br />

of reduced survival following liver transplantation. The aim<br />

of this study is to determine the effect of the combination of<br />

pre-transplant recipient diabetes and donor diabetes on patient<br />

and graft survival, according to the underlying liver disease.<br />

Methods Adults (>18 years) who received a liver transplant<br />

in the United States between 2005 and 2010 were identified<br />

from the United Network for Organ Sharing registry. Exclusion<br />

criteria included living related transplant, ABO incompatible,<br />

multi-organ transplant, donation after cardiac death,<br />

retransplantation and acute liver failure. We examined the<br />

impact of donor diabetes and pre-transplant recipient diabetes<br />

on 30 day, one-year and five-year patient and graft survival<br />

with subanalysis according to the underlying etiology of liver<br />

disease. Results Of the 25280 liver transplant recipients identified,<br />

6053 (23.9%) recipients had a diagnosis of diabetes and<br />

2849 (11.3%) of the grafts originated from a diabetic donor.<br />

A total of 705 (2.8%) recipients were diabetic and received<br />

a liver from a diabetic donor. The commonest indications for<br />

transplantation included hepatitis C (HCV) (31%), hepatocellular<br />

carcinoma (HCC) (22.4%) and alcohol (12.4%). The<br />

median (range) of follow-up was 4.4 (0-9.4) years. The 5-year<br />

overall patient and graft survival was 72% and 70% respectively.<br />

Recipient and donor diabetes independently reduced<br />

5-year survival (69.6% vs. 74% Adjusted Hazard Ratio 1.193,<br />

95% CI 1.126-1.264 p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 631A<br />

Disclosures:<br />

The following authors have nothing to disclose: Daniela Ladner, Kathryn Skibba,<br />

Kathryn Jackson, Niloufar Safaeinili, David S. Goldberg, Lisa B. VanWagner,<br />

Satyender Goel, Abel Kho, Amna Daud, Nazanin Salehitezangi, Anton I. Skaro<br />

850<br />

Objective Measurement of Karnofsky Performance Status<br />

in End Stage Liver Disease Patients<br />

Lisa Louwers, Emmanouil Palaios, Galal El-Gazzaz, Mario Spaggiari,<br />

Bijan Eghtesad, Dympna Kelly; Transplantation and Hepatobiliary<br />

Surgery, Cleveland Clinic, Cleveland, OH<br />

Purpose In OLT the Karnofsky performance score (KPS) is a<br />

heavily weighted component in risk stratification in the Scientific<br />

Registry of Transplant Recipients (SRTR), but there are no<br />

guidelines as to its use in patients with ESLD. Our goal was to<br />

develop an objective designation of the KPS in ESLD patients.<br />

Methods We developed and implemented an algorithm to<br />

objectively assign KPS to OLT patients, used at the time of OLT<br />

offer (see figure). We reviewed KPS in our patients before and<br />

after the tool was implemented. Group 1: 2004 – 2009, 788<br />

patients, 69% male, mean age 53±13 yrs. Group 2: 2011-<br />

2014, 468 patients, 68% male, mean age 53±16 yrs. Group<br />

2 patients had a higher mean chemical MELD score (18±8 vs<br />

20±11, p=0.001), were more often hospitalized at the time<br />

of transplant (17.5% vs 23.1%, p=0.024), had a higher incidence<br />

of COPD and tobacco use (2.6% vs 8.8% and 44.7%<br />

vs 54.5% respectively, p


632A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of the ≥80y livers (range: 22-65 livers per OPO; “high-yield”).<br />

Of these high-yield OPOs, 5 of 6 were high-volume OPOs,<br />

with ≥80y donors accounting for 1.2-2.9% of their total donor<br />

volume. Discard rates of ≥80y procured livers were similar<br />

between the 6 high-yield OPOs and all other OPOs that procured<br />

at least one ≥80y liver (24 vs 26%, p=0.86). Conclusion:<br />

The majority of oldest old (≥80y) donor livers originated from<br />

6 high-yield OPOs. These OPOs increased their procured liver<br />

volume by up to 3% using ≥80y livers, without an increased<br />

rate of ≥80y liver discard. Developing algorithms to strategically<br />

evaluate and guide pursuit of the oldest old donors may<br />

improve efficiency and yield in this rapidly growing population.<br />

Disclosures:<br />

Sandy Feng - Consulting: Novartis<br />

The following authors have nothing to disclose: Suzanne R. Sharpton, Dorry L.<br />

Segev, Jennifer C. Lai<br />

Disclosures:<br />

Jean-Frédéric Blanc - Advisory Committees or Review Panels: Lilly / Imclone,<br />

Merck Serono, Sanofi; Speaking and Teaching: Bayer, BMS, Roche, Abbvie<br />

Masatoshi Kudo - Advisory Committees or Review Panels: Bayer HealthCare;<br />

Grant/Research Support: Bayer HealthCare<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

Morris Sherman - Consulting: Gilead, Merck, Bayer, Arqule, Celsion, Boehringer<br />

Ingelheim, Janssen, Abbvie; Grant/Research Support: WAKO; Speaking and<br />

Teaching: WAKO<br />

The following authors have nothing to disclose: Sasan Roayaie, Ghalib Jibara,<br />

Sarah Berhane, Parissa Tabrizian, Joong-Won Park, Jijin Yang, Lunan Yan, Guohong<br />

Han, Francesco Izzo, Minshan Chen, Phillip Johnson<br />

852<br />

“Oldest Old” (≥80 years) Donors in Liver Transplantation:<br />

National U.S. Organ Procurement Organization<br />

(OPO) Practices<br />

Suzanne R. Sharpton 1 , Sandy Feng 2 , Dorry L. Segev 3 , Jennifer C.<br />

Lai 1 ; 1 Medicine, University of California, San Francisco, San Francisco,<br />

CA; 2 Surgery, University of California, San Francisco, San<br />

Francisco, CA; 3 Surgery, Johns Hopkins, Baltimore, MD<br />

Background: Reports from non-U.S. single centers have shown<br />

that increased utilization of the “oldest old” donor livers [≥80<br />

years (y)] can significantly expand the donor pool. Little is<br />

known about U.S. OPO practice patterns surrounding the<br />

oldest old donors. Screening criteria to identify ≥80y donors<br />

most likely to yield organs are lacking. Methods: We examined<br />

all U.S. adult deceased donors from 2/2005 to 1/2014<br />

(n=63,550). OPOs (n=58) were categorized as low-, mid-,<br />

or high-volume based on tertiles of the total number of donors<br />

pursued for donation. Logistic regression models investigated<br />

donor and OPO factors associated with ≥80y liver discard.<br />

Results: From 403 ≥80y donors, 389 (97%) livers were procured,<br />

of which 291/389 (75%) were transplanted. Donors<br />

≥80y vs 2 drinks per day; 6 vs 16%), median AST/ALT (36/23 vs<br />

45/34), % with liver biopsy (68 vs 40%), and % macrosteatosis<br />

≥30% (5 vs 15%) [p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 633A<br />

854<br />

Outcomes of Patients Undergoing Liver Transplantation<br />

using Donors with Confirmed Positive Blood Cultures: A<br />

review of the UNOS dataset<br />

Moises Huaman 2 , Valery Vilchez 1 , Xiaonan Mei 1 , Malay B.<br />

Shah 1 , Michael F. Daily 1 , Roberto Gedaly 1 ; 1 Surgery, Univ of Kentucky,<br />

Lexington, KY; 2 Division of Infectious Diseases, Department<br />

of Medicin, University of Kentucky, Lexington, KY<br />

Background: The potential effect of blood culture positive donor<br />

(BCPD) on liver transplant outcomes has not been well established.<br />

We aimed to evaluate the impact of BCPD in patient<br />

and graft survival after liver transplantation. Methods: We<br />

retrieved data from the United Network for Organ Sharing<br />

(UNOS) registry on all adults who underwent primary, single<br />

organ deceased-donor liver transplantation in the US between<br />

2008 and 2013. Patients were classified in two cohorts: the<br />

BCPD cohort and the non-BCPD cohort. Graft and patient survival<br />

at 30 days and 1 year were compared between cohorts.<br />

Multivariable analysis using Cox Proportional Hazard models<br />

for graft survival was performed. Results: A total of 28,961<br />

patients were included in the analysis. There were 2,316<br />

(8.0%) recipients of BCPD. BCPD were more likely to received<br />

organs from female diabetic donors with higher body mass<br />

index compared to non-BCPD. Although patient survival was<br />

similar at 7 days, 1-, 6- and 12 months we found a significant<br />

difference in graft survival at this time lines (p=0.009). On multivariable<br />

analysis, BCPD was independently associated with<br />

decreased graft survival (adjusted OR; 1.10, 95% CI 1.005<br />

– 1.203; P


634A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

scores 6-19 or 30-40 did not vary with BMI (47% and 77%<br />

post-LT mortality reduction, respectively, regardless of BMI).<br />

Conclusion: Low BMI is associated with decreased LT survival<br />

benefit in patients with moderate MELD scores. If validated,<br />

these results support the consideration of BMI during LT listing.<br />

significantly increased as LT survival benefit decreased. ROC<br />

analysis showed excellent accuracy of LT futility prediction in<br />

both cohorts. Conclusion: We describe a scoring system that<br />

predicts LT futility based on LT survival benefit decrement. This<br />

has the potential to improve organ allocation.<br />

Association of BMI with LT survival benefit by MELD scores<br />

Disclosures:<br />

The following authors have nothing to disclose: Mohannad Dugum, Nizar N.<br />

Zein, Alex Dugum, Rocio Lopez, Bijan Eghtesad, Ibrahim A. Hanouneh<br />

856<br />

Predicting Futility of Liver Transplantation (LT): A Scoring<br />

System Based on LT Survival Benefit<br />

Mohannad Dugum 1,2 , Nizar N. Zein 3 , Alex Dugum 4 , Rocio<br />

Lopez 5 , Bijan Eghtesad 6 , Ibrahim A. Hanouneh 3 ; 1 Internal Medicine,<br />

Cleveland Clinic, Cleveland, OH; 2 Gastroenterology,<br />

Hepatology and Nutrition, University of Pittsburgh, Pittsburgh, PA;<br />

3 Gastroenterology and Hepatology, Cleveland Clinic, Cleveland,<br />

OH; 4 University of California, San Diego, San Diego, CA; 5 Quantitative<br />

Health Sciences, Cleveland Clinic, Cleveland, OH; 6 Transplantation<br />

Center, Cleveland Clinic, Cleveland, OH<br />

Background and aim: Assessment of LT futility is essential to<br />

improve organ allocation. We aimed to devise a scoring system<br />

to predict LT futility based on factors affecting LT survival benefit.<br />

Methods: We defined LT survival benefit as the difference<br />

between life expectancy if transplanted and life expectancy if<br />

the patient remains on the waiting list (WL). LT was considered<br />

futile if the LT survival benefit was negative. Adult patients listed<br />

for LT in the US 1987-2012 were identified from the UNOS<br />

database. After excluding malignant diagnoses and status 1,<br />

patients were randomly divided to 2 cohorts: 2/3 training<br />

(for model building) and 1/3 validation. Pre-LT survival was<br />

modeled using competing risks analysis and post-LT survival<br />

was assessed using Cox regression. Using these models, life<br />

expectancy on WL and post-LT was estimated for each patient<br />

by calculating the area under survival curve up to 5 years using<br />

the trapezoidal rule. Using the training cohort, multivariable<br />

logistic regression analysis was performed to build a model<br />

for prediction of LT futility. Multiple factors were considered<br />

for inclusion in the model and backward elimination was used<br />

to remove factors with p>0.05. The model was then applied<br />

in the validation cohort, and ROC analysis was performed to<br />

assess the accuracy of LT futility prediction. Results: 33,376<br />

patients were included in the training cohort. Average age<br />

at listing was 53±10 years and 44% underwent LT. Median<br />

LT survival benefit was 6.5 months [0.2, 19.1]. Older age<br />

at listing, higher BMI, lower serum sodium, and higher MELD<br />

predicted survival. Similar results were observed in the LT population.<br />

The proposed score consists of age, BMI, serum sodium<br />

and MELD. When applied in the validation cohort, the score<br />

Disclosures:<br />

The following authors have nothing to disclose: Mohannad Dugum, Nizar N.<br />

Zein, Alex Dugum, Rocio Lopez, Bijan Eghtesad, Ibrahim A. Hanouneh<br />

857<br />

Remnant Liver Ischemia Is Associated With Early Recurrence<br />

and Poorer Survival after Liver Resection<br />

Jai Young Cho, Ho Seong Han, YoungRok Choi, Yoo-Seok Yoon,<br />

Jae Yool Jang, Hanlim Choi, Jae Seong Jang, Seong Uk Kwon;<br />

Seoul National University Bundang Hospital, Seongnam, Korea<br />

(the Republic of)<br />

Background: Non-anatomical hepatectomy or compromised<br />

blood supply to the remnant liver may result in remnant liver<br />

ischemia (RLI) of various extension and severity. There are a<br />

few reports showing association between ischemia-reperfusion<br />

injury and tumor recurrence in animal liver transplantation<br />

model. However, there is no report to evaluate the impact of<br />

RLI after hepatectomy for hepatocellular carcinoma (HCC) on<br />

patient survival. Method: Remnant liver ischemia was graded<br />

on postoperative CT scan in 328 patients who underwent<br />

hepatectomy for HCC between January 2004 and December<br />

2013. We defined RLI as reduced or absent contrast enhancement<br />

during the venous phase and classified to minimal (none<br />

or marginal) or significant (partial, segmental, and necrosis).<br />

Results: We observed radiologic signs of significant RLI in 98<br />

patients (29.9%): 63 partial, 16 segmental, and 19 necrosis,<br />

and these patients showed more complications (P < 0.001)<br />

and longer hospital stays (P = 0.002). Preoperative history of<br />

transarterial embolization (P = 0.040), use of Pringle maneuver<br />

(P = 0.028), and longer operation time (P < 0.001) were independent<br />

risk factors for developing RLI. Patients with significant<br />

RLI showed higher rates of early recurrence within 6 or 12<br />

months after hepatectomy compared those without (P < 0.001).<br />

Moreover, RLI was independent risk factors for both overall<br />

patient (P < 0.001; RR = 6.984; 95% CI, 4.268-11.426) and<br />

disease-free survival (P < 0.001; RR = 5.153; 95% CI, 3.615-<br />

7.345). Conclusion: Partial hepatectomy without RLI and medical<br />

treatment for RLI are highly recommended in patients with<br />

HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jai Young Cho, Ho Seong Han,<br />

YoungRok Choi, Yoo-Seok Yoon, Jae Yool Jang, Hanlim Choi, Jae Seong Jang,<br />

Seong Uk Kwon


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 635A<br />

858<br />

Novel Sonar Guided Technique For Management Of<br />

Incarcerated Umbilical Hernia In Cirrhotic Patients<br />

Awaiting Liver Transplantation<br />

Abd Elrazek A. Hussein 1 , Shymaa E. Bilasy 2 , Hamdy M. Moustafa 1 ,<br />

Mohammed Nafady 3 ; 1 Hepatology and Liver Transplantation, Al<br />

Azhar School of Medicine, Asuit, Egypt; 2 Biochemistry, Faculty of<br />

Pharmacy, Suez Canal University, Ismalia, Egypt; 3 Radiology, Al<br />

Azhar Faculty of Medicine, Asuit, Egypt<br />

Background and Aim: Umbilical hernias pose a management<br />

dilemma in patients with cirrhosis, especially those awaiting<br />

liver transplantation. Complications of umbilical hernias in cirrhotic<br />

patients with ascites include leakage, ulceration, rupture<br />

and incarceration thus leading to a higher mortality rate after<br />

surgical repair. Hernias are usually repaired at the time of<br />

transplantation. However, as the waiting time prior to liver<br />

transplantation is prolonged, the likelihood of an incarcerated<br />

hernia developing is increased. Here, we investigated the feasibility<br />

of using sonar guided repair for incarcerated hernia.<br />

Patients and Methods: This study included 23 cirrhotic patients<br />

examined at the centers of GIT, Hepatology, Radiology,<br />

Faculty of Medicine- Al Azhar University- Egypt. All patients<br />

showed the clinical manifestations of incarceration particularly,<br />

tense feeling of the hernia and manifestation of acute intestinal<br />

obstruction. Both conventional and color doppler abdominal<br />

sonography revealed the extent of obstruction and the vasculature.<br />

Three essential steps were followed to repair the hernia<br />

under ultrasonography guidance. First, we preformed complete<br />

aspiration of the ascetic fluid and the fluid inside the hernia<br />

sac. Then, the distended intestine inside the sac was deflated<br />

to some extent by syringe needle of 16-18 gauge which was<br />

introduced obliquely through the intestinal wall under sonography<br />

guidance. Finally, the hernia was completely reduced<br />

by gentle compression upon the hernia using sterilized jell as<br />

a lubricant and cold fomentation to decrease the size of the<br />

herniated sac. After reduction of the hernia contents, a pressure<br />

pad of dressing was applied at the site of the hernia and<br />

fixed with adhesive tapes to prevent immediate recurrence.<br />

The patient was informed to leave the pressure dressing in situ<br />

and avoid any straining. Results: This study included 20 males<br />

and 3 females between the age of 40-70 yr with incarcerated<br />

hernia between 5 – 9 cm and incarceration duration between<br />

6 – 30 hours. Color doppler sonography revealed inappropriate<br />

intestinal vasculature in three cases and those were referred<br />

back to their surgeons for immediate operation. The remaining<br />

20 patients, showed appropriate intestinal vasculature. Using<br />

the above mentioned sonar guided technique, we succeeded<br />

in repairing hernia in 17 patients (7male and 10 female) with<br />

overall success rate of 73.9%. Conclusion: To our knowledge,<br />

this is the first report to describe sonar guided manipulation<br />

for fixing incarcerated umbilical hernia. Thus, this described<br />

procedure can be either a safe alternative to surgery or at least<br />

as a bridge for later safe elective surgery.<br />

Disclosures:<br />

The following authors have nothing to disclose: Abd Elrazek A. Hussein, Shymaa<br />

E. Bilasy, Hamdy M. Moustafa, Mohammed Nafady<br />

859<br />

Safety of live liver donation by individuals with G6PD<br />

deficiency: Initial results and comparative study<br />

Mettu S. Reddy, Manoj V. Shrivastav, Ilankumaran Kaliamoorthy,<br />

Mohamed Rela; Institute of liver disease and transplantation,<br />

Global health city, Chennai, India<br />

Background G6PD deficiency is the most common genetic<br />

human enzyme defect in the world. Baring a single case report,<br />

there is no published literature regarding the safety of donor<br />

hepatectomy in G6PD deficient individuals. We established a<br />

protocol of selective utilization of G6PD deficient liver donors in<br />

our LDLT program in September 2013. This manuscript presents<br />

the initial results ofour donors with G6PD deficiency (G6PDd).<br />

Methods All potential donors underwent qualitative and quantitative<br />

assessment of G6PDd. Donors with WHO Class III or<br />

Class IV G6PDdwere evaluated for donation if there was no<br />

other suitable donor in the recipient’s family and pre-operative<br />

screening revealed no evidence of previous or ongoing haemolysis.<br />

Care was taken to avoid medications which induce<br />

haemolysis in the peri-operative period. Donors were monitored<br />

for haemolysis. Post-operative course of these 14 patients was compared<br />

with a cohort of 30 non-G6PDd donors case-matched<br />

for age, BMI, type of graft and residual liver volume.><br />

Results Between September 2013 and May 2015, 14 donors<br />

with G6PDd underwent donor hepatectomy. There were 3 right<br />

lobe, 2 left lobe and 9 left lateral segment donors. No donor<br />

received intra-operative transfusion. Two donors with G6PDd<br />

had biochemical evidence of post-operative haemolysis which<br />

was self-limiting. No donor either in the G6PDd or non-G6PDd<br />

group had greater than Clavien 3A complication. Post-op liver<br />

function tests, ICU stay (4 vs 3.5 days, p=0.117), hospital<br />

stay (9 vs 9 days, p=0.517) and % with greater than Clavien<br />

II grading was similar in the two groups (7.1% vs 6.6%,<br />

p=0.187). Donors in the G6PDd group had similar pre-op haemoglobin<br />

(12.4g% vs 12.9g%, p=0.404) but had lower trough<br />

haemoglobin in post-op period (8.1g% vs 92g%, p=0.006),<br />

greater drop in post-op haemoglobin (33.6% vs 25.7%, p=<br />

0.007) and a higher need for post-op blood transfusion (4/14<br />

vs 2/30, p=0.071). Conclusions We present the first reported<br />

series of 14 G6PD deficient individuals who underwent donor<br />

hepatectomy. G6PDd donors can have self-limiting haemolysis<br />

post-operatively and may have a slightly greater need for<br />

post-operative transfusion. Donor hepatectomy can be carried<br />

out safely in carefully selected donors with G6PD deficiency as<br />

long as they are managed in a vigilant clinical setting.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mettu S. Reddy, Manoj V. Shrivastav,<br />

Ilankumaran Kaliamoorthy, Mohamed Rela


636A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

860<br />

Living donor liver transplantation for patients with HCC<br />

exceeding Milan criteria: results of pilot study.<br />

Josep M. Llovet 1,2 , Mihai-Calin Pavel 3 , Jordi Rimola 1 , Maria<br />

Alba Diaz 4 , Jordi Colmenero 3,5 , Constantino Fondevila 3 , Josep<br />

Fuster 1,3 , Pere Gines 3,5 , Jordi Bruix 1,5 , Juan Carlos García-Valdecasas<br />

3 ; 1 Liver Unit, BCLC, IDIBAPS, CIBEREHD, Universitat de Barcelona,<br />

Barcelona, Spain; 2 Institució Catalana de Recerca i Estudis<br />

Avançats, Barcelona, Spain; 3 Liver Transplant Unit, Digestive and<br />

Metabolic Diseases Institute, Hospital Clinic, University of Barcelona,<br />

Barcelona, Spain; 4 Pathology Department, Hospital Clinic,<br />

University of Barcelona, Barcelona, Spain; 5 Liver Unit, Digestive<br />

and Metabolic Diseases Institute, Hospital Clinic, University of Barcelona,<br />

Barcelona, Spain<br />

Introduction: Liver transplantation is the standard of care for<br />

patients with hepatocellular carcinoma (HCC) within Milan criteria<br />

not suitable for resection. Nonetheless, during the last<br />

20 years it became apparent that a subset of patients beyond<br />

Milan criteria might obtain acceptable survival outcomes. In<br />

parallel, living donor liver transplantation has emerged as a<br />

feasible alternative to overcome the paucity of donors. Methods:<br />

In 2002, we proposed for LDLT in Child A-B patients with<br />

HCC a set of criteria that substantially expanded the conventional<br />

indications of transplantation (1 tumor up to 7cm, 5<br />

tumors of < 3cm, 3 tumors < 5cm, or downstaging to Milan<br />

lasting 6 mo after loco-regional therapies) (Bruix& Llovet, Hepatology<br />

2002). Results: We herein present a prospective cohort<br />

of 22 patients with HCC fulfilling these criteria that were treated<br />

with LDLT between 2002 and 2014. Median age was 57 yrs,<br />

20 men, HCV related 14, Child-Pugh A:16, B:6, AFP


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 637A<br />

862<br />

Clinical impact of 18 F-FDG-PET/CT in living donor liver<br />

transplantation for advanced hepatocellular carcinoma<br />

Seung Duk Lee, Eung Chang Lee, Seong Hoon Kim; Liver Cancer<br />

Center, National Cancer Center, Republic of Korea, Goyang-si,<br />

Korea (the Republic of)<br />

Background: The relevant number of patients with hepatocellular<br />

carcinoma (HCC) beyond the Milan criteria has undergone<br />

living donor liver transplantation (LDLT). However, the prognostic<br />

factors for these patients with advanced HCC are not<br />

well established. Methods: From March 2005 to May 2013,<br />

280 patients with HCC underwent LDLT in the National Cancer<br />

Center. Among these, patients beyond the Milan criteria<br />

were retrospectively enrolled. We analyzed the prognostic significance<br />

of 18 F-fluorodeoxyglucose positron emission tomography/computed<br />

tomography ( 18 F-FDG-PET/CT) for selecting<br />

appropriate candidates. Results: Among total 280 patients,<br />

147 patients (52.5%) were confirmed to have HCC beyond<br />

the Milan criteria using pathological reports. The patients who<br />

met and exceeded the Milan criteria had 5-year overall survival<br />

(OS) rates of 87.2% and 64.6%, respectively, (p < 0.001). In<br />

multivariable analysis for OS and DFS in patients beyond the<br />

Milan criteria, PET/CT positivity [hazard ratio (HR) 2.714, p<br />

= 0.013 for OS; HR 3.803, p < 0.001 for DFS], total tumor<br />

size over 10 cm (HR 2.333, p = 0.035 for OS; HR 3.334, p<br />

= 0.001 for DFS), and microvascular invasion (HR 2.917, p =<br />

0.025 for DFS) were significant prognostic factors. In particular,<br />

patients beyond the Milan criteria with a PET/CT-negative<br />

status and total tumor size < 10 cm showed similar OS and<br />

DFS in comparison with those within the Milan criteria. Conclusions:<br />

A PET/CT status in LDLT is a useful marker to predict<br />

survival of patients with advanced HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Seung Duk Lee, Eung Chang Lee,<br />

Seong Hoon Kim<br />

863<br />

Living or Deceased Donor Liver Transplantation for<br />

Hepatocellular Carcinoma: A Western multicenter, intention<br />

to treat, cohort study<br />

Daniel Azoulay; HPB Surgery and liver transplant, Hopital Mondor<br />

APHP, Creteil, France<br />

Purpose. An intention-to-treat analysis of overall survival (ITT-OS)<br />

in cirrhotic patients with hepatocellular carcinoma (HCC) listed<br />

for either living donor (LDLT) [Group LiLDLT] or deceased donor<br />

liver transplantation (DDLT), [Group LiDDLT] across 5 French<br />

liver transplantation (LT) centers. Methods. Records of 861 cirrhotic<br />

patients with HCC who were consecutively listed for<br />

either LDLT (n=79) or DDLT (n=782) from 2000 to 2009 at 5<br />

French centers were analyzed for ITT-OS using Cox model, and<br />

for tumor recurrence using a competitive risk model. Results.<br />

Tumour staging in both groups was similar. Among the listed<br />

patients, 162 dropped-out (20.7%), all from Group LiDDLT (p <<br />

0.0001). At 5-years, ITT-OS was significantly better for Group<br />

LiLDLT vs Group LiDDLT (73.2% vs 66.7%;p = 0.062). Being<br />

listed for LDLT (HR: 0.62 (0.40-0.98); p = 0.039), and MELD<br />

score ≥ 25 at listing (HR: 1.79 (1.11-2.88); p = 0.017) were<br />

independent predictors of ITT-OS. For the 699 transplanted<br />

patients, 5-year OS post-LT (73.2% and 73.0%, p = 0.407)<br />

and HCC recurrence (10.9 % and 11.2 %, p = 0.753) were<br />

similar between those who had a living donor LT [Group LDLT]<br />

and deceased donor LT [Group DDLT], respectively. Tumor<br />

beyond Milan criteria (HR = 2.67 (1.5-4.7); p < 0.001), and<br />

vascular invasion (HR = 2.52 (1.38;4.59); p = 0.003) were<br />

independent predictors of recurrence whereas the type of LT<br />

was not. Conclusion. LDLT improves ITT-OS by obviating dropout,<br />

is not a risk factor of tumor recurrence, and hence should<br />

be equally encouraged in countries where both, LDLT and DDLT<br />

are available.<br />

Disclosures:<br />

The following authors have nothing to disclose: Daniel Azoulay<br />

864<br />

Third Liver Transplantation - Utility and Principles to<br />

Guide Future Transplants; an Analysis of UNOS data<br />

Vandana Khungar 1 , David S. Goldberg 1 , Peter L. Abt 2 ; 1 Medicine,<br />

Division of Gastroenterology, University of Pennsylvania, Philadelphia,<br />

PA; 2 Surgery, University of Pennsylvania, Philadelphia, PA<br />

Background: Repeat liver transplantation (reLT), whether for<br />

recurrent disease, vascular complications, or chronic rejection,<br />

can be a lifesaving therapy. However, donor livers are<br />

a scarce resource, with up to 20% of patients dying on the<br />

waitlist before receiving their first transplant. As more patients<br />

get further out from their 1 st (or 2 nd ) transplant, the transplant<br />

community will need to consider the utility of 3 rd liver transplants.<br />

Yet no data specifically addressing this question exist.<br />

Methods: We performed a retrospective cohort study of all<br />

adult patients listed for their third transplant, who were on<br />

the waiting list between 2/28/02-12/31/2014, using UNOS<br />

data. Kaplan-Meier analyses and log-rank tests were used to<br />

evaluate post-transplant outcomes among the subset of patients<br />

who received their third transplant. Results: 696 adult patients<br />

were on the waitlist for a 3 rd transplant in this period. 362<br />

(52.0%) received a 3 rd transplant, while 218 (31.3%) were<br />

removed from the list for death or clinical deterioration. The<br />

median recipient age was 47 (IQR: 34-45), median donor<br />

risk index (DRI) was 1.33 (IQR: 1.15-1.58), with 211 (58.3%)<br />

donors being under the age of 40. Median laboratory MELD at<br />

transplant was 29 (21-36). The 1,3, and 5 year post-transplant<br />

patient survival rates were 74.5%, 64.4%, and 57.2% respectively.<br />

Graft and patient survival were not statistically different<br />

between centers who had performed 1-9 3 rd transplants vs<br />

≥10 3 rd transplants (p>0.05). Post-transplant patient survival<br />

differed significantly (p=0.001) based on the transplant recipient’s<br />

location at the time of 3 rd transplant. Conclusions: Patients<br />

undergoing reLT were given optimal organs from young donors<br />

with low DRI. Despite this, their post-transplant survival is lower<br />

when compared to patients undergoing 1 st transplantation.<br />

The survival rates among the subset transplanted from the ICU<br />

are dramatically lower, and raise questions about the utility of<br />

transplantation in these high-risk patients outside of the setting<br />

of primary non function (PNF) or hepatic artery thrombosis<br />

(HAT).


638A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Patient survival by location at time of 3rd transplant<br />

Disclosures:<br />

The following authors have nothing to disclose: Vandana Khungar, David S.<br />

Goldberg, Peter L. Abt<br />

865<br />

Mechanisms of Gender Disparities in Liver Transplantation:<br />

An Examination of Organ Offers<br />

Lauren D. Nephew, James D. Lewis, David S. Goldberg, Kimberly<br />

A. Forde; Gastroenterology, Hospital of the University of Pennsylvania,<br />

Philadelphia, PA<br />

Background: In the MELD era, women are more likely than men<br />

to become too sick or die on the liver transplant list and less<br />

likely than men to receive liver transplantation.Gender differences<br />

in creatinine, a component of the MELD score, is a potential<br />

contributing factor. However corrected MELD scores that<br />

account for these differences have not been shown to improve<br />

estimates of 3-month mortality or rates of transplantation in<br />

women. However, receiving a liver transplant is a complex<br />

process that involves first receiving an organ offer. The role of<br />

organ offers in gender disparities in liver transplant have yet to<br />

be explored. We hypothesize that female waitlist candidates<br />

will have lower MELD scores compared to male waitlist candidates<br />

and as a result will receive fewer organ offers. Methods:<br />

A retrospective cohort study of patients listed for liver transplantation<br />

in the United States was conducted. We included adult<br />

match runs from May 10, 2007 through June 17, 2013. Given<br />

different criteria for wait listing and organ allocation for multiorgan<br />

transplantation, retransplantation, bypasses, and status<br />

1, these patients were excluded from the analysis. The primary<br />

outcome was receipt of an organ offer defined as appearing<br />

in the first position on any match run. Multilevel mixed effects<br />

logistic regression modeling was used to account for clustering<br />

of the data by donor service area. Univariable analysis was<br />

performed to evaluate unadjusted gender differences in the<br />

odds of receiving an organ offer. Multivariable analysis then<br />

evaluated region, blood type, race, exception point status,<br />

etiology of liver disease, and listing MELD as potential confounding<br />

variables. Results: There were 64,995 candidates<br />

who appeared on a match run during our study period. Organ<br />

offers were made to 10,714 candidates (65.1% male, 35%<br />

female). On univariable analysis, there was no association<br />

between gender and organ offer (OR=0.98, CI 0.94-1.02).<br />

When blood type, ethnicity/race, exception point status, etiology<br />

of liver disease, and region were added to the model,<br />

their remained no association between gender and organ offer<br />

(OR=.98, CI 0.94-1.03). Conclusions: We hypothesized that<br />

lower creatinine values could result in lower MELD scores and<br />

hence fewer organ offers for female transplant candidates.<br />

However, female waitlist candidates are just as likely as men<br />

to receive an organ offer. These analyses suggest that gender<br />

disparities in liver transplantation do not occur at the level of<br />

the organ offer. Future work will explore gender disparities in<br />

organ refusals.<br />

Disclosures:<br />

James D. Lewis - Grant/Research Support: Bayer<br />

The following authors have nothing to disclose: Lauren D. Nephew, David S.<br />

Goldberg, Kimberly A. Forde<br />

866<br />

Significance of functional hepatic resection rate using<br />

3D fusion image of CT and 99m Tc-galactosyl human<br />

serum albumin scintigraphy<br />

Yosuke Tsuruga, Toshiya Kamiyama, Hirofumi Kamachi, Shingo<br />

Shimada, Kenji Wakayama, Tatsuya Orimo, Hideki Yokoo, Akinobu<br />

Taketomi; Gastroenterological Surgery I, Hokkaido University<br />

Graduate School of Medicine, Sapporo, Japan<br />

Background: Preoperative estimation of hepatic functional<br />

reserve is important for major hepatectomy. Though future<br />

remnant hepatic “volume” is calculated by CT images as volumetry,<br />

it was difficult to estimate the “function” of future liver<br />

remnant. Hepatic receptor imaging with 99m Tc-galactosyl-human<br />

serum albumin ( 99m Tc-GSA) scintigraphy is frequently<br />

used for evaluating hepatic functional reserve. In this study,<br />

we evaluated the usefulness of calculation of functional hepatic<br />

resection rate (FHRR) using 3D fusion image of CT and 99m Tc-<br />

GSA single-photon emission computed tomography (SPECT)<br />

scintigraphy. Methods: 57 patients who underwent more than<br />

one sectionectomy at our institution between October 2013<br />

and March 2015 were enrolled in this study. Of these, 26<br />

patients presented with hepatocellular carcinoma, 12 with hilar<br />

cholangiocarcinoma, 6 with intrahepatic cholangiocarcinoma,<br />

4 with liver metastasis, and 9 with other diseases. 50 patients<br />

underwent bisectionectomy, and 7 trisectionectomy. We compared<br />

the volume hepatic resection rate (VHRR) with FHRR.<br />

FHRR was defined as the resection volume counts per total liver<br />

volume counts of 99m Tc-GSA SPECT from 3D fusion image with<br />

CT. Results: FHRR and VHRR were 38.6 ± 19.9 and 44.5 ±<br />

16.0 (mean ± standard deviation) respectively. The regression<br />

coefficient of FHRR on VHRR was 1.16 and the coefficient of<br />

determination was 0.87 (p 5.0mg/dl) was<br />

observed in 6 of 57 patients. There was no operation-related<br />

death. Conclusion: Calculation of FHRR was important for<br />

major hepatectomy because there was a discrepancy between<br />

FHRR and VHRR in some cases due to hepatic insufficient inflow<br />

and congestion.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yosuke Tsuruga, Toshiya Kamiyama,<br />

Hirofumi Kamachi, Shingo Shimada, Kenji Wakayama, Tatsuya Orimo,<br />

Hideki Yokoo, Akinobu Taketomi<br />

867<br />

Auxiliary partial orthotopic liver transplantation (APOLT)<br />

for metabolic liver disease- is it still relevant?<br />

Mohamed Rela 2 , Priya Ramachandran 1 , Mohamed Safwan 2 ,<br />

Naresh P. Shanmugam 2 , Sanjay Govil 2 , Mettu S. Reddy 2 ; 1 Pediatric<br />

Surgery, CHILDS Trust Hospital, Chennai, India; 2 Institute of<br />

Liver Disease and Transplantation, Global Hospital & Health City,<br />

Chennai, India<br />

Purpose: The role of auxillary partial orthotopic liver transplantation<br />

(APOLT) for non-cirrhotic metabolic liver disease<br />

(MLD) is controversial because of greater technical challenges<br />

and higher rejection rates when compared to orthotopic liver<br />

transplantation (OLT). The aim of our study was to examine<br />

the outcome of APOLT in our series of patients with selected<br />

non-cirrhotic MLD. Methods: We retrospectively reviewed the<br />

case notes of all patients who underwent APOLT for MLD in our


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 639A<br />

centre. Operative details, post-operative course and current<br />

status of these patients were reviewed. Results: Over a 5 year<br />

period (2010-2015), 13 patients underwent APOLT in our centre.<br />

Of these, 8 patients (5 males) underwent APOLT for MLD.<br />

The indications were Propionic acidemia (PA) (n=4), Citrullinemia<br />

(n=1) and Crigler-Najjar syndrome (CNS) (n=3). The<br />

median age at transplantation for PA and citrullinemia was 38<br />

months (9-52 months) while the median age for patients with<br />

CNS was 14 years (6-23 years). Five patients received living<br />

related donor grafts, two (1 Citrullinemia and 1 CNS) received<br />

deceased donor grafts, and one child with CNS received a<br />

partial left lobe graft recovered from the child with PA who<br />

underwent a left lobe APOLT (Domino graft). Graded portal<br />

vein banding (to prevent portal steal to the graft) was done<br />

in all cases. One child had hepatic artery thrombosis on day<br />

4 which was successfully managed by thrombectomy. One<br />

child developed two episodes of acute rejection diagnosed by<br />

biopsy and managed with steroid pulse therapy. All patients<br />

are well (median follow up 9 months) (range 14 days-55<br />

months) with normal liver function tests and unrestricted protein<br />

intake and activity. Conclusion: The outcome of APOLT in our<br />

series is comparable to the reported outcomes for OLT in MLD.<br />

APOLT is an acceptable treatment for non-cirrhotic MLD. The<br />

major advantages are the safety of the remnant native liver in<br />

case of technical complications which can translate into better<br />

patient survival and the lack of an anhepatic phase, which<br />

ensures a smooth intra-operative period. In addition the patient<br />

will have the option of future corrective therapies in the native<br />

liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mohamed Rela, Priya Ramachandran,<br />

Mohamed Safwan, Naresh P. Shanmugam, Sanjay Govil, Mettu S. Reddy<br />

868<br />

Graft Survival following Domino Liver Transplantation:<br />

The U.S. Experience<br />

Ian W. Folkert 1 , David S. Goldberg 2 , Matthew H. Levine 1 , Kim M.<br />

Olthoff 1 , Abraham Shaked 1 , Peter L. Abt 1 ; 1 Surgery, Hospital of<br />

the University of Pennsylvania, Philadelphia, PA; 2 Gastroenterology,<br />

Hospital of the University of Pennsylvania, Philadelphia, PA<br />

Introduction: Domino liver transplantation, in which the<br />

explanted liver from a patient with a hepatic metabolic disorder<br />

(domino donor - DD) is subsequently transplanted into<br />

a second patient (domino recipient - DR), has been utilized<br />

as a method to increase the donor organ supply. Outcomes<br />

of domino transplantation in the U.S. have not yet been fully<br />

evaluated, nor compared to recipients of deceased donor and<br />

living donor grafts. Methods: We performed a retrospective<br />

review of all domino liver transplants performed in the United<br />

States utilizing the UNOS liver transplant database. Graft survival<br />

of DD and DR were compared to recipients of deceased<br />

donor and living donor liver grafts respectively. Results: From<br />

2000-2014, there were 144 domino transplants performed at<br />

38 different centers. Mean DD age was 47; 66% were male.<br />

Mean DD waitlist time was 276 days. The most common diagnosis<br />

leading to transplant in the domino donor was familial<br />

amyloidosis (80%). 26 (18%) DD received multiple organs at<br />

the time of transplant, with 22 receiving a simultaneous heart<br />

transplant. Mean age of the DR was 57 years; 67% were male.<br />

Mean waitlist time for the DR was 543 days. The most common<br />

diagnosis leading to transplant in the domino recipient was<br />

hepatitis C (26%). Mean MELD score at time of transplant for<br />

domino donors vs. deceased donor controls was 8 vs. 21.<br />

Mean MELD score at time of transplant for domino recipients<br />

vs. living donor controls was 15 vs. 15. Kaplan-Meier survival<br />

analysis demonstrated no statistical difference in graft<br />

survival between domino donors and a control group receiving<br />

deceased donor grafts at 1 (88% vs. 85%, p=.34) and 5 years<br />

(71% vs. 69%, p=.56). There was also no difference in graft<br />

survival between domino recipients and living donor controls at<br />

1 (85% vs. 83%, p=.57) and 5 years (68% vs. 71%, p=.66).<br />

Conclusions: Domino transplantation is an effective strategy to<br />

increase liver transplantation, with post-transplant graft survival<br />

that is not different when compared to transplantation using<br />

deceased donor or living donor grafts.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ian W. Folkert, David S. Goldberg,<br />

Matthew H. Levine, Kim M. Olthoff, Abraham Shaked, Peter L. Abt<br />

869<br />

Application of Gadoxetate Disodium-Enhanced MR<br />

Imaging in a Rat Model of Acute Partial Liver Congestion<br />

Akira Shimizu, Akira Kobayashi, Takahide Yokoyama, Hiroaki<br />

Motoyama, Hiroshi Sakai, Noriyuki Kitagawa, Tsuyoshi Notake,<br />

Tomoki Shirota, Kentaro Fukushima, Shinichi Miyagawa; Surgery,<br />

Shinshu University, School of medicine, Matsumoto, Japan<br />

Purpose: To evaluate the features of hepatic congestion on<br />

gadoxetate disodium (Gd-EOB-DTPA)-enhanced magnetic resonance<br />

imaging (MRI) and the mechanisms responsible for<br />

the radiological findings in a rat model of acute partial liver<br />

congestion. Materials and Methods: We used male Wistar rats<br />

weighing 250–320 g. A conventional T1-weighted spin-echo<br />

sequence of the liver was performed using a 1.5T magnetic<br />

resonance imager with an 80-mm magnetic aperture for animal<br />

<strong>studies</strong>. We induced regional liver congestion using partial left<br />

lateral hepatic vein ligation (n = 5) and evaluated the following<br />

in both congestive liver (CL) and non-congestive liver (non-CL):<br />

1) chronological changes in the relative enhancement (RE) up<br />

to 60 minutes after Gd-EOB-DTPA administration and other<br />

parameters including the maximum RE value (C max<br />

), time to<br />

maximum RE (T max<br />

), elimination half-life of RE (T 1/2<br />

), and the<br />

hepatocellular uptake index per unit volume (HUI/V), 2) the<br />

amount of bile secretion and ATP concentration, and 3) mRNA<br />

and protein expression of rat organic anion transporting protein<br />

1a1 (Oatp1a1) using real-time PCR and immunohistological<br />

examination. Results: 1) The RE in the CL reached a small<br />

peak (18%) at 5 minutes, corresponding to approximately half<br />

of the value observed in the non-CL, then slowly decreased in<br />

a linear manner thereafter. The degree of RE in the CL was<br />

significantly lower than that in the non-CL for up to 30 minutes<br />

(P < 0.05). 2) The amount of bile secretion was significantly<br />

decreased in the CL (3.6 ± 4.1 μL/h/g of liver) compared with<br />

those in the non-CL (90.2 ± 32.2 μL/h/g of liver, P < 0.001).<br />

Similarly, the ATP concentration was significantly reduced in<br />

the CL (0.5 ± 0.2 μmol/h/g wet liver) compared with those in<br />

the non-CL (3.3 ± 1.3μmol /g wet liver, P < 0.001). There were<br />

significant correlations between the ATP concentration and the<br />

HUI/V (r 2 = 0.593, 95% CI 0.587-0.599, P = 0.004), C max<br />

(0.546, 0.538-0.555, P = 0.008) and T max<br />

(0.458, 0.487-<br />

0.500, P = 0.012); the closest correlation was observed with<br />

the HUI/V. 3) An immunohistological examination showed that<br />

Oatp1a1 protein expression was downregulated in the CL.<br />

The mRNA level of Oatp1a1 in the CL was significantly upregulated,<br />

compared with that in control rat liver (P = 0.046),<br />

whereas no significant difference was observed between the<br />

CL and the non-CL (P = 0.698). Conclusion: Gd-EOB-DTPA-enhanced<br />

MRI is useful for detection of the congestive liver area<br />

clearly depicted as hypointense. Reduced signal intensity in<br />

the congestive area on Gd-EOB-DTPA-enhanced MRI could


640A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

be explained by the decreased uptake of contrast agent via<br />

Oatp1a1 protein.<br />

Disclosures:<br />

The following authors have nothing to disclose: Akira Shimizu, Akira Kobayashi,<br />

Takahide Yokoyama, Hiroaki Motoyama, Hiroshi Sakai, Noriyuki Kitagawa,<br />

Tsuyoshi Notake, Tomoki Shirota, Kentaro Fukushima, Shinichi Miyagawa<br />

Disclosures:<br />

Thomas D. Schiano - Advisory Committees or Review Panels: salix, merck, gilead,<br />

pfizer; Grant/Research Support: galectin, massbiologics, biotest<br />

The following authors have nothing to disclose: Amy Tan, M. Isabel Fiel, Stephen<br />

C. Ward, Marcelo Facciuto, Swan N. Thung, Myron E. Schwartz, Sander S.<br />

Florman<br />

870<br />

Lack of Radiologic-Pathologic Correlation and Associated<br />

Clinical Features of Patients with Hepatocellular<br />

Carcinoma Undergoing Liver Transplantation<br />

Amy Tan 1 , M. Isabel Fiel 3 , Stephen C. Ward 3 , Marcelo Facciuto 4 ,<br />

Swan N. Thung 3 , Myron E. Schwartz 4 , Sander S. Florman 4 ,<br />

Thomas D. Schiano 2 ; 1 Internal Medicine, Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 2 Liver Diseases, Icahn School<br />

of Medicine at Mount Sinai, New York, NY; 3 Pathology, Icahn<br />

School of Medicine at Mount Sinai, New York, NY; 4 Surgery,<br />

Recanati/Miller Transplantation Institute; Icahn School of Medicine<br />

at Mount Sinai, New York, NY<br />

Background: The Milan criteria are widely applied in selecting<br />

and listing patients with hepatocellular carcinoma (HCC) for<br />

liver transplantation (LT). They fundamentally depend on the<br />

sensitivity of radiologic imaging in detecting lesions. We aimed<br />

to compare radiologic detection with pathologic assessment of<br />

HCC in explanted livers as well as to identify clinical features<br />

that might impact this correlation. Design: At our institution, LT<br />

candidates undergo CT scan or MRI every 3-6 months depending<br />

on if they have HCC and if it is being actively treated.<br />

Per protocol, all explanted livers are weighed, sectioned at<br />

0.5cm intervals, and reviewed by the pathologist. For the current<br />

study, retrospective clinical data were gathered, including<br />

demographics, MELD score, # of lesions seen on imaging, and<br />

history of locoregional therapy (LRT). Retrospective histopathological<br />

review of liver explant sections was performed by a<br />

liver pathologist (MIF) blinded to patient identities. The degree<br />

of cirrhosis based on the Laennec sub-classification (4A=incomplete<br />

cirrhosis, 4B=nodules with thin fibrous septa, 4C=nodules<br />

surrounded by thick fibrous septa) was assessed. Correlation of<br />

the number of tumors found on radiologic imaging and pathologic<br />

examination was performed. Results: 148 patients with<br />

HCC who underwent LT over a 36-month period were studied.<br />

There were 114M & 34F, 60.2±7.2 years of age. 105 had<br />

LT for HCV, 17 for NASH, 19 for HBV, and the rest for other<br />

etiologies. 66 patients (45%, Group A) showed concordance<br />

between imaging and pathologic findings of HCC whereas 82<br />

(55%, Group B) did not. Laennec staging was performed on<br />

patients over the first 21-month period. 32/34 in Group A and<br />

26/27 in Group B had cirrhosis. More Group B than Group<br />

A patients had class 4C cirrhosis (p=0.028). Patients with a<br />

lack of concordance also had a higher creatinine (p=0.019)<br />

and MELD score (p=0.021). There was a trend towards smaller<br />

liver volumes in this group as well. No significant difference<br />

was found in age, sex, time between LT and last imaging<br />

study, and LRT. 39 (48%) in Group B had tumors ≥1cm not<br />

detected by pre-LT imaging. Conclusion: There was a high rate<br />

of non-correlation between radiologic and pathologic findings<br />

in our cohort of patients with known HCC. We found that<br />

HCC is detected less well on imaging in patients having more<br />

advanced fibrosis, as well as in patients with renal dysfunction<br />

and higher MELD scores. Current radiologic modalities may<br />

not be sensitive enough to identify HCC lesions in this population<br />

and thus underestimate tumor burden at time of LT. This<br />

may be due to less optimal penetration of radiologic contrast<br />

into their liver parenchyma.<br />

871<br />

Using Biomarkers to Predict Progression to End-stage<br />

Renal Disease within 6 Months of Liver Transplant<br />

Joseph Bahng, Peter L. Abt, Deirdre Sawinski, Anirban Ganguli,<br />

Debra McCorriston, Mary Ann Lim, Kimberly A. Forde; Perelman<br />

School of Medicine at the University of Pennsylvania, Philadelphia,<br />

PA<br />

Background Liver transplant (LT) recipients are at increased<br />

risk for chronic kidney disease and end-stage renal disease<br />

(ESRD). Creatinine, the standard for assessment of renal function,<br />

provides limited prognostic information about the likelihood<br />

of recovery from acute kidney injury (AKI). Our aim was<br />

to create a predictive model for progression to ESRD within 6<br />

months of LT that incorporates clinical data and biomarkers.<br />

Methods We enrolled 202 patients listed for LT at our center<br />

from 2012-2014. Serum and urine samples were collected at<br />

enrollment, at LT, and 1 year thereafter. Clinical and demographic<br />

data were also collected. Patients were followed until<br />

they developed the primary outcome, ESRD, defined as glomerular<br />

filtration rate


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 641A<br />

872<br />

Traveling for a Liver Transplant in the US does not<br />

Result in a Lower Match MELD for the Majority of<br />

Patients<br />

Catherine T. Frenette 1 , Ann M. Harper 2 , Angeles Baquerizo 1 , Randolph<br />

L. Schaffer 1 , Jonathan S. Fisher 1 , Christopher L. Marsh 1 ;<br />

1 Organ Transplantation, Scripps Clinic, La Jolla, CA; 2 United Network<br />

for Organ Sharing, Richmond, VA<br />

Aim: Given the current geographic disparities in the Unites<br />

States in liver allocation for orthotopic liver transplantation<br />

(OLT), there is a segment of the population that may travel<br />

to a different donor service area other than their home. The<br />

purpose of this study was to determine if patients who traveled<br />

for OLT were more likely to be transplanted at a lower MELD.<br />

Methods: The OPTN database was queried for all patients who<br />

underwent a deceased donor liver transplant from 2010-2014.<br />

Patients with unknown home zip code were excluded (N=340).<br />

Demographic, insurance, transplant, exception status and survival<br />

data were analyzed (N=30,436). Patients who traveled<br />

outside of their home area (Tr) were compared to patients<br />

who underwent OLT at home (NTr). Results: From 2010-2014,<br />

25.6% of patients traveled outside of their home donor service<br />

area (N=7780). Of these patients, 6925 (89%) traveled outside<br />

of their home state. The mean distance traveled was 368<br />

miles for Tr vs 50.3 miles for NTr. There was no difference in<br />

gender between the groups (men 66.9% Tr vs 65.1% NTr).<br />

There was a higher proportion of pediatric patients in the Tr<br />

group (10.2% vs 6.9%, p


642A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

874<br />

Coronary Angiography in Patients being evaluated for<br />

Orthotopic Liver Transplantation<br />

Feng Li, A. James Hanje, Khalid Mumtaz, Robert B. Kirkpatrick,<br />

Douglas M. Levin, Elmahdi Elkhammas, Sylvester Black, Alice Hinton,<br />

Anthony Michaels; The Ohio State University Medical Center,<br />

Columbus, OH<br />

Purpose: The ideal preoperative cardiac screening study in<br />

patients with end-stage liver disease undergoing an orthotopic<br />

liver transplantation (OLT) evaluation has yet to be determined.<br />

Left heart catheterization (LHC) can be considered for cardiac<br />

screening in patients with coronary artery disease (CAD) risk<br />

factors. We aim to identify the patients that would most benefit<br />

from a LHC during an OLT evaluation. Methods: We retrospectively<br />

identified patients who underwent a LHC during an<br />

OLT evaluation at our center between May 2009 and February<br />

2015. Demographic, clinical, laboratory, and procedural<br />

data were collected and analyzed. Patients considered for LHC<br />

included patients with abnormal non-invasive testing, symptoms<br />

consistent with CAD, evidence of structural abnormality<br />

in a technically sufficient study, or associated risk factors such<br />

as tobacco use, diabetes mellitus, family history of premature<br />

CAD, age (> 45 men and >55 women), hyperlipidemia, hypertension,<br />

left ventricular hypertrophy, obesity (BMI > 35), and<br />

non alcoholic fatty liver disease (NASH) or alcoholic liver disease.<br />

Results: 126 patients were included in the study. 25%<br />

had a negative LHC, 64% had a positive LHC without intervention,<br />

and 11% had a positive LHC with non-surgical intervention.<br />

82% of patients with NASH had a positive LHC and<br />

14% required non-surgical intervention; 62% of patients with<br />

alcoholic liver disease had a positive LHC and 8% required<br />

non-surgical intervention; 77% of patients with hepatitis C had<br />

a positive LHC and 17% required non-surgical intervention. Of<br />

the 75% of patients with a positive LHC, 25% underwent OLT<br />

at our medical center, and 21% of the patients were deferred<br />

for transplant listing at our center with the LHC findings contributing<br />

to their deferment. Age (p=0.009) and a history of<br />

hyperlipidemia (p=0.025) were the only two statistically significant<br />

risk factors in predicting a positive coronary angiography<br />

on univariate analysis. A forward stepwise process was used<br />

to build a model containing significant predictors of a positive<br />

coronary angiography. Age was a significant predictor<br />

(p=0.023) and was associated with higher odds of a positive<br />

coronary angiography with Odds Ratios of 1.07 (95% Confidence<br />

Interval (CI), 1.01-1.13), 1.40 (95% CI, 1.05-1.87) and<br />

1.95 (95% CI, 1.10-3.48) for a 1 year, 5 year, and 10 year<br />

age increase, respectively. Conclusion: Coronary Angiography<br />

is an appropriate cardiac screening tool in a select group of<br />

patients undergoing an OLT evaluation particularly in patients<br />

with increasing age and history of hyperlipidemia.<br />

Disclosures:<br />

A. James Hanje - Consulting: Salix pharmaceutical<br />

Anthony Michaels - Advisory Committees or Review Panels: Gilead, BMS, Janssen;<br />

Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Feng Li, Khalid Mumtaz, Robert<br />

B. Kirkpatrick, Douglas M. Levin, Elmahdi Elkhammas, Sylvester Black, Alice<br />

Hinton<br />

875<br />

Long Term Survival Of Liver Grafts From Elderly Donors<br />

Marius Braun 1 , Eviatar Nesher 2 , Michal Cohen-Naftaly 1 , Assaf<br />

Issachar 1 , Amir Shlomai 1 , Yael Harif 1 , Sigal Eisner 2 , Ran Tur-<br />

Kaspa 1 , Eytan Mor 1 ; 1 Liver institute, Rabin Medical Center, Petach<br />

tiqva, Israel; 2 Transplant department, Rabin Medical Center,<br />

Petach tiqva, Israel<br />

Introduction ; There is an increasing gap between the supply<br />

of organ donations and need for liver grafts. The decreasing<br />

quality of the grafts due to aging and overweight population<br />

is expected to lead to a decrease in transplant volume<br />

from deceased donors. The use of extended criteria for grafts<br />

including aged donors may expand the dwindling donor<br />

pool. Traditionally, older donors have been regarded as highrisk<br />

transplants, but recently encouraging results have been<br />

reported. We analyzed our own experience in recipients<br />

of grafts from donors older than 70 years. Methods : Retrospective<br />

analysis of a prospectively maintained transplantation<br />

data base. Recipients of livers older than 70 years were<br />

compared to recipients from donors younger than 60 y and<br />

donors between 61 and 69 y with regard to long term survival<br />

and complications. Statistical analysis included Kaplan Meyer<br />

survival curves, chi squre and regression analysis. Results :<br />

42 patients (25 males) received grafts from donors older than<br />

70 years. Their characteristics and outcomes were compared<br />

with 238 patients who received grafts from donors less than<br />

60 years of age and with 60 patients with grafts from donors<br />

between 61 and 69 years of age. In the group of donors over<br />

70, the mean donor age was 74.3 y (SD 2.8 ) and the recipients<br />

age was 57.8y (range 22-74, SD 9.4).The indications for<br />

transplantation varied: eleven cases of HBV cirrhosis, five HCV<br />

cirrhosis and one hepatocellular carcinoma. Mean MELD at<br />

transplant was 20.8. The long term patient survival at 5 years<br />

was 70% and was similar among all the age groups. Conclusions:<br />

Good quality grafts from elderly donors maybe used<br />

to extend the donor pool. New techniques might improve the<br />

preservation and assessment of these grafts to decrease shortterm<br />

complications and mortality<br />

Disclosures:<br />

The following authors have nothing to disclose: Marius Braun, Eviatar Nesher,<br />

Michal Cohen-Naftaly, Assaf Issachar, Amir Shlomai, Yael Harif, Sigal Eisner,<br />

Ran Tur-Kaspa, Eytan Mor<br />

876<br />

Controlled attenuation parameter (CAP) assessment may<br />

be clinically useful to screen hepatic steatosis for liver<br />

transplant purposes<br />

Youngmi Hong 1 , Ki Tae Yoon 1 , Mong Cho 1 , Daehwan Kang 1 ,<br />

Hyungwook Kim 1 , Cheol Woong Choi 1 , SuBum Park 1 , Jeong<br />

Heo 2 , Hyun Young Woo 2 , Won Lim 2 ; 1 Pusan National University<br />

Yangsan Hospital, Yangsan, Korea (the Republic of); 2 Pusan<br />

National University Hospital, Busan, Korea (the Republic of)<br />

Backgroud & Aims: The presence of hepatic steatosis is associated<br />

with an increased risk of graft loss. The controlled<br />

attenuation parameter (CAP) using transient elastogarphy is a<br />

noninvasive method of assessing hepatic steatosis. However,<br />

the data on the diagnosis performance of CAP is still limited.<br />

Therefore, we assessed the accuracy and the efficacy of CAP<br />

for the detection of hepatic steatosis in potential liver donors.<br />

Methods: We enrolled potential liver donors and all patients<br />

underwent CAP assessment, MRI (Dixon in phase/out of phase<br />

(Dixon IP/OP) with/without fat saturation images) and ultrasonography-guided<br />

liver biopsy. Results: A total of 30 potential<br />

liver donors were included: 24 were male and median<br />

age was 27 years. Clinical and laboratory variables were


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 643A<br />

analyzed according to CAP values. The median CAP value<br />

was 246.4 dB/m (range, 129-371) and CAP value was positively<br />

correlated with alanine aminotransferase (ALT) (r=0.516,<br />

P


644A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

879<br />

Ethanol-Mediated Lipocalin 2 Induction Selectively<br />

Impairs Hepatic Chaperone-Mediated Autophagy and<br />

Causes Alcoholic Steatosis in Mice<br />

Xudong Hu 1,2 , Jiayou Wang 1,3 , Alvin Jogasuria 1 , Kwangwon Lee 1 ,<br />

Jiashin Wu 1 , Min You 1 ; 1 Pharmaceutical Sciences, Northeast Ohio<br />

Medical University (NEOMED), Rootstown, OH; 2 Department of<br />

Biology, Shanghai University of Traditional Chinese Medicine,<br />

Shanghai, China; 3 Department of Anatomy, Guangzhou University<br />

of Chinese Medicine, Guangzhou, China<br />

Lipocalin-2 (LCN2) [also named as SIP24/24 p<br />

3 in mouse and<br />

neutrophil gelatinase-associated lipocalin (NGAL) in human]<br />

has been recently implicated as a critical player in alcohol-induced<br />

fatty liver disease (AFLD). However, the cellular and<br />

molecular mechanisms for its role in mediating detrimental<br />

action of ethanol in the liver are poorly understood. We investigated<br />

the in vivo function of LCN2 in the development of AFLD<br />

by pair-feeding wild-type (WT) and LCN2 knockout (Lcn2KO)<br />

mice using Lieber-DeCarli ethanol-containing diets for 4 weeks.<br />

In WT mice, chronic ethanol administration significantly inhibited<br />

the mRNA expression of hepatic chaperone-mediated autophagy<br />

activity with decreased heat shock protein (Hsp) a8<br />

(Hspa8), Hsp90aa1 and hsp70 compared to pair-fed controls.<br />

Remarkably, mRNA levels of Hspa8, Hsp90aa1 or hsp70 in<br />

the livers of chronically ethanol-fed Lcn2KO mice were completely<br />

restored to the pair-fed control levels. Consistently, the<br />

mRNA levels of lysosome-associated membrane protein type<br />

2A (LAMP-2A) and co-chaperone BAG2, two molecules also<br />

known to participate in chaperone-mediated autophagy, were<br />

significantly inhibited in ethanol-fed WT mice but partially<br />

and significantly restored in the livers of ethanol-fed Lcn2KO<br />

mice. Those results clearly demonstrate that LCN2 deficiency<br />

selectively stabilizes and activates hepatic chaperone-mediated<br />

autophagy in chronically ethanol-fed Lcn2KO mice. The<br />

autophagic flux status was further estimated by quantitating the<br />

protein levels of p62, an autophagy-selective substrate. Compared<br />

to WT control mice, chronic ethanol feeding substantially<br />

inhibited p62 protein levels, and LCN2 deficiency suppressed<br />

p62 to the similar extent. Interestingly, chronic ethanol administration<br />

to Lcn2KO mice augmented inhibition of p62 protein<br />

expression compared to all other groups. Consistent with these<br />

in vivo findings, the mRNAs of Hspa8, Hsp90aa1 and hsp70<br />

were significantly reduced by treatment of recombinant LCN2<br />

in a dose-dependent manner in mouse AML-12 hepatocytes.<br />

Collectively, our <strong>studies</strong> demonstrate that ethanol exposure<br />

selectively inhibits hepatic chaperone-mediated autophagy<br />

and disrupts autophagic flux via LCN2 induction in mice. Ethanol-mediated<br />

disruption of hepatic chaperone-mediated autophagy<br />

and autophagic flux via LCN2 contributes, at least in<br />

part, to the development of alcoholic fatty liver in mice.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xudong Hu, Jiayou Wang, Alvin<br />

Jogasuria, Kwangwon Lee, Jiashin Wu, Min You<br />

Disclosures:<br />

The following authors have nothing to disclose: Kyota Fukazawa, Alexander A.<br />

Vitin, Kenneth Martay, Jorge Reyes<br />

880<br />

Hepatocyte-specific depletion of UBXD8 accelerates<br />

fibrosis in a mouse non-alcoholic steatohepatitis model<br />

Norihiro Imai 1,2 , Yoji Ishizu 1 , Teiji Kuzuya 1 , Takashi Honda 1 ,<br />

Kazuhiko Hayashi 1 , Masatoshi Ishigami 1 , Yoshiki Hirooka 1 , Toyoshi<br />

Fujimoto 2 , Hidemi Goto 1 ; 1 Gastroenterology and Hepatology,<br />

Nagoya University Graduate School of Medicine, Nagoya, Japan;<br />

2 Anatomy and Molecular Cell Biology, Nagoya University Graduate<br />

School of Medicine, Nagoya, Japan<br />

[Background] We previously reported that UBXD8 in lipid<br />

droplets is a key molecule for the degradation of lipidated<br />

ApoB in hepatocytes. More recently, we found that hepatocyte-specific<br />

UBXD8-deficient mice (U8-LKO) fed a high-fat<br />

diet develop periportal macrovesicular steatosis accompanied<br />

by a decrease in VLDL secretion (Imai et al, PLOS ONE 10,<br />

e0127114, 2015). These results demonstrated that UBXD8 is<br />

a crucial factor to regulate VLDL secretion from the liver and<br />

suggested that dysfunction of UBXD8 may aggravate or even<br />

cause NAFLD/NASH. To address this possibility, we studied<br />

U8-LKO mice fed a NASH model diet. [Method] U8-LKO mice<br />

and their age-matched littermates (control) were fed with three<br />

different diets: (1) normal chow diet (CLEA Rodent Diet CE-2),<br />

(2) high-fat diet (CLEA Rodent Diet Quick Fat; 14.4% fat), and<br />

(3) choline-deficient high-fat diet (CDHF, Oriental Yeast; 60%<br />

fat). Blood samples were analyzed biochemically. Paraffin sections<br />

were stained with hematoxylin and eosin or Sirius red to<br />

evaluate morphological changes. [Results] (1) With a normal<br />

diet, liver sections did not exhibit any morphological difference<br />

between U8-LKO and control mice. No sign of hepatic fibrosis<br />

or steatosis was observed. (2) When mice were fed a high-fat<br />

diet for 30 weeks, U8-LKO mice showed periportal steatosis,<br />

as we reported previously. Even when steatosis was prominent,<br />

no sign of fibrosis was observed in control or U8-LKO mice.<br />

(3) When mice were fed a CDHF diet for two weeks, serum<br />

AST and ALT levels were greater in U8-LKO mice than in control<br />

mice (AST, 331 IU/l vs. 227 IU/l; ALT, 334 IU/l vs. 238<br />

IU/l). Serum bilirubin and ALP levels were markedly higher<br />

in U8-LKO mice than in control mice (bilirubin, 96.7 μg/dl<br />

vs. 47.5 μg/dl, p = 0.036; ALP, 840 IU/l vs. 288 IU/L, p =<br />

0.023), although food consumption and body weight changes<br />

were similar in both groups. Under these conditions, the liver<br />

of U8-LKO mice had a lumpy surface, and marked periportal<br />

and perisinusoidal fibrosis was observed histologically. Control<br />

mice showed no fibrosis of the liver. The Sirius red-positive<br />

area was also significantly greater in U8-LKO mice than in<br />

control mice (4.67% vs. 0.75%, p = 0.004). [Conclusion] With<br />

only two weeks of CDHF diet feeding, U8-LKO mice exhibited<br />

remarkable fibrosis; this was not observed with a normal or<br />

simple high-fat diet. These results indicate that UBXD8 functionality<br />

can be largely compensated for in the normal dietary<br />

condition or even with mild fat loading, but that UBXD8 plays<br />

a crucial role when the liver is exposed to intense dietary challenge.<br />

It would be worthwhile to determine if UBXD8 functions<br />

similarly in the human liver.<br />

Disclosures:<br />

Hidemi Goto - Grant/Research Support: MSD, Roche, Bayer, Bristol-Myers, Eisai,<br />

Ajinomoto, Otsuka, Astra, Tanabe, Takeda<br />

The following authors have nothing to disclose: Norihiro Imai, Yoji Ishizu,<br />

Teiji Kuzuya, Takashi Honda, Kazuhiko Hayashi, Masatoshi Ishigami, Yoshiki<br />

Hirooka, Toyoshi Fujimoto


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 645A<br />

881<br />

Regulation of Adaptive Thermogenesis by the Gut<br />

Microbiome: Implications for Non-Alcoholic Fatty Liver<br />

Disease (NAFLD) Pathogenesis<br />

Tibor I. Krisko 1,2 , Hayley T. Nicholls 1 , Katherine B. Leclair 1 , Alexander<br />

S. Banks 1 , David E. Cohen 1 ; 1 Medicine, Brigham and Women’s<br />

Hospital, Boston, MA; 2 Medicine, VA Medical Center, Boston,<br />

MA<br />

Background: The gut microbiome plays a critical role in the<br />

pathophysiology of both obesity and non-alcoholic fatty liver<br />

disease (NAFLD). While an altered energy balance is known<br />

to be associated with liver injury, the mechanisms by which the<br />

gut microbiome controls energy homeostasis and thus may contribute<br />

to NAFLD are incompletely understood. Recent <strong>studies</strong><br />

have shown that increasing thermogenesis ameliorates hepatic<br />

steatosis via both direct and indirect mechanisms (Wang, Nat<br />

Med 2014;20:1436). Aim: This study was designed to elucidate<br />

the contribution of the gut microbiome to adaptive thermogenesis,<br />

as reflected by alterations in energy expenditure<br />

under conditions of varied thermal stress. Methods: We utilized<br />

a temperature-controlled comprehensive lab animal monitoring<br />

system that was custom-designed to maintain sterile conditions.<br />

Rates of O 2<br />

consumption (VO 2<br />

) and CO 2<br />

production (VCO 2<br />

)<br />

were quantified by indirect calorimetry in control and germ-free<br />

C57BL/6J mice as functions of variations in ambient temperatures<br />

that ranged from thermoneutrality (30 °C) to severe cold<br />

stress (4 °C). Energy expenditure (kcal/72 h) was estimated<br />

according to the formula VO 2<br />

(3.82 + 1.22RER) and respiratory<br />

exchange ratio (RER) was calculated as (VCO 2<br />

/VO 2<br />

).<br />

Physical activity was recorded as cumulative beam breaks per<br />

hour over 72 h. Body composition was determined by NMR<br />

spectroscopy and blood glucose concentrations using a glucometer.<br />

Germ-free status was confirmed at the beginning and<br />

end of experiments by quantitative culture and gram staining of<br />

stool. Results: Compared to controls, germ-free mice exhibited<br />

lower energy expenditure, both at thermoneutrality (18.9 ± 0.2<br />

vs 21.1 ± 0.3, p < 0.01) and in the cold (54.7 ± 0.5 vs 58.2<br />

± 0.2, p < 0.05). This occurred despite increased physical<br />

activity in germ-free mice that was most pronounced at 4 °C<br />

(1023 ± 54 vs 1542 ±127, p < 0.05). Germ-free mice exhibited<br />

a greater percentage of adipose tissue mass compared<br />

to controls. Plasma glucose concentrations were reduced and<br />

values of RER were lower in germ-free mice at 4 °C, indicating<br />

increased lipid utilization. Conclusions: Reductions in energy<br />

expenditure and increased adiposity despite increased physical<br />

activity support a key role for the gut microbiome in promoting<br />

adaptive thermogenesis across a broad range of ambient<br />

temperatures. In addition, reduced RER values and plasma<br />

glucose concentrations in germ-free mice indicate that the gut<br />

microbiome promotes the utilization of dietary carbohydrate as<br />

an energy substrate. We speculate that impaired thermogenesis<br />

in the setting of dysbiosis contributes to the pathogenesis<br />

of NAFLD.<br />

Disclosures:<br />

David E. Cohen - Advisory Committees or Review Panels: Merck, Aegerion,<br />

Genzyme; Consulting: Intercept<br />

The following authors have nothing to disclose: Tibor I. Krisko, Hayley T. Nicholls,<br />

Katherine B. Leclair, Alexander S. Banks<br />

882<br />

The dual FXR/TGR5 agonist INT-767 counteracts nonalcoholic<br />

steatohepatitis in a rabbit model of high fat<br />

diet-induced metabolic syndrome<br />

Linda Vignozzi 1 , Ilaria Cellai 1 , Sandra Filippi 4 , Paolo Comeglio 1 ,<br />

Erica Sarchielli 2 , Annamaria Morelli 2 , Elena Maneschi 1 , Gabriella<br />

Barbara Vannelli 2 , Luciano Adorini 3 , Mario Maggi 1 ; 1 Dep.<br />

of Experimental and Clinical Biomedical Sciences, University of<br />

Florence, Florence, Italy; 2 Dep. of Experimental and Clinical Medicine,<br />

University of Florence, Florence, Italy; 3 Intercept Pharmaceuticals,<br />

New York, NY 10011, NY; 4 Interdepartmental Laboratory<br />

of Functional and Cellular Pharmacology of Reproduction, Dep. of<br />

NEUROFARBA and Dep. of Experimental and Clinical Biomedical<br />

Sciences, University of Florence, Florence, Italy<br />

Farnesoid X receptor (FXR) and Takeda G protein-coupled<br />

receptor 5 (TGR5) are interesting pharmacological targets for<br />

the treatment of liver and metabolic diseases. FXR-deficient<br />

mice on a high-fat diet (HFD) exhibit massive hepatic steatosis,<br />

necro-inflammation and fibrogenesis. Moreover, pharmacological<br />

activation of TGR5 in mice promotes protective mechanisms<br />

in biliary epithelial cells, inhibits hepatic and systemic<br />

inflammation. Thus, we hypothesized that a FXR/TGR5 dual<br />

agonist would ameliorate features of nonalcoholic steatohepatitis<br />

(NASH) in a rabbit model of metabolic syndrome (MetS).<br />

Treatment with increasing doses of the dual FXR/TGR5 agonist<br />

INT-767 (3,10,30mg/Kg/day, 5 days a week for 12 weeks)<br />

in a rabbit model of HFD-induced MetS, characterized also<br />

by NASH, dose-dependently reduced several MetS-associated<br />

alterations, including hepatomegaly, insulin resistance,<br />

increase of ALT, glucose and cholesterol levels, while significantly<br />

increasing HDL levels. ALT was positively associated<br />

with all MetS parameters; however introducing all MetS factors<br />

in a multivariate analysis, only total cholesterol levels resulted<br />

positively associated with ALT level (Adj.r: 0.493, p=0.014).<br />

High macrophage M1pro-inflammatory/M2 anti-inflammatory<br />

ratio was observed in MetS-induced NASH, which was<br />

independently associated with serum ALT levels (Adj.r: 0.322,<br />

p=0.032). HFD-induced increase in M1/M2 ratio was reduced<br />

by INT-767 treatment and M2 macrophage markers (IL-10,T-<br />

GFβ) were increased. Genes related to neutrophil apoptosis/<br />

apoptotic-neutrophil clearance (lactoferrin, eNOS, RAGE)<br />

and to extracellular matrix degradation (MMP2, TIMP2) were<br />

also increased by INT-767 treatment. INT-767 also reduced<br />

liver expression of IL-6, which preferentially skews the Th cell<br />

response towards a Th17-phenotype, while increasing Foxp3<br />

expression, a Treg cell marker. Thus these data indicate that<br />

INT-767 can promote the neutrophil and macrophage-driven<br />

resolution phase of inflammation and fibrosis regression. In<br />

addition, INT-767 increased genes related to hepatic fatty acid<br />

metabolism (PPARa, AR, CD36) and lipid droplet formation<br />

(SNAP23, VAMP4,syntaxin5, perilipin) therefore suggesting<br />

that INT-767 counteracts excess fatty acid mediated lipotoxicity<br />

in the liver. Genes related to insulin signaling (IRS1, SREBP1,<br />

G6Pase, and PEPCK) were also increased by INT-767. Finally,<br />

immunohistochemical <strong>studies</strong> demonstrated that INT-767 treatment<br />

significantly reduced both HFD-induced liver inflammation<br />

and fibrosis. In conclusion, INT-767 treatment counteracts<br />

NASH in a rabbit model of HFD-induced MetS by promoting<br />

insulin sensitivity, resolution of inflammation and fibrosis regression.<br />

Disclosures:<br />

Luciano Adorini - Employment: Intercept Pharmaceuticals<br />

Mario Maggi - Consulting: BAYER, ELI LILLY, MENARINI, PROSTRAKAN, INTER-<br />

CEPT<br />

The following authors have nothing to disclose: Linda Vignozzi, Ilaria Cellai,<br />

Sandra Filippi, Paolo Comeglio, Erica Sarchielli, Annamaria Morelli, Elena Maneschi,<br />

Gabriella Barbara Vannelli


646A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

883<br />

Lipid-induced Endoplasmic Reticulum Stress But Not<br />

Hepatic Steatosis Contributes to Blockage of Autophagosome-Lysosome<br />

Fusion<br />

Koichiro Miyagawa, Shinji Oe, Yuichi Honma, Masaru Harada;<br />

University of Occupational and Environmental Health, Kitakyushu,<br />

Japan<br />

Purpose: Autophagy is crucial for intracellular quality control of<br />

proteins and intracellular organelles in various tissues. Several<br />

<strong>studies</strong> showed that blockage of hepatic autophagic degradation<br />

system occurred in obese, and that had an important role<br />

in the development of nonalcoholic fatty liver disease (NAFLD).<br />

However, the mechanism of this blockage has not been fully<br />

elucidated. In this study, we evaluated how lipid overload is<br />

linked to impairment of autophagy in hepatocytes. Methods:<br />

We investigated the effect of lipid overload on hepatic autophagy<br />

in cultured hepatocytes-derived cells treated with monounsaturated<br />

fatty acids (MUFAs) or saturated fatty acids (SFAs).<br />

Autophagic flux was analyzed by immunoblotting and fluorescence<br />

microscopy. We also evaluated the relationship between<br />

hepatic steatosis, endoplasmic reticulum (ER) stress, and autophagy<br />

in this model. Results: SFAs but not MUFAs induced ER<br />

stress and accumulation of autophagosomes despite of lower<br />

accumulation of lipid droplets compared with MUFAs. Microtubule-associated<br />

protein 1 light chain 3 (LC3) turnover assay<br />

demonstrated that SFAs induced the reduction of the rate of<br />

lysosomal degradation of autophagosomes without change<br />

in autophagosome synthesis. Moreover, we monitored autophagic<br />

flux in cells transfected with mRFP-GFP tandem fluorescence-tagged<br />

LC3 (tf-LC3). GFP fluorescence is quenched in<br />

the lysosome but mRFP is not, indicating that GFP and mRFP<br />

are co-localized before the fusion with lysosomes. SFAs but not<br />

MUFAs induced GFP and mRFP overlapping, indicating that<br />

SFAs induced accumulation of autophagosomes. SFAs also<br />

suppressed co-localization of mRFP and lysosome-associated<br />

membrane protein 1 (Lamp1), suggesting that SFAs-induced<br />

impairment of autophagic flux was due to blockage of autophagosome-lysosome<br />

fusion. Furthermore, we assessed lysosomal<br />

acidification. Cells treated with SFA were labeled for<br />

Lysotracker Red and then stained with Lamp1. We could not<br />

identify Lysotracker-negative Lamp1 in SFAs-treated cells as<br />

well as vehicle, suggesting that lysosomal acidification was not<br />

impaired by SFAs treatment. We also revealed that SFAs did<br />

not affect cathepsin activity. These results suggested that SFAs<br />

impaired autophagosome-lysosome fusion without change in<br />

lysosomal function. Moreover, we found that this impairment<br />

occurred in an ER stress-dependent manner. Conclusions: The<br />

disturbance of autophagic flux in NAFLD is due to the blockage<br />

of autophagosome-lysosome fusion, and that involves in the<br />

degree of ER stress but not the intracellular accumulation of<br />

lipid droplets.<br />

Disclosures:<br />

The following authors have nothing to disclose: Koichiro Miyagawa, Shinji Oe,<br />

Yuichi Honma, Masaru Harada<br />

884<br />

CHOP and the Unfolded Protein Response Regulate<br />

NF-κB Activity through IRAK2 in the Pathogenesis of<br />

Non-Alcoholic Steatohepatitis<br />

Jeffrey A. Willy 1 , James L. Stevens 1 , Howard C. Masuoka 2 , Ronald<br />

C. Wek 1 ; 1 Biochemistry and Molecular Biology, Indiana University<br />

School of Medicine, Indianapolis, IN; 2 Department of Medicine,<br />

Indiana University School of Medicine, Indianapolis, IN<br />

Free fatty acid (FFA) induction of cell death is an important feature<br />

of NASH and has been associated with disruption of the<br />

endoplasmic reticulum and activation of the Unfolded Protein<br />

Response (UPR). The mechanisms by which this stress pathway<br />

results in deleterious effects at the organ and cellular levels<br />

are not well understood. In this study, we examined the role of<br />

the UPR in NASH. The UPR features transcriptional and translational<br />

control of gene expression involved in stress remediation,<br />

including the transcription factor CHOP, which can trigger<br />

death processes during chronic endoplasmic reticulum stress.<br />

Primary hepatocytes and HepG2 cells were exposed to physiologic<br />

levels of either saturated or unsaturated FFA. We found<br />

that the saturated FFA palmitate, but not the unsaturated FFA<br />

oleate, induced hepatocyte cell death as a result of inhibition<br />

of autophagic flux in a CHOP-dependent manner. Additionally,<br />

saturated FFA exposure resulted in activation of NF-κB through<br />

expression of IRAK2, resulting in secretion of many cytokines,<br />

including TNFα and IL-8, which contribute to both hepatic cell<br />

death and inflammation. Depletion of CHOP or the RelA subunit<br />

of NF-κB in hepatocytes by shRNA alleviated both autophagy<br />

and cytokine secretion resulting in enhanced cell viability<br />

and lowered inflammatory responses during exposure to saturated<br />

FFA. We next investigated the relevance of these findings<br />

in human liver disease. We showed that the CHOP-IRAK2-<br />

NF-κB pathway is upregulated in liver biopsies from patients<br />

with NASH, in combination with inhibition of autophagic flux,<br />

as measured by accumulation of LC3b and P62. Secretion of<br />

CHOP-dependent cytokines such as TNFα and IL-8 are also<br />

increased in the sera of NASH patients. Interestingly, while<br />

UPR activation was conserved in primary rodent hepatocytes,<br />

there was no induction of IRAK2 and NF-κB in primary rodent<br />

hepatocytes following saturated FFA exposure, providing a<br />

potential explanation for the shortcomings in regards to inflammation<br />

and fibrosis in current rodent models of NASH relative<br />

to the human disease. In conclusion, we have identified a novel<br />

pathway that is upregulated in NASH patients that plays an<br />

active role in both the hepatotoxicity and inflammation at the<br />

level of the hepatocyte following excess saturated FFA. A better<br />

understanding of the mechanism by which the UPR contributes<br />

to hepatocyte toxicity during FFA exposure may provide new<br />

approaches for the diagnosis and treatment of liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jeffrey A. Willy, James L. Stevens,<br />

Howard C. Masuoka, Ronald C. Wek<br />

885<br />

Deletion of G-protein-coupled bile acid receptor (Gpbar-<br />

1, Tgr5) alleviates fasting induced hepatic steatosis by<br />

altering hepatic lipid metabolism<br />

Ajay C. Donepudi, Shannon M. Boehme, John Chiang; Northeast<br />

Ohio Medical University, Rootstown, OH<br />

Introduction: Bile acids and its receptors such as farnesoid X<br />

receptor (Fxr) and G-protein-coupled bile acid receptor (Gpbar-<br />

1, Tgr5) play a major role in nutrient homeostasis. Tgr5 has<br />

been shown to regulate glucose homeostasis by regulating insulin<br />

secretion. Tgr5 is also involved in regulating basal metabolic


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 647A<br />

rates by regulating endocrine hormone levels such as thyroid<br />

hormone levels. Fasting is an adaptive mechanism developed<br />

in mammals for nutrient depravation. Fasting is characterized<br />

with increase in free fatty acid levels in serum and liver causing<br />

hepatic steatosis along with several other physiological<br />

changes. These physiological changes in fatty acid metabolism<br />

during fasting are similar to diabetes condition. In this study<br />

we aim to identify role of Tgr5 in fatty acid metabolism using<br />

fasting model. Methods: Male C57BL/6J (WT) and Tgr5 knock<br />

out (Tgr5-/-) mice in pure C57BL/6J background were either<br />

fed ad libitum or fasted for 24 hrs. Lipid profiles of serum, liver<br />

and adipose tissues were analyzed. Quantitative real-time PCR<br />

was used to assay mRNA expression levels of the genes in bile<br />

acid, cholesterol and lipid metabolism. A Tgr5 specific agonist<br />

INT-777 (Intercept Pharma) was used to activate Tgr5 signaling<br />

in mice. Results: Fasting increased hepatic lipid accumulation<br />

in both WT and Tgr5-/- mice. Lack of Tgr5 attenuated fasting-induced<br />

hepatic triglyceride and fatty acids accumulation compared<br />

to WT mice, without changes in adipose tissue lipolysis.<br />

Gene expression analysis showed Tgr5-/- mice have increased<br />

expression of hepatic lipolysis genes such as Cpt1, Pgc-1α<br />

and autophagy gene LC3II levels. Comprehensive laboratory<br />

animal monitoring system (CLAMS) metabolic <strong>studies</strong> showed<br />

Tgr5-/- mice have increased physical activity and energy<br />

expenditure. Tgr5-/- mice also showed decrease in expression<br />

of hepatic fatty acid transporter Cd36 and decreased hepatic<br />

fatty acid uptake. Additionally, Tgr5-/- mice had increased<br />

hepatic Stat5 activation, which is known to regulate hepatic<br />

lipid metabolism. Moreover, a Tgr5 agonist INT-777 treatment<br />

increased hepatic Cd36 expression and decreased hepatic<br />

Lc3II levels in fasted mice. Conclusion: Although Tgr5 is not<br />

expressed in hepatocytes, Tgr5 may play a role in regulation<br />

of hepatic fatty acid metabolism during fasting. Lack of<br />

Tgr5 attenuates fasting-induced hepatic steatosis by decreasing<br />

hepatic fatty acid uptake and increasing hepatic lipolysis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ajay C. Donepudi, Shannon M.<br />

Boehme, John Chiang<br />

886<br />

Hepatic NOD-like receptors, pyrin domain-containing<br />

3 (NLRP3) inflammasome activation is associated with<br />

histological severity in patients with nonalcoholic fatty<br />

liver disease<br />

Hironori Mitsuyoshi, Kohichiroh Yasui, Tasuku Hara, Hiroyoshi<br />

Taketani, Hiroshi Ishiba, Akira Okajima, Yuya Seko, Atsushi<br />

Umemura, Taichiro Nishikawa, Kanji Yamaguchi, Michihisa Moriguchi,<br />

Yoshio Sumida, Masahito Minami, Yoshito Itoh; Kyoto Prefectural<br />

University of Medicine, Kyoto, Japan<br />

Background and Aims: Nonalcoholic fatty liver disease<br />

(NAFLD) is the hepatic manifestation of metabolic syndrome,<br />

which is associated with dysregulated inflammasome activation<br />

caused by obesity. Inflammasomes are intracellular multiprotein<br />

complexes, comprising nucleotide-oligomerization and binding<br />

domain (NOD)-like receptors, the apoptosis-associated specklike<br />

protein containing a CARD (ASC), and procaspase-1.<br />

Assembly of these components in response to metabolic stress<br />

results in caspase-1 activation, leading to interleukin-1 beta<br />

(IL-1β) and IL-18 activation, which exacerbates inflammation<br />

in systemic organs, including the liver. Here, we examined the<br />

role of inflammasomes in the development of NAFLD. Methods:<br />

Biopsy-proven patients with NAFLD (n = 91) were enrolled.<br />

Liver histology was assessed according to the NAFLD activity<br />

score (NAS). Hepatic (n = 91) and blood (n = 37) levels of the<br />

NOD-like receptor family, pyrin domain-containing 3 (NLRP3),<br />

ASC, procaspase-1, IL-1β, and IL-18 were quantified by realtime<br />

polymerase chain reaction (PCR). Adiponutrin polymorphisms<br />

(rs738409, C>G) were determined by TaqMan PCR.<br />

Serum IL-1β and IL-18 levels were measured by ELISA and<br />

liver tissue caspase-1 expression was evaluated by immunostaining.<br />

Results: Hepatic mRNA levels of NLRP3, ASC, procaspase-1,<br />

IL-1β, and IL-18 were significantly higher in patients<br />

with NAFLD compared with control livers. Blood procaspase-1<br />

mRNA levels were significantly higher in NAFLD patients compared<br />

with healthy controls. In contrast to the blood mRNA<br />

levels, hepatic mRNA levels of NLRP3 inflammasome components<br />

were significantly associated with adiponutrin G alleles.<br />

Hepatic procaspase-1 and IL-1β mRNA levels correlated significantly<br />

with lobular inflammation, hepatocyte ballooning,<br />

and total NAS. Serum IL-18 levels were significantly higher<br />

in NAFLD patients compared with controls, while IL-1β levels<br />

exhibited a non-significant increase. Serum IL-1β and IL-18 levels<br />

were significantly correlated with steatosis, total NAS and<br />

serum transaminase levels. Caspase-1-positive cells were scattered<br />

within the portal tracts, inflammatory foci, and ballooning<br />

hepatocytes. Immunofluorescence staining showed colocalization<br />

of caspase-1 with the macrophage marker CD68. Conclusions:<br />

Our results suggest that NLRP3 inflammasomes are<br />

primed in the liver in NAFLD patients and that this process is<br />

influenced by adiponutrin genotypes. Furthermore, our results<br />

indicate that NLRP3 inflammasomes are activated exclusively<br />

in Kupffer cells or infiltrating macrophages, leading to histological<br />

NAFLD progression through IL-1β and IL-18 production.<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD, Chugai Phamaceutical CO., Bristol-Myers<br />

Squibb, Dainippon Sumitomo, Otsuka, Takeda, Fujifilm medical, Eisai,<br />

Mitsubishi Tanabe, AstraZeneca; Speaking and Teaching: MSD<br />

The following authors have nothing to disclose: Hironori Mitsuyoshi, Kohichiroh<br />

Yasui, Tasuku Hara, Hiroyoshi Taketani, Hiroshi Ishiba, Akira Okajima, Yuya<br />

Seko, Atsushi Umemura, Taichiro Nishikawa, Kanji Yamaguchi, Michihisa Moriguchi,<br />

Yoshio Sumida, Masahito Minami<br />

887<br />

Liver MicroRNA-21 is Overexpressed in Non Alcoholic<br />

Steatohepatitis in Patients and Contributes to this Disease<br />

In Two Mouse Models by restoring PPARα expression<br />

Xavier Loyer 1 , Valerie Paradis 3 , Carole Hénique 1 , Anne-Clémence<br />

Vion 1 , Nathalie Colnot 3 , Coralie L. Guerin 1 , Cécile Devue 1 , Sissi<br />

On 1 , Jérémy Scetbun 1 , Mélissa Romain 1 , Jean-Louis Paul 4 , Marc<br />

E. Rothenberg 5 , Patrick Marcellin 2 , Francois Durand 2 , Pierre<br />

Bedossa 3 , Carina Prip-Buus 6 , Eric Baugé 7 , Bart Staels 7 , Chantal<br />

Boulanger 1 , Alain Tedgui 1 , Pierre-Emmanuel Rautou 1,2 ; 1 INSERM,<br />

U970, Paris Cardiovascular Research Center - PARCC@HEGP,<br />

Paris, France; 2 Hepatology, Hôpital Beaujon, APHP, Clichy,<br />

France; 3 Pathology, Hôpital Beaujon, APHP, Clichy, France; 4 Service<br />

de Biochimie, Hôpital Européen Georges Pompidou, AP-HP,<br />

Paris, France; 5 Cincinnati Children’s Hospital Medical Center, University<br />

of Cincinnati College of Medicine, Division of Allergy and<br />

Immunology, Department of Pediatrics, Cincinnati, OH; 6 INSERM,<br />

U1016, Institut Cochin, Université Paris Descartes, Sorbonne Paris<br />

Cité, Paris, France; 7 European Genomic Institute for Diabetes<br />

(EGID); INSERM UMR1011, University of Lille; and Institut Pasteur<br />

de Lille, Lille, France<br />

Objective: Previous <strong>studies</strong> suggested that microRNA-21<br />

may be up-regulated in the liver in non alcoholic steatohepatitis<br />

(NASH), but its role in the development of this disease<br />

remains unknown. This study aimed to determine the role of<br />

microRNA-21 in NASH. Design: We assessed features of<br />

NASH in two mouse models of NASH where microRNA-21<br />

was either pharmacologically inhibited or genetically deleted:


648A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(a) low-density lipoprotein receptor deficient (Ldlr −/− ) mice fed<br />

a high fat diet and treated with either antagomir-21, antagomir<br />

control (without specific target) or phosphate buffer saline;<br />

(b) microRNA-21 deficient and wild-type mice fed a methionine-choline-deficient<br />

diet (MCD). Wild-type mice fed a chow<br />

diet served as control. We quantified microRNA-21 levels and<br />

assessed cell localization in the liver using real-time quantitative<br />

PCR and in situ hybridization. To determine whether<br />

the effect of inhibiting or deleting microRNA-21 involved<br />

PPARα, a known microRNA-21 target implicated in NASH,<br />

we also fed PPARα deficient mice a MCD diet and treated<br />

them with either antagomir-21 or antagomir control. Finally,<br />

microRNA-21 levels were also quantified in the liver of patients<br />

with NASH or bland steatosis and in normal liver; its cellular<br />

localization was also determined. Results: Inhibiting or deleting<br />

liver microRNA-21 expression in both murine models of NASH<br />

reduced liver cell injury, inflammation, and fibrogenesis without<br />

affecting liver lipid accumulation or betaoxidation. PPARα<br />

was decreased in the liver of mice with NASH and restored following<br />

microRNA-21 inhibition or deletion. In PPARα deficient<br />

mice fed a MCD diet, antagomir-21 had no effect on liver cell<br />

injury, inflammation, and fibrogenesis. Liver microRNA-21 was<br />

overexpressed, primarily in inflammatory and biliary cells, in<br />

both mouse models as well as in patients with NASH, but not<br />

in patients with bland steatosis. Conclusion: This study demonstrates<br />

that microRNA-21 is increased in NASH in liver inflammatory<br />

cells in mice and in patients. MicroRNA-21 inhibition<br />

or deletion strongly decreases liver injury, inflammation and<br />

fibrosis, through PPARα normalization. These findings underscore<br />

the potential interest of antagomir-21 as a therapeutic<br />

strategy for NASH.<br />

Disclosures:<br />

Jérémy Scetbun - Employment: Merck & Co.<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Francois Durand - Advisory Committees or Review Panels: Astellas, Novartis,<br />

BMS; Speaking and Teaching: Gilead<br />

Bart Staels - Advisory Committees or Review Panels: MSD; Consulting: Genfit<br />

Pierre-Emmanuel Rautou - Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Xavier Loyer, Valerie Paradis,<br />

Carole Hénique, Anne-Clémence Vion, Nathalie Colnot, Coralie L. Guerin,<br />

Cécile Devue, Sissi On, Mélissa Romain, Jean-Louis Paul, Marc E. Rothenberg,<br />

Pierre Bedossa, Carina Prip-Buus, Eric Baugé, Chantal Boulanger, Alain Tedgui<br />

888<br />

Two different aspects of PD-1 expression on intrahepatic<br />

CD4+ and CD8+ T cells in a High Fat High Cholesterol<br />

Diet-induced murine NASH<br />

Kanji Yamaguchi 1 , Akira Okajima 1 , Yuya Seko 1 , Hiroshi Ishiba 1 ,<br />

Hiroyoshi Taketani 1 , Atsushi Umemura 1 , Taichiro Nishikawa 1 ,<br />

Yoshio Sumida 1 , Hironori Mitsuyoshi 1 , Kohichiroh Yasui 1 , Takeshi<br />

Okanoue 2 , Yoshito Itoh 1 ; 1 Gastroenterology, Kyoto Prefectural University<br />

of Medicine, Kyoto, Japan; 2 Hepatology Center, Saiseikai<br />

Suita Hospital, Osaka, Japan<br />

Background: Programmed death (PD)-1 signaling plays an<br />

important role of the negative regulation of immune responses<br />

in chronic inflamed liver. Its ligand, PD-L1, is expressed on<br />

hepatocytes in non-alcoholic steatohepatitis (NASH) livers, and<br />

that hepatocytes induce apoptosis in T-cells via their interaction.<br />

The aim of this study is to clarify the pathophysiological<br />

role of PD-1 expressing CD4+ and CD8+ T cells in murine<br />

NASH livers. Methods: Eight-week-old male C57/BL6 mice<br />

were fed either chow or high fat high cholesterol (HFHC) diets<br />

for 24 weeks. Hepatic lymphocytes were isolated by density<br />

gradient centrifuge with 35% isotonic Percoll solution after the<br />

perfusion with 10 ml PBS via portal vein and expression of<br />

PD-1 was analyzed by FACS at the time points of 8, 16, and<br />

24 weeks. Moreover, after sorting the fractions of CD4+/PD-1-,<br />

CD4+/PD-1+, CD8+/PD-1-, CD8+/PD-1+ T cells, mRNA levels<br />

of TNFα, IL-4, IFNγ, IL-6, foxp3, IL-10 in each group of CD4+ T<br />

cells and TNFα, IFNγ, granzyme B, perforin, IL-6 in each group<br />

of CD8+ T cells were assessed by real-time PCR. For further<br />

study of PD-1 expressing T cells, we analyzed the blocking<br />

effect of PD-1 signaling on liver injury by continuously using<br />

PD-1 and PDL-1 neutralizing antibody in HFHC-fed mice for<br />

24 weeks. Results: The number of PD-1 expressing CD4+ or<br />

CD8+ T cells was higher in HFHC diet-fed mouse livers and<br />

increased in a time and the degree of hepatic injury dependent<br />

manner. Furthermore, HFHC diets significantly induced<br />

hepatic PD-L1 mRNA at 24 weeks as previously reported. Realtime<br />

PCR analysis revealed that whereas PD-1 expressions on<br />

CD4+ T cells suppressed TNFα and IL-6 expression but rather<br />

increased IL-10 expression, those on CD8+ T cells significantly<br />

increased granzyme B and perforin despite of decreasing IL-6<br />

expression. Finally, blockade of PD-1 signaling for 24 weeks<br />

greatly ameliorated plasma levels of AST and ALT. Conclusion:<br />

PD-1 expressing T cells were increased in murine NASH livers<br />

with the elevation of hepatic PD-L1 expression. This regulatory<br />

molecule may act as a suppressor in CD4+ T cells but an activator<br />

in CD8+ T cells in this murine NASH model and blockade<br />

of these PD-1 signaling led to the amelioration of liver injury.<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

The following authors have nothing to disclose: Kanji Yamaguchi, Akira Okajima,<br />

Yuya Seko, Hiroshi Ishiba, Hiroyoshi Taketani, Atsushi Umemura, Taichiro Nishikawa,<br />

Yoshio Sumida, Hironori Mitsuyoshi, Kohichiroh Yasui, Takeshi Okanoue<br />

889<br />

DNA Methylation Profiles of Fibrosis-Associated Gene in<br />

Blood Reflects Liver Fibrosis Severity in Human Nonalcoholic<br />

Fatty Liver Disease<br />

Cynthia A. Moylan 1,2 , Susan K. Murphy 3 , Carole Grenier 3 , Manal<br />

F. Abdelmalek 1 , Anna Mae Diehl 1 ; 1 Duke University Medical Center,<br />

Durham, NC; 2 Medicine, Durham Veterans Affairs Medical<br />

Center, Durham, NC; 3 Obstetrics & Gynecology, Duke University,<br />

Durham, NC<br />

Background: Nonalcoholic fatty liver disease (NAFLD) can lead<br />

to cirrhosis, but disease progression is highly variable. Fibrosis<br />

severity is the only independent predictor of cirrhosis risk in<br />

NAFLD. DNA methylation and other epigenetic mechanisms<br />

play a role in instructing fibrosis progression. We reported<br />

that DNA methylation at specific loci of certain liver genes correlated<br />

with their expression in liver and with fibrosis severity<br />

in human NAFLD. Aim: We sought to determine if DNA methylation<br />

profiles of fibrosis-associated genes can be measured in<br />

blood and whether they reflect fibrosis severity (i.e., cirrhosis<br />

risk) in NAFLD patients. Methods: Patients with near-normal histology<br />

and biopsy-proven NAFLD were selected from the Duke<br />

NAFLD Biobank. The cohort was stratified into 3 groups as<br />

shown in the Table: Advanced NAFLD, Mild NAFLD, and Controls.<br />

DNA was isolated from liver and blood, bisulfite modified<br />

using the Zymo EZ DNA Methylation Kit, and then analyzed<br />

by pyrosequencing (PS). Site-specific methylation levels (%met)<br />

for two genes of interest (FGFR2, cg1385327; CASP1,<br />

cg13802966) were correlated by tissue type and compared<br />

by NAFLD severity using Pearson correlation and non-para-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 649A<br />

metric t-tests and ANOVA. Results: Age, gender and BMI were<br />

similar across the 3 groups. More advanced NAFLD patients<br />

had diabetes mellitus and there were more black patients in the<br />

control group. A NAFLD Activity Score greater than or equal to<br />

5 was more common in the advanced NAFLD group. CASP1<br />

and FGFR2 % met could be measured in both liver and blood.<br />

CASP1 %met was similar between tissue types and differed<br />

significantly by group in both liver and blood (controls: 2.8,<br />

mild NAFLD: 3.1, advanced NAFLD: 3.6; p=0.04). Although<br />

FGFR2 %met somewhat correlated in paired normal liver and<br />

blood samples (r=0.55, p=0.01), only FGFR2 %met in liver<br />

differed significantly by group (controls: 9.1, mild NAFLD: 8.6,<br />

advanced NAFLD: 4.8; p


650A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

quent cellular stress may promote redistribution and retention of<br />

LC3 in the nuclei. Further <strong>studies</strong> to explore the mechanism of<br />

nuclear LC3 redistribution are needed.<br />

Disclosures:<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Aaron B. Koenig, Rohini Mehta,<br />

Gary Bratthauer, Fanny Monge<br />

Disclosures:<br />

Brian P. Lam - Advisory Committees or Review Panels: BMS; Speaking and Teaching:<br />

Gilead; Stock Shareholder: Gilead, Vertex<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Elzafir Elsheikh, Mehmet Sayiner,<br />

Zahra Younoszai, Fanny Monge, Lakshmi Alaparthi, Munkhzul Otgonsuren, Hussain<br />

Allawi, Dinan Abdelatif<br />

892<br />

Soluble Intercellular Adhesion Molecule-1 is Independently<br />

Associated with Increased Collagen Deposition<br />

in Patients with Nonalcoholic Steatohepatitis<br />

(NASH)<br />

Elzafir Elsheikh 1,2 , Mehmet Sayiner 2 , Zahra Younoszai 2 , Fanny<br />

Monge 1 , Lakshmi Alaparthi 1 , Munkhzul Otgonsuren 2 , Brian P.<br />

Lam 1 , Hussain Allawi 2 , Dinan Abdelatif 2 , Zachary D. Goodman 1 ,<br />

Zobair M. Younossi 1,2 ; 1 Center for Liver Diseases, Inova Fairfax<br />

Hospital, Falls Church, VA; 2 Betty and Guy Beatty Center for Integrated<br />

Research, Inova Health System, Falls Church, VA<br />

Background: Non-alcoholic fatty liver disease (NAFLD) and its<br />

progressive subtype NASH, are closely associated with chronic<br />

inflammation and fibrosis. Circulating soluble adhesion molecules<br />

are important aspects of this inflammatory response. The<br />

association of serum circulating soluble adhesion molecules<br />

in NASH patients with different degree of hepatic collagen<br />

deposition has not been fully investigated. Aim: To assess the<br />

relationship between hepatic collagen depositions quantified<br />

by morphometry and circulating soluble adhesion molecules<br />

in patients with biopsy-proven NAFLD. Methods: We included<br />

114 subjects [biopsy-proven NAFLD (n=94) and control subjects<br />

without NAFLD (n=20)]. Sections of liver biopsies were<br />

stained with Sirius Red and used for measurement of the percentages<br />

of collagen with morphometry. Concentrations of the<br />

adhesion molecules: sICAM-1 (soluble intercellular adhesion<br />

molecule-1), sVCAM-1 (soluble vascular adhesion molecule-1),<br />

sP-selectin, and sL-selectin (ng/mL) were determined in the<br />

serum collected at the time of liver biopsy using Multiplex platform.<br />

Results: Histologic NASH was seen in 55% of NAFLD.<br />

There were no significant differences between groups based<br />

on gender, ethnicity and BMI. sICAM-1 levels were higher in<br />

NASH group (median, 86) as compared to patients with only<br />

hepatic steatosis (median, 70, p=0.025 ) or controls without<br />

NAFLD (median, 65, p=0.033). We also found that, sVCAM-1<br />

levels were higher in NASH group (median, 409) as compared<br />

to patients with only hepatic steatosis (median, 357, p=0.025).<br />

In fact, higher levels of sICAM-1 positively correlated with the<br />

increased collagen deposition n(r=0.50; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 651A<br />

Giuseppe Penna - Consulting: Intercept Pharmaceuticals; Grant/Research Support:<br />

Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: Edina Hot, Ilaria Spadoni, Maria<br />

Rescigno<br />

Disclosures:<br />

The following authors have nothing to disclose: Wilhelmus J. Kwanten, Wim Martinet,<br />

Benedicte Y. De Winter, Peter P. Michielsen, Sven M. Francque<br />

894<br />

Hepatocyte-specific autophagy-deficiency prevents obesity<br />

in HFD-fed mice and improves glucose tolerance<br />

Wilhelmus J. Kwanten 1 , Wim Martinet 2 , Benedicte Y. De Winter<br />

1 , Peter P. Michielsen 1,3 , Sven M. Francque 1,3 ; 1 Laboratory of<br />

Experimental Medicine and Pediatrics (LEMP) - Gastroenterology<br />

& Hepatology, University of Antwerp, Wilrijk, Belgium; 2 Laboratory<br />

of Physiopharmacology, University of Antwerp, Wilrijk, Belgium;<br />

3 Gastroenterology Hepatology, Antwerp University Hospital<br />

(UZA), Edegem, Belgium<br />

Autophagy is a cellular degradation pathway delivering cytoplasmic<br />

content to lysosomes. There’s increasing evidence that<br />

autophagy interferes with the pathogenesis of NAFLD. Compared<br />

to hepatic lipid metabolism, its role in glucose metabolism<br />

has hardly been studied. AIM To study the influence of<br />

hepatocellular autophagy on HFD and in glucose metabolism.<br />

METHODS Hepatocyte-specific autophagy-deficient C57Bl/6J<br />

mice (Atg7-F) were created using Cre-LoxP with an albumin-promoter<br />

to delete the autophagy gene Atg7. These mice were<br />

compared with their controls (Atg7-WT). Mice were fed control<br />

diet (CD) or high fat diet (HFD) for 16 weeks, with follow-up<br />

of weight/glycaemia. Glucose (ipGTT) and insulin (ipITT) tolerance<br />

tests were performed in week 15. RESULTS Atg7-WT/<br />

HFD mice show a significant increase in body weight compared<br />

to Atg7-WT/CD (36.1 vs 27.2g, p


652A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

atohepatitis (NASH). We have also shown that nitro-oxidative<br />

stress might play a key role in the pathogenesis of this dysfunction<br />

and that activity of the NADPH oxidase is increased in the<br />

liver of mice with NASH. The aim of this study was to evaluate<br />

the role played by the NADPHox in the pathogenesis of the<br />

OXPHOS dysfunction in high fat diet (HFD)-fed mice. Material<br />

and Methods. Twenty four C57BL/6J mice were distributed in<br />

four groups: (1) six wild-type (WT) mice fed a standard chow<br />

diet (SCD) (WT/SCD); (2) six WT mice fed a HFD (WT/HFD);<br />

(3) six NADPHox deficient mice on a SCD (NOX —/— /SCD); (4)<br />

six NADPHox deficient mice on a HFD (NOX —/— /HFD). After<br />

six months on these diets, we studied in the liver: histology,<br />

MRC activity (spectrophotometry), fully assembled MRC complexes<br />

(BN-PAGE), subunits of the MRC complexes (BN/SDS-<br />

PAGE), gene expression of these subunits, as well as of TNFα,<br />

IFNγ, MCP-1, caspase-3, collagen α1(I), and TGFβ1 (RT-PCR),<br />

oxidative (thiobarbituric acid-reactive substances, glutathione)<br />

and nitrosative (immunohistochemistry, nitration of MRC proteins)<br />

stress, and oxidative DNA damage [8-hydroxy-2ʹ-deoxyguanosine<br />

(8-OHdG)]. Results: (1) In HFD mice, we found:<br />

severe steatosis, mild inflammatory infiltrates, ballooning<br />

degeneration, pericellular fibrosis, 3-tyrosine nitrated proteins,<br />

and marked increased gene expression of TNFα, IFNγ,<br />

MCP-1, caspase-3, collagen α1(I), and TGFβ1. Activity of all<br />

complexes of the MRC was decreased to about 50-60% of<br />

activity in WT/SCD. Fully assembled complexes were reduced<br />

to 50% to 60% of that found in control mice. The amount of<br />

all studied subunits was markedly decreased, particularly, the<br />

mitochondrial DNA-encoded subunits. Gene expression of<br />

mitochondrial, but not of genomic, DNA-encoded subunits was<br />

decreased to about 60% of control gene expression. 8-OHdG<br />

was increased in mitochondrial, but not in genomic DNA. (2)<br />

The liver of NOX —/— /HFD mice showed steatosis but not infiltrates,<br />

ballooning degeneration pericellular fibrosis, 3-tyrosine<br />

nitrated proteins, or increased gene expression of inflammatory,<br />

apoptosis or fibrosis related proteins. Likewise, OXPHOS<br />

activity, subunits, their gene expression, and assembly of these<br />

subunits in OXPHOS complexes, and 8OHdG were normal in<br />

the liver of these NOX —/— mice on a HFD. Conclusions. This<br />

study shows that NADPHox plays a key role in the pathogenesis<br />

of the OXPHOS dysfunction found in mice on a HFD, as<br />

this dysfunction was not found in NADPHox deficient mice fed<br />

a HFD.<br />

Disclosures:<br />

The following authors have nothing to disclose: José A. Solís-Herruzo, Pablo<br />

Solis-Munoz, Daniel Fernández-Moreira, Teresa Muñoz-Yagüe, Inmaculada<br />

García-Ruiz<br />

897<br />

Dual targeting of nuclear receptors ameliorates NAFLD<br />

pathogenesis in mice<br />

Pedro M. Rodrigues 2 , Marta B. Afonso 2 , André L. Simão 2 , Marta<br />

Caridade 2 , Catarina C. Carvalho 3 , Alexandre Trindade 3 , António<br />

Duarte 3 , Pedro M. Borralho 1 , Mariana V. Machado 4 , Helena Cortez-Pinto<br />

4,5 , Cecília M. Rodrigues 1 , Rui E. Castro 1 ; 1 iMed.ULisboa<br />

& Dep. of Biochemistry and Human Biology, Faculty of Pharmacy,<br />

University of Lisbon, Lisboa, Portugal; 2 iMed.ULisboa, Faculty of<br />

Pharmacy, University of Lisbon, Lisbon, Portugal; 3 Reproduction<br />

and Development, Interdisciplinary Centre of Research in Animal<br />

Health (CIISA), Faculty of Veterinary Medicine, University of Lisbon,<br />

Lisbon, Portugal; 4 Gastrenterology, Hospital Santa Maria,<br />

Lisbon, Portugal; 5 Faculty of Medicine, University of Lisbon, Lisbon,<br />

Portugal<br />

Non-alcoholic fatty liver disease (NAFLD) burden is rapidly<br />

increasing and disease pathogenesis remains incomplete.<br />

We and others have recently characterized miRNAs as novel<br />

NAFLD pathogenic factors. In addition, nuclear receptors (NRs),<br />

namely peroxisome proliferator-activated receptors (PPARs) and<br />

the farnesoid X receptor (FXR) are currently under scrutiny as<br />

promising therapeutic targets for non-alcoholic steatohepatitis<br />

(NASH). In fact, the FXR agonist obeticholic acid (OCA) significantly<br />

ameliorated NASH in a phase 2 clinical trial. We aimed<br />

to: 1) elucidate the role of miR-21 during NAFLD pathogenesis<br />

in mice, particularly regarding modulation of the miR-21 target<br />

PPARα; and 2) evaluate the therapeutic potential of a dual<br />

NR-targeting approach, specifically activation of PPARα and<br />

FXR through miR-21 ablation and OCA administration, respectively.<br />

C57BL/6 wild type (WT; n=24) and miR-21 knockout<br />

(KO; n=24) mice were fed either a standard diet (SD; n=12)<br />

or a fast food diet (FF; n=12) for 25 weeks. Six animals from<br />

each group received OCA (10 mg/kg/day; Intercept Pharmaceuticals,<br />

Inc.) as a dietary admixture. Human liver biopsies<br />

were obtained from morbidly obese NAFLD patients at different<br />

disease stages (n=28). Human and mice liver samples<br />

were processed for histological analysis and for expression of<br />

miR-21, pro-inflammatory cytokines, NRs and proteins of the<br />

insulin signaling pathway, by qRT-PCR and immunoblotting.<br />

Our results showed that WT FF-fed mice developed hepatomegaly,<br />

with livers presenting macrovesicular steatosis and<br />

inflammatory infiltrates, as well as deregulated insulin signaling.<br />

mRNA levels of TNF-α, IL-6, TLR4, IL-1β and NLRP3 were<br />

increased, further evidencing development of steatohepatitis.<br />

miR-21 levels were increased in WT FF-fed mice concomitantly<br />

with decreased expression of PPARα, a correlation also<br />

found in NAFLD patients, increasing from steatosis to less and<br />

more-severe NASH. OCA-fed animals displayed decreased<br />

CYP7A1 and SREBP-1c liver expression. Further, WT FF+O-<br />

CA-fed mice exhibited decreased steatosis and miR-21 expression,<br />

compared with WT FF-fed mice. In addition, KO FF-fed<br />

mice exhibited significantly reduced liver inflammation, as well<br />

as steatosis severity, in parallel with increased PPARα, comparing<br />

with WT FF-fed mice. Significantly, improvement of the<br />

above histological, biochemical, metabolic and inflammatory<br />

parameters was augmented in KO FF+OCA-fed mice. In conclusion,<br />

our results indicate that miR-21 downregulation, leading<br />

to increased PPARα, together with FXR activation by OCA,<br />

strongly ameliorate NASH in mice, highlighting the therapeutic<br />

potential of novel dual-targeting therapies for human NAFLD.<br />

Disclosures:<br />

Helena Cortez-Pinto - Advisory Committees or Review Panels: Norgine, Lundbeck;<br />

Speaking and Teaching: Janssen, Gilead Janssen<br />

The following authors have nothing to disclose: Pedro M. Rodrigues, Marta B.<br />

Afonso, André L. Simão, Marta Caridade, Catarina C. Carvalho, Alexandre<br />

Trindade, António Duarte, Pedro M. Borralho, Mariana V. Machado, Cecília M.<br />

Rodrigues, Rui E. Castro<br />

898<br />

Alteration of Neutrophil-derived Lipocalin 2 in Bone<br />

Marrow Cells Modulates Non-alcoholic Steatohepatitis<br />

in Mice<br />

Dewei Ye 1,2 , Yu Wang 2,1 , Karen S. L. Lam 1,2 , Aimin Xu 1,2 ; 1 State<br />

Key Laboratory of Pharmaceutical Biotechnology, The University of<br />

Hong Kong, Hong Kong, Hong Kong; 2 Department of Medicine,<br />

The University of Hong Kong, Hong Kong, Hong Kong<br />

Background and Aims: Hepatic infiltration of neutrophils has<br />

been characterized as one of key histological features of nonalcoholic<br />

steatohepatitis (NASH) in rodents and humans. Lipocalin-2<br />

(LCN2) is a secretory glycoprotein originally identified<br />

from neutrophil granules. This study aims to investigate the<br />

potential role of LCN2-chimerism in bone marrow-derived cells<br />

in the pathogenesis of NASH in mice. Methods: LCN2 knockout


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 653A<br />

(KO) mice were backcrossed into ApoE -/- mice, which are susceptible<br />

to diet-induced NASH, to generate ApoE-LCN2 double<br />

knockout (DKO) mice and ApoE -/- /LCN2-wild type (AKO)<br />

control mice. LCN2-chimeric mice were generated through<br />

a combined strategy of lethal irradiation and bone marrow<br />

transplantation. After confirmation of successful reconstitution<br />

of bone marrow cells, chimeric mice (n= 5 - 6 per group) were<br />

fed with high fat high cholesterol (HFHC) diet for 12 weeks to<br />

induce NASH. Results: HFHC diet feeding resulted in a marked<br />

elevation in hepatic expression and circulating levels of LCN2<br />

in AKO mice. Transplantation with LCN2-null bone marrow<br />

cells dramatically abolished HFHC diet-evoked LCN2 elevation<br />

in both liver and blood in AKO mice, whereas reconstitution of<br />

wild-type bone marrow cells in DKO mice restored the hepatic<br />

expression and circulating levels of LCN2. Functionally, transplantation<br />

of LCN2-expressing bone marrow cells from AKO<br />

mice into DKO mice was sufficient to restore the susceptibility<br />

of the recipient mice to HFHC-induced liver inflammation<br />

and damage, as evidenced by deteriorated liver histology<br />

and elevated serum activities of alanine aminotransferase and<br />

aspartate aminotransferase. By contrast, liver lesions triggered<br />

by HFHC diet were dramatically inhibited in chimeric mice<br />

bearing LCN2-deficient bone marrow cells. Likewise, HFHC<br />

diet-elicited hepatic infiltration of inflammatory cells (including<br />

macrophages and neutrophils) and expression of pro-inflammatory<br />

cytokines (TNF-α and IL-1β) and chemokines (MCP-1<br />

and CXCL-2) were significantly augmented in DKO mice by<br />

transplantation of LCN2-expressing bone marrow cells from<br />

AKO mice, but were markedly suppressed by reconstitution of<br />

LCN2-null bone marrow cells from DKO mice into AKO mice<br />

(P 0.05). Conclusions: Bone marrow-derived cells are the major<br />

contributor to elevated LCN2 in both circulation and the liver in<br />

murine model of NASH. LCN2 chimerism in bone marrow-derived<br />

cells is sufficient to exert significant influence on HFHC<br />

diet-induced NASH, suggesting that bone marrow-derived<br />

inflammatory cells and neutrophil-derived LCN2 may serve as<br />

promising therapeutic target for NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Dewei Ye, Yu Wang, Karen S.<br />

L. Lam, Aimin Xu<br />

899<br />

Key role of ASMase in the susceptibility of MAT1A deficient<br />

mice to choline deficient diet-induced steatohepatitis<br />

Cristina Alarcón-Vila 1 , Vicent Ribas 1 , Susana Nuñez 1 , Carmen<br />

Garcia-Ruiz 1,2 , Jose Fernandez-Checa 1,2 ; 1 Instituto Investigaciones<br />

Biomedicas Barcelona, CSIC, Barcelona, Spain; 2 Liver Unit, Hospital<br />

Clinic-IDIBAPS, Barcelona, Spain<br />

Non-alcoholic steatohepatitis (NASH) is an advanced stage of<br />

fatty liver disease characterized by steatosis, oxidative stress,<br />

fibrosis, inflammation and hepatocellular death, which can<br />

progress to cirrhosis. Available treatments are ineffective due<br />

to our poor understanding of the underlined mechanisms. Disturbed<br />

methionine metabolism (e.g. increased homocysteine<br />

and decreased S-adenosyl-L-methionine, SAMe) is known to<br />

contribute to NASH. The first step in methionine metabolism<br />

is its conversion to SAMe by the enzyme methionine adenosyltransferase,<br />

MAT1A, which is expressed mainly in adult<br />

liver. MAT1A deletion markedly depletes hepatic SAMe levels,<br />

induces NASH and increases liver proliferation. Acid sphingomyelinase<br />

(ASMase)-induced ceramide generation regulates<br />

hepatic steatosis and liver fibrogenesis. Our previous observations<br />

in NASH indicated the association between decreased<br />

SAMe and ASMase activation. Thus, we investigated the role<br />

of ASMase in the NASH model of MAT1A null mice fed a<br />

diet deficient in choline (CD). Methods: MAT1A -/- mice were<br />

fed a CD diet for 2 weeks with or without amitriptyline (5mg/<br />

Kg) to inhibit ASMase or after injection with an adenoviral<br />

vector expressing ASMase shRNA (ASMSH) to silence hepatic<br />

ASMase. Moreover, we generated double null mice deficient<br />

in ASMase and MAT1A to examine their susceptibility to CD<br />

feeding for 7 days. Serum and liver samples were processed<br />

to assess ALT levels, ASMase activity, H&E, Oil Red, myeloperoxidase<br />

staining, inflammation and RT-PCR and Western Blot<br />

analyses. Results: MAT1A -/- mice fed CD diet and treated with<br />

amitriptyline or injected with ASMSH exhibited 40-60% lower<br />

mRNA levels of inflammatory markers such as TNFα, IL-1β and<br />

IL-6 and less MPO positive cells. Liver damage monitored by<br />

serum ALT and H&E analyses was reduced in MAT1A KO mice<br />

after amitriptyline treatment or ASMase silencing (ASMSH).<br />

Furthermore, markers of endoplasmic reticulum stress PDI,<br />

CHOP, BiP due to CD diet feeding as well as overexpression of<br />

lipogenic genes SREBP-1c, FAS, HMGCoA, and SREBP2 and<br />

fibrosis, Col1A and TGFb, were markedly reduced (50-60%)<br />

in MAT1A null mice after amitriptyline treatment or ASMase<br />

silencing. These findings correlated with reduced hepatic steatosis<br />

monitored by oil-red staining and lower triglycerides<br />

levels. Similar results regarding steatosis, inflammation, fibrosis<br />

and liver injury were observed in double knockout mice<br />

MAT1A -/- /ASMase -/- fed a CD diet. Conclusions: Modulation<br />

of ASMase activity, through pharmacological o genetic intervention,<br />

protects MAT1A null mice for developing steatohepatitis,<br />

highlighting the role of ASMase in the progression of<br />

steatohepatitis<br />

Disclosures:<br />

The following authors have nothing to disclose: Cristina Alarcón-Vila, Vicent<br />

Ribas, Susana Nuñez, Carmen Garcia-Ruiz, Jose Fernandez-Checa<br />

900<br />

Genetic Ablation of Phosphatidylcholine Transfer Protein<br />

(PC-TP) in ob/ob Mice Improves Glucose Homeostasis<br />

Without Increasing Energy Expenditure: Evidence for the<br />

Direct Role for PC-TP in Hepatic Insulin Resistance<br />

Tibor I. Krisko 1,2 , Katherine B. Leclair 1 , David E. Cohen 1 ; 1 Medicine,<br />

Brigham and Women’s Hospital, Boston, MA; 2 Medicine, VA<br />

Medical Center, Boston, MA<br />

Background: Hepatic insulin resistance is a hallmark of non-alcoholic<br />

fatty liver disease (NAFLD). PC-TP is a specific phosphatidylcholine<br />

binding protein that is highly expressed in liver and<br />

oxidative tissues. Genetic ablation or small molecule inhibition<br />

of PC-TP increases both hepatic insulin sensitivity and energy<br />

expenditure in high fat fed mice. We have demonstrated that<br />

PC-TP suppresses insulin signaling in the liver. However, its<br />

direct contribution to hepatic insulin resistance remains unclear<br />

owing to the capacity of PC-TP to also reduce energy expenditure.<br />

Aim: We utilized leptin-deficient obese (ob/ob) mice,<br />

which exhibit both glucose intolerance and defective thermogenesis,<br />

to dissect the direct contribution of PC-TP to the regulation<br />

of glucose metabolism. Methods: Mice deficient in both<br />

leptin and PC-TP were prepared by crossing Pctp -/- mice on a<br />

C57BL/6J genetic background with ob/ob mice of the same<br />

background. Tolerance tests to glucose (GTT) were performed<br />

by measuring blood glucose concentrations at baseline and<br />

following intraperitoneal administration of glucose. We measured<br />

energy expenditure by indirect calorimetry as functions<br />

of variations in ambient temperatures ranging from thermoneutrality<br />

(30°C) to cold stress (15°C) using a comprehensive<br />

laboratory animal monitoring system, which also recorded<br />

physical activity and food intake. Body composition was deter-


654A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

mined by NMR spectroscopy. Core body temperature was<br />

measured by rectal thermocouple probe. Results: As evidenced<br />

by a decrease in area under the curve (AUC, mg/dl*min),<br />

ob/ob mice lacking PC-TP demonstrated improvement in the<br />

GTT (PC-TP-deficient ob/ob 10,846 ± 1229, ob/ob 17,297 ±<br />

2677, P < 0.05). There were no differences in energy expenditure<br />

when adjusted for lean body mass at any ambient temperature.<br />

There were also no effects of PC-TP expression on<br />

physical activity, food intake or core body temperature. Conclusion:<br />

Marked improvements in GTT in PC-TP-deficient ob/ob<br />

mice in the absence of increases in energy expenditure indicate<br />

a direct role for PC-TP in regulating glucose metabolism.<br />

These findings support the potential utility of PC-TP inhibitors in<br />

the management of NAFLD.<br />

Disclosures:<br />

David E. Cohen - Advisory Committees or Review Panels: Merck, Aegerion,<br />

Genzyme; Consulting: Intercept<br />

The following authors have nothing to disclose: Tibor I. Krisko, Katherine B.<br />

Leclair<br />

901<br />

Aging associated impaired metabolic compensation in<br />

adipose tissues promotes diet-induced murine nonalcoholic<br />

fatty liver disease<br />

Hiroyoshi Taketani 1 , Taichiro Nishikawa 1 , Hisakazu Nakajima<br />

2 , Satoru Sugimoto 2 , Ikuyo Itoh 2 , Kazuki Koudou 2 , Atsushi<br />

Umemura 1 , Kanji Yamaguchi 1 , Michihisa Moriguchi 1 , Yoshio<br />

Sumida 1 , Hironori Mitsuyoshi 1 , Kohichiroh Yasui 1 , Yoshito Itoh 1 ;<br />

1 Gastroenterology and Hepatology of Kyoto Prefectural University<br />

of Medicine, Kyoto-shi, Japan; 2 Pediatrics, Kyoto Prefectural University<br />

of Medicine, Kyoto, Japan<br />

Background: The prevalence of nonalcoholic fatty liver disease<br />

(NAFLD), a hepatic manifestation of metabolic syndrome,<br />

generally increases with age in humans. The progression of<br />

NAFLD is expected to result from a failed metabolic homeostasis.<br />

Recently, adipose tissues lead to renewed interest on<br />

energy metabolism as brown adipose tissues with huge energy<br />

expenditure was demonstrated to be inducible (iBAT) from<br />

progenitors within subcutaneous white adipose tissues (sWAT)<br />

aside from classical BAT (cBAT) in adult human. On the other<br />

hand, age related change in sWAT is reported to cause a loss<br />

of iBAT, though its detailed mechanism is still unclear. Here<br />

we evaluated detailed adipose tissues profiling associated<br />

with iBAT under aging and its association with progression of<br />

NAFLD. Methods: Six-week-old male C57BL/6 mice were fed<br />

with control chow (C) or high-fat diet (60% fat; HF) for 12 or<br />

24 weeks as Short-term (St) or Long-term (Lt) study group, or<br />

switched C to HF diet at 18 weeks age (C/HF) for 24 weeks.<br />

All animals were evaluated on body weight gain, energy<br />

expenditure and blood biochemical assays including lipid and<br />

glucose tolerance test. At necropsy liver and adipose tissues<br />

were examined for histological analysis containing NAFLD<br />

Activity Score (NAS) and beta-gal staining as a hallmark of<br />

cellular senescence. The study of gene expression profiling<br />

on self-renewal of progenitors, iBAT and adipocytokines in<br />

adipose tissues was also evaluated in Quantitative reverse transcriptase<br />

polymerase chain reaction and flow cytometry. Inducing<br />

capacity of BAT from progenitors in adipose tissues was<br />

also examined in vitro culture system. Results: Animals in St/HF<br />

group gained body weight, subcutaneous and abdominal adipose<br />

storages, and showed two types of BAT compensations<br />

(hypertrophy of classical BAT and appearance of subcutaneous<br />

induced BAT) and mild hepatic steatosis. On the other<br />

hand, animals both in Lt/HF and Lt/C/HF groups decreased<br />

energy expenditure and developed more severe lipid and glucose<br />

intolerance, higher NAS with lost BAT compensations<br />

and advanced adipose senescence, where there was no significant<br />

difference between both groups. Adipose progenitor<br />

and induced BA associated factors (FGF4, and UCP1) were<br />

highly expressed in St/HF group with abundant differentiation<br />

capacity of progenitor cells, and these findings disappeared in<br />

aged groups (Lt/HF and Lt/C/HF). Conclusion: Aging caused<br />

impaired metabolic homeostasis and progression of NAFLD<br />

through deteriorated adipose compensation. Aging associated<br />

adipose tissues profiling might provide a new insight to go<br />

toward a fully-integrated understanding of NAFLD<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

The following authors have nothing to disclose: Hiroyoshi Taketani, Taichiro<br />

Nishikawa, Hisakazu Nakajima, Satoru Sugimoto, Ikuyo Itoh, Kazuki Koudou,<br />

Atsushi Umemura, Kanji Yamaguchi, Michihisa Moriguchi, Yoshio Sumida,<br />

Hironori Mitsuyoshi, Kohichiroh Yasui<br />

902<br />

The dual FXR/TGR5 agonist INT-767reduces visceral fat<br />

mass, promoting preadipocyte brown differentiation,<br />

mitochondrial function and insulin sensitivity in a rabbit<br />

model of high fat diet-induced metabolic syndrome<br />

Linda Vignozzi 2 , Ilaria Cellai 2 , Sandra Filippi 1 , Paolo Comeglio 2 ,<br />

Tommaso Mello 2 , Daniele Bani 3 , Daniele Guasti 3 , Erica Sarchielli 3 ,<br />

Annamaria Morelli 3 , Elena Maneschi 2 , Gabriella Barbara Vannelli<br />

3 , Luciano Adorini 4 , Mario Maggi 2 ; 1 Interdepartmental Laboratory<br />

of Functional and Cellular Pharmacology of Reproduction,<br />

Dep. of NEUROFARBA and Dep. of Experimental and Clinical<br />

Biomedical Sciences, University of Florence, Florence, Italy; 2 Dep.<br />

of Experimental and Clinical Biomedical Sciences, University of<br />

Florence, Florence, Italy; 3 Dep. of Experimental and Clinical Medicine,<br />

University of Florence, Florence, Italy; 4 Intercept Pharmaceuticals,<br />

New York, NY<br />

Expanding brown adipose tissue is a potential therapeutic strategy<br />

to counteract insulin resistance and metabolic syndrome<br />

(MetS).Farnesoid X receptor (FXR) and Takeda G protein-coupled<br />

receptor 5 (TGR5)activation enhances insulin sensitivity,<br />

suggesting the capacity of FXR/TGR5 agonists to promote<br />

brown differentiation in adipose tissue. Treatment with increasing<br />

doses of the dual FXR/TGR5 agonist INT-767 (3,10,30mg/<br />

Kg, orally, daily, 5 days a week for 12 weeks) in a rabbit model<br />

of high fat diet (HFD)-induced MetS, characterized by insulin<br />

resistance, hypertension, atherogenic dyslipidemia and visceral<br />

adipose tissue (VAT) accumulation, dose-dependently reduced<br />

VAT mass, glycemia, insulin resistance and cholesterol levels<br />

while significantly increasing HDL levels. INT-767 (30mg/Kg/<br />

day) also significantly reduced HFD-induced hepatomegaly and<br />

ALT increase. Treatment with INT-767 significantly decreased<br />

HFD-induced adipocyte hypertrophy and reduced GLUT4 translocation<br />

to the plasma membrane, both considered hallmarks<br />

of insulin resistance in VAT. Treatment with INT-767 also significantly<br />

induced VAT expression of genes related to nitric oxide<br />

(NO)/cyclic guanosine-3’,5’-monophosphate (cGMP)/protein<br />

kinase G (PKG) signaling, mediating both mitochondriogenesis<br />

and brown adipocyte differentiation. Rabbit preadipocytes<br />

(rPADs), isolated from the different experimental groups, were<br />

investigated for their spontaneous adipogenic potential. The<br />

expression of genes specific for: (i) brown fat (UCP1, CIDEA,<br />

BMP4, BMP7, TMEM26,HOXC9, and LHX8),(ii) mitochondrial<br />

biogenesis (Tfam, NRF1,GCa1, GCb1, PKG1), (iii)membrane<br />

respiratory chain (SLC25A12, NDUFB3, NDUFB5, SDHB,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 655A<br />

(iv) pro-fusion (Mfn2) and pro-fission (Fis1) proteins of mitochondria,<br />

were all significantly increased in rPADs from INT-<br />

767-treated compared to HFD rabbits. Transmission electron<br />

microscopy demonstrated that INT-767 treatment normalized<br />

HFD-induced reduction of mitochondrial cristae. INT-767 treatment<br />

was also able to: (i) improve mitochondrial architecture<br />

and dynamic, with the majority of mitochondria continuously<br />

moving and changing shape, as assessed by time lapse imaging<br />

with mitochondria-targeted fluorescent probe MitoTracker,<br />

(ii) reduce superoxide production, assessed by measuring the<br />

time-dependent accumulation of dihydroethidium-derived fluorescence,<br />

(iii) improve insulin sensitivity assessed by measuring<br />

3 H-2-deoxy-D-glucose uptake in response to increasing insulin<br />

concentrations. In conclusion, the dual FXR/TGR5 agonist INT-<br />

767 ameliorates the metabolic profile and reduces visceral<br />

adiposity by improving insulin sensitivity and promoting brown<br />

differentiation in visceral adipose tissue.<br />

Disclosures:<br />

Luciano Adorini - Employment: Intercept Pharmaceuticals<br />

Mario Maggi - Consulting: BAYER, ELI LILLY, MENARINI, PROSTRAKAN, INTER-<br />

CEPT<br />

The following authors have nothing to disclose: Linda Vignozzi, Ilaria Cellai,<br />

Sandra Filippi, Paolo Comeglio, Tommaso Mello, Daniele Bani, Daniele Guasti,<br />

Erica Sarchielli, Annamaria Morelli, Elena Maneschi, Gabriella Barbara Vannelli<br />

903<br />

CXCL1 promotes steatohepatitis and fibrosis in a mouse<br />

model of high-fat diet-plus-binge ethanol feeding<br />

Zhou Zhou 1 , Ming-Jiang Xu 1 , Binxia Chang 1,2 , Yan Cai 1 , Bin<br />

Gao 1 ; 1 Laboratory of Liver Diseases, National Institute of Alcohol<br />

Abuse and Alcoholism, Rockville, MD; 2 Diagnosis and Treatment<br />

Center for Non-Infectious Liver Diseases, Institute of Alcoholic Liver<br />

Disease, Beijing 302 Hospital, Beijing, China<br />

Background and Aims: Obesity and alcohol abuse are independent<br />

risk factors of human liver diseases. Although some<br />

epidemic <strong>studies</strong> suggested a synergistic effect, the exact function<br />

and mechanism of liver lipid accumulation and alcohol<br />

intake on liver injury is elusive. Methods: In our present study<br />

we developed a murine model of high fat diet (HFD)-plus-binge<br />

ethanol feeding. By using this model, we examined the role of<br />

CXCL1 in steatohepatitis and fibrosis. Results: This HFD-plusbinge<br />

ethanol feeding model induced significant steatohepatitis<br />

and liver fibrosis, as demonstrated by increased expression<br />

of collagen genes and obvious collagen deposition shown in<br />

Sirius Red staining. When we analyzed the mechanisms, we<br />

found that HFD-plus-binge resulted in a significant increase<br />

of neutrophil infiltration in the liver by flow cytometry analysis<br />

of Gr-1 hi CD11b + cells and immunohistochemistry staining of<br />

MPO. Hepatic and serum CXCL1 levels were highly elevated<br />

after HFD-plus-binge ethanol feeding. This upregulation was<br />

mediated by free fatty acid accumulation and the downstream<br />

ERK, JNK and NF-kB pathways. Blocking CXCL1 by genetic<br />

disruption or a neutralizing antibody improved steatohepatitis.<br />

In contrast, overexpressing CXCL1 in hepatocytes with adenovirus<br />

induced significant liver inflammation and injury. Moreover,<br />

hepatocyte derived CXCL1 also contributed to liver fibrosis.<br />

Overexpression of CXCL1 in hepatocytes devastated collagen<br />

deposition in HFD mice, while neutralizing CXCL1 ameliorated<br />

fibrosis induced by HFD-plus-binge ethanol feeding. Conclusions:<br />

Our results revealed a pivotal role of hepatocyte derived<br />

CXCL1 in the liver damage and fibrosis caused by fatty liver<br />

disease and alcohol abuse. Inhibiting the production of CXCL1<br />

or blocking its function might be a useful therapeutic strategy in<br />

preventing alcoholic hepatitis and the related fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Zhou Zhou, Ming-Jiang Xu, Binxia<br />

Chang, Yan Cai, Bin Gao<br />

904<br />

Vitamin D through Induction of Paneth Cell Defensins<br />

Attenuates Gut Dysbiosis and Improves Metabolic Disorders<br />

in Animal Models<br />

Yuan-Ping Han 1 , Danmei Su 1 , Yuanyang Nie 1 , Airu Zhu 1 , Zishuo<br />

Chen 1 , Li Zhang 1 , Pengfei Wu 1 , Mei Luo 1 , Guihui Wu 2 , Richard<br />

Hu 3 , Aurelia Lugea 4 , Jun Xu 5 , Hidekazu Tsukamoto 5 , Stephen J.<br />

Pandol 4 ; 1 College of Life Sciences, Sichuan University, Chengdu<br />

City, China; 2 Chengdu Public Health Clinical Center, Chengdu,<br />

China; 3 Olive View-UCLA Medical Center, Los Angeles, CA;<br />

4 Cedars-Sinai Medical Center, Los Angeles, CA; 5 University of<br />

Southern California, Los Angeles, CA<br />

Metabolic syndrome (MetS) and vitamin D insufficiency/deficiency<br />

are co-prevalent in general population worldwide, but<br />

their causal relationship and mechanisms remain to be determined.<br />

Large body of clinical surveys shows vitamin D insufficiency/deficiency<br />

is associated with type-2 diabetes, obesity,<br />

non-alcoholic steatohepatitis (NASH), and metabolic disorders.<br />

Our previous work also demonstrated dietary vitamin D deficiency<br />

promotes the high fat feeding initiated NASH, in part<br />

through impaired enterohepatic circulation. Since vitamin D<br />

receptor (VDR) is highly expressed in ileum epithelia, we tested<br />

the role of vitamin D signaling in maintaining intestinal integrity<br />

and eubiosis. Balb/C and C57/B6 mice were fed for 4-5<br />

months by (1) AIN93 chow with vitamin D 3<br />

supplementation<br />

as control, or (2) AIN93 with vitamin D depletion, VDD, or (3)<br />

high fat (60% calorie) with vitamin D supplementation, HFD,<br />

and (4) high fat chow with vitamin D depletion, HFD+VDD.<br />

Our results showed that HFD alone was not sufficient to induce<br />

NASH and metabolic disorders, but additional dietary vitamin<br />

D deficiency (HFD+VDD) resulted in severe NASH and insulin<br />

resistance. The interface of ileum—being defined by the intestinal<br />

epithelia and the adjacent gut microbiota in the lumen—<br />

was impaired by the mice fed with HFD+VDD. Specifically, the<br />

Paneth cell-specific defensins including a-defensin 5 (DEFA5),<br />

MMP7 which activates defensins, tight junction genes, and<br />

goblet cell produced mucous protein MUC2 were all thoroughly<br />

suppressed or down regulated in the ileum, resulting<br />

in mucosal collapse, gut leakiness, dysbiosis, endotoxemia,<br />

systemic inflammation, insulin resistance, hepatic steatosis,<br />

and metabolic disorders. Moreover, Helicobacter hepaticus,<br />

a known murine hepatic-pathogen causing chronic hepatitis,<br />

was substantially amplified in within the ileum of the mice fed<br />

with HFD+VDD, while the beneficial symbiotic Akkermansia<br />

muciniphila was diminished. Mice deficient in VDR exhibited<br />

spontaneous NASH and very similar dysbiosis to the mice fed<br />

with HFD+VDD. Remarkably, oral administration of DEFA5 to<br />

the metabolic mice restored gut eubiosis, showing suppression<br />

of Helicobacter hepaticus and reviving Akkermansia muciniphila<br />

within the lumen of ileum, where VD signaling and DEFA5<br />

were impaired by VDD. Along with rebalancing of gut microbiota,<br />

oral administration of DEFA5 resolved hepatic steatosis<br />

in the HFD+VDD fed mice. These results reveal the critical roles<br />

of vitamin D/VDR pathway in optimal expression of defensins,<br />

which in turn regulate intestinal integrity and eubiosis to protect<br />

the host from metabolic disorders including NASH.<br />

Disclosures:<br />

Stephen J. Pandol - Consulting: Calcimedica, Shire, Takeda; Grant/Research<br />

Support: NIH, Department of Veterans Affairs, Calcimedica; Stock Shareholder:<br />

GIRx, Phyteau, Pandol Bros. Inc., Ranch 50


656A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Yuan-Ping Han, Danmei Su,<br />

Yuanyang Nie, Airu Zhu, Zishuo Chen, Li Zhang, Pengfei Wu, Mei Luo, Guihui<br />

Wu, Richard Hu, Aurelia Lugea, Jun Xu, Hidekazu Tsukamoto<br />

905<br />

Fecal Alkaline Phosphatase Is a Determinant Of Metabolic<br />

Endotoxemia And Severity Of Nonalcoholic Fatty<br />

Liver Disease (NAFLD)<br />

Kalyani Daita 1 , Lorenzo Mulazzani 2 , Robert Vincent 1 , Amon<br />

Asgharpour 1 , Sophie C. Cazanave 1 , Mohammad S. Siddiqui 1 ,<br />

Puneet Puri 1 , Michael Idowu 1 , Sherry L. Boyett 1 , Faridoddin Mirshahi<br />

1 , Hae-Ki Min 1 , Robert E. Brown 3 , Masoumeh Sikaroodi 3 ,<br />

Patrick M. Gillevet 3 , Arun J. Sanyal 1 ; 1 VCU, Richmond, VA; 2 Università<br />

di Bologna, Bologna, Italy; 3 George mason university,<br />

Manassas, VA<br />

BACKGROUND: Endotoxemia plays a key role in the genesis of<br />

insulin resistance (IR) and activation of the innate immune system<br />

(IAS) in the metabolic syndrome (Nature 2012; 485:S12).<br />

IR and activation of the IAS are important in the pathogenesis<br />

of NAFLD. Alkaline phosphatase (AP) dephosphorylates<br />

endotoxin (ET) and renders it inactive and intestinal AP (IAP)<br />

is critical to limit ET exposure (AmJPath151.4;1997,1163).<br />

HYPOTHESIS: We hypothesized that fecal IAP (F-IAP) activity<br />

would be inversely related to severity of endotoxemia and the<br />

severity of underlying NAFLD. AIMS: (1) to determine if NASH<br />

was associated with lower fecal IAP and higher endotoxin levels<br />

compared to controls and NAFL, (2) to determine the relationship<br />

between fecal IAP and plasma ET and liver histology<br />

across the spectrum of NAFLD, (3) to determine if changes in<br />

fecal IAP were linked to intestinal inflammation as assessed by<br />

fecal calprotectin, and (4) to identify changes in the intestinal<br />

microbiome associated with altered fecal IAP. METHODS: Subjects<br />

with nonalcoholic fatty liver (NAFL, n= 9) or nonalcoholic<br />

steatohepatitis (NASH, n=15) were compared to 5 weight<br />

matched controls. Fecal IAP was measured by chromogenic<br />

quantification with Dako 5-bromo-4-chloro-3-indoyl-phosphate/<br />

nitroblue with calf AP as an internal control. Plasma ET and<br />

fecal calprotectin were measured by a chromogenic quantification<br />

kit and ELISA respectively. Liver histology was scored<br />

according to NASH CRN criteria. Stool microbiome was analyzed<br />

using 16s pyrosequencing. RESULTS: F-IAP was significantly<br />

lower in NASH compared to NAFL as well as controls<br />

(237.13 U/mg dry weight vs 306.03 and 314.2, p< 0.01<br />

and 0.03 respectively). The plasma AP was similar between<br />

groups. NASH was associated with higher endotoxemia than<br />

both NAFL and controls (1.53 EU/ml vs 1.3 and 1.26, p<<br />

0.02 and 0.03 respectively). There was a tight inverse relationship<br />

between F-IAP and ET levels (r=-0.42, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 657A<br />

Disclosures:<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yusuke Ebisutani, Yoshihiro<br />

Kamada, Sachiho KIda, Kayo Mizutani, Maaya Akita, Akiko Yamamoto,<br />

Hironobu Fujii, Tomoaki Sobajima, Naoko Terao, Shinji Takamatsu, Yuichi<br />

Yoshida, Eiji Miyoshi<br />

907<br />

High Milkfat Diet Causes More Severe NASH Compared<br />

to High Lard Diet: The Dual FXR-TGR5 Agonist INT-767<br />

Prevents NASH in both Models<br />

Xiaoxin Wang 1 , Yuhuan Luo 1 , David J. Orlicky 2 , Luciano Adorini 3 ,<br />

Moshe Levi 1 ; 1 Medicine, Univ Colorado, Aurora, CO; 2 Pathology,<br />

University of Colorado, Aurora, CO; 3 Intercept Pharmaceuticals,<br />

New York, NY<br />

Diets enriched with saturated fatty acids may cause increased<br />

incidence of obesity, insulin resistance, nonalcoholic fatty liver<br />

disease (NAFLD) and nonalcoholic steatohepatitis (NASH). In<br />

this study in C57BL/6J mice we compared the effects of high<br />

milkfat diet (60 kcal% milkfat) versus high lard fat diet (60<br />

kcal% lard fat). We found that both diets induced equal degree<br />

of weight gain, increased fasting glucose, increased serum<br />

insulin, and increased serum cholesterol levels. However,<br />

following a milkfat diet in the liver more inflammatory foci,<br />

lipogranulomas, ballooned cells, and fibrosis were observed.<br />

Oral treatment with the dual FXR-TGR5 agonist INT-767, 10<br />

mg/kg bw/day, or 30 mg/kg bw/day, or 100 mg/kg bw/<br />

day, starting from 8 weeks of age and continuing for 8 months,<br />

in a dose dependent way significantly decreased microsteatosis,<br />

inflammation, fibrosis, and liver cell injury. In mice fed<br />

the milkfat and lard diets, treatment with 100 mg/kg bw/day<br />

of INT-767 was able to abrogate, in all mice tested, hepatic<br />

macrosteatosis, microsteatosis, as well as fibrosis, decreasing<br />

them to control levels (0%). However, in contrast to mice fed the<br />

high lard fat diet where treatment with 100 mg/kg bw/day of<br />

INT-767 was able to decrease inflammation to 0%, in mice fed<br />

the milkfat diet 33% of the livers still demonstrated evidence of<br />

inflammation. INT-767 is therefore able to effectively prevent<br />

NASH in two different models of diet induced obesity and<br />

insulin resistance.<br />

Disclosures:<br />

Luciano Adorini - Employment: Intercept Pharmaceuticals<br />

Moshe Levi - Grant/Research Support: Intercept, Genzyme-Sanofi, Daiichi Sankyo,<br />

Merck, Novartis<br />

The following authors have nothing to disclose: Xiaoxin Wang, Yuhuan Luo,<br />

David J. Orlicky<br />

908<br />

Ornithine transcarbamylase gene expression and<br />

hepatic urea nitrogen handling are reduced in models<br />

of NAFLD and recovers with dietary modulation and<br />

reducing bacterial translocation: Rationale for ammonia<br />

lowering therapy in NASH patients<br />

Karen Louise Thomsen 1 , Francesco De Chiara 1 , Fausto Andreola 1 ,<br />

Krista Rombouts 1 , Jane MacNaughtan 1 , Jane Grove 2 , Guruprasad<br />

P. Aithal 2 , Rajeshwar Mookerjee 1 , Rajiv Jalan 1 ; 1 Institute of Liver<br />

and Digestive Health, University College London, London, United<br />

Kingdom; 2 Nottingham Digestive Diseases Biomedical Research<br />

Unit, Nottingham University Hospital, Nottingham, United Kingdom<br />

Background: Non-alcoholic fatty liver disease (NAFLD) is a<br />

spectrum of liver disease ranging from steatosis, through non-alcoholic<br />

steatohepatitis (NASH) to cirrhosis. In NASH, bacterial<br />

translocation and dysfunctional mitochondria may affect the<br />

function of the mitochondrial urea cycle enzyme ornithine transcarbamylase<br />

(OTC), and result in hyperammonemia, which<br />

has been shown to activate hepatic stellate cells in vivo and in<br />

vitro. This may favour the progression of NAFLD. The aims of<br />

this study were to determine whether gene and protein expression<br />

and function of OTC are altered in animal models of NASH<br />

and NAFLD patients. We also determined if this was reversible<br />

with recovery of animals by restoring the diet and by reducing<br />

bacterial translocation. Methods: Two animal models of NASH<br />

were studied. (a) high-fat, high-cholesterol diet (HFC) for 10<br />

months and then recovery for 2 months (b) methionine-choline<br />

deficient diet (MCD) for 4 weeks treated with or without Yaq-<br />

001 (Yaqrit Ltd.), a nanoporous carbon which has been shown<br />

to reduce bacterial translocation. In humans, we obtained liver<br />

biopsies from 16 NAFLD patients during bariatric surgery and<br />

measured the OTC gene expression. Results: In both animal<br />

models, gene and protein expression of OTC was reduced<br />

significantly. In the HFC rats, reversal of NASH by changing<br />

the diet to normal chow restored OTC gene (0.53 (0.41-0.68)<br />

vs. 0.32 (0.28-0.37), P


658A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

fourth week and twice at 1-week intervals to investigate their<br />

importance in the pathogenesis. Functional effects of Tregs on<br />

lipid metabolism, oxidative stress, and macrophage activation<br />

were then examined. Results: Chronic alcohol feeding induced<br />

liver steatosis and increased serum alanine aminotransferase. A<br />

decrease of hepatice Tregs was observed in these mice. Adaptive<br />

transfer of Tregs was able to abolish hepatic inflammation<br />

and liver injury. Adaptive transfer of Tregs downregulated the<br />

enhanced expression of nuclear Sterol regulatory element-binding<br />

protein 1c (SREBP1c) and its downstream genes induced<br />

in alcohol-fed mice. This intervention also ameliorated hepatic<br />

oxidative stress, increased the level of phosphorylatied adenosine<br />

monophosphate-activated protein kinase (AMPK), peroxisome<br />

proliferator-activated receptor (PPAR) α in the nucleus,<br />

and its downstream genes involved in fatty acid metabolism,<br />

which were inhibited by alcohol-feeding. Furthermore, Tregs<br />

inhibited MCP-1 and TNF- overproduction and macrophage<br />

activation in the liver. In vitro, Tregs suppressed the expression<br />

of MCP-1, TNFα, and CD14 on monocytes/macrophages that<br />

underwent LPS and alcohol co-treatment. This suppression was<br />

markedly abrogated by neutralizing anti-IL-10 mAbs. Conclusions:<br />

Our results indicate that Tregs suppress the development<br />

of AFL, in part through modulating lipid metabolism, oxidative<br />

stress, and the macrophage pro-inflammatory response.<br />

Disclosures:<br />

Qin Ning - Advisory Committees or Review Panels: ROCHE, NOVARTIS, BMS,<br />

MSD, GSK; Consulting: ROCHE, NOVARTIS, BMS, MSD, GSK; Grant/Research<br />

Support: ROCHE, NOVARTIS, BMS; Speaking and Teaching: ROCHE, NOVAR-<br />

TIS, BMS, MSD, GSK<br />

The following authors have nothing to disclose: Hongwu Wang, Ting Wu, Yaqi<br />

Wang, Xiaojing Wang, Xiaoping Luo<br />

910<br />

NOX2 deficiency attenuates ethanol-induced hepatic<br />

steatosis by down-regulating CB1R-mediated lipogenesis<br />

in mice<br />

Jong-Min Jeong, So Yeon Kim, Won-IL Jeong; KAIST, Daejeon,<br />

Korea (the Republic of)<br />

Background: Lines of evidence have suggested that alcohol-mediated<br />

oxidative stress plays important roles in hepatic<br />

steatosis, in which hepatic stellate cells (HSCs) stimulate CB1<br />

receptor (CB1R)-mediated hepatic lipogenesis by producing<br />

an endocannabinoid, 2-arachidonoylglycerol (2-AG). However,<br />

it is not clear yet whether oxidative stress has an effect<br />

on endocannabinoid/CB1R signaling pathway. In the present<br />

study, we investigated the role of NADPH oxidase 2 (NOX2) in<br />

hepatic CB1R signaling pathway during alcoholic steatosis in<br />

mice. Methods: WT and NOX2 deficient (NOX2 -/- ) mice were<br />

fed with liquid ethanol diet for 8 weeks. Sera and liver tissues<br />

of mice were used for biochemical analyses, hitopathological<br />

observation, fluorescence-activated sorting (FACS) analysis<br />

and real-time PCR analyses. LPS treatment, NOX2 siRNA, measurement<br />

of reactive oxygen species (ROS), immunoblotting<br />

and immunostaining were performed in primary hepatic macrophages<br />

and Raw 264.7 cell lines. Results: In histological findings,<br />

livers of NOX2 -/- mice showed less hepatic steatosis than<br />

WT mice. In FACS analyses, populations of Gr1 + CD11b + and<br />

F4/80 + CD11b + cells in WT mice were similar with those of<br />

NOX2 -/- mice. However, gene expression of inflammatory cytokines<br />

in WT mice was significantly higher than that of NOX2 -<br />

/-<br />

mice. In addition, LPS-mediated ROS generation in primary<br />

hepatic and Raw 264.7 macrophages was remarkably attenuated<br />

as NOX2 and toll-like receptor 4 (TLR4) were depleted in<br />

macrophages, suggesting that TLR4 activation was implicated<br />

in NOX2-mediated ROS generation. Furthermore, in co-culturing<br />

macrophages with HSCs, we found that NOX2-mediated<br />

ROS generation in hepatic macrophages increased gene<br />

expression of diacylglycerol lipases, specific enzymes for 2-AG<br />

production, in HSCs, leading to enhance hepatic lipogenesis<br />

through CB1R signaling in hepatocyte. Conclusions: Based on<br />

our data, LPS-mediated NOX2 activation in hepatic macrophage<br />

played a decisive role on alcoholic hepatic steatosis by<br />

stimulating endocannabinoid production of HSCs and CB1R<br />

signaling in hepatocyte. Therefore, NOX2 might be a therapeutic<br />

candidate for alcoholic liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jong-Min Jeong, So Yeon Kim,<br />

Won-IL Jeong<br />

911<br />

L-Fabp deletion attenuates both hepatic steatosis and<br />

fibrosis resulting from hepatic microsomal triglyceride<br />

transfer protein (Mttp) deletion<br />

Elizabeth P. Newberry 2 , Susan M. Kennedy 2 , Yan Xie 2 , Hui<br />

Jiang 2 , Anping Chen 1 , Daniel S. Ory 2 , Nicholas O. Davidson 2 ;<br />

1 St. Louis University, St Louis, MO; 2 Washington University School<br />

of Medicine, St Louis, MO<br />

Genetic or pharmacological inhibition of hepatic microsomal<br />

triglyceride transfer protein (Mttp) blocks VLDL secretion<br />

causing hepatic steatosis. However, the effects on fibrosis are<br />

unknown. Here we asked if liver specific (Alb Cre) Mttp deletion<br />

(Mttp-LKO) promotes hepatic stellate cell activation and<br />

fibrosis and if this phenotype is altered by germline L-Fabp<br />

deletion (DKO mice). Mttp-LKO mice fed low fat chow showed<br />

~10 fold increased hepatic triglyceride (TG) vs C57BL/6J controls<br />

at 12 weeks, with upregulated expression of fibrotic genes<br />

(Col1a1, αSMA, Ctgf) and increased collagen staining (C57,<br />

4.1 ± 0.6 foci/field; Mttp-LKO, 10.1 ± 0.9, n=5-7/group).<br />

Steatosis was significantly attenuated in DKO mice (223 ± 13<br />

mg TG/mg protein vs 331 ± 36, Mttp-LKO, n=6), with reduced<br />

expression of fibrogenic mRNAs and decreased fibrotic staining<br />

(DKO, 5.1 ± 0.5 foci/field). High fat diet fed Mttp-LKO<br />

mice also exhibited 2.5-fold increased hepatic TG, which<br />

was attenuated in DKO mice, again with attenuated Col1a1<br />

gene expression. Expression of lipid droplet proteins Pln2 and<br />

Cidec mRNA were reduced in DKO vs Mttp-LKO, consistent<br />

with reduced steatosis. In vitro FA trafficking revealed ~25%<br />

decreased incorporation of FA into TG in DKO hepatocytes,<br />

yet reduced FA oxidation compared to Mttp-LKO, suggesting<br />

altered FA compartmentalization. ER stress response (Xbp1s,<br />

Atf4, Ern1) was decreased ~40% in DKO livers vs both C57<br />

and Mttp-LKO. Oxidative stress, assessed by lipoperoxide levels,<br />

was increased 4-fold in Mttp-LKO vs DKO (Mttp-LKO, 6.8<br />

± 0.9 nmol LPO/mg protein; DKO, 1.7 ± 0.3, n=4). To understand<br />

the basis for reduced fibrosis in mice lacking L-Fabp, we<br />

undertook broad range lipidomic analyses in livers of Mttp-<br />

LKO and DKO mice. As expected, TG mass was decreased<br />

~25% in DKO, with a disproportionate reduction in TG species<br />

containing essential fatty acids C18:2 and C18:3 (Mttp-LKO,<br />

11.9 ± 0.6; DKO, 8.0 ± 0.6, n=4-5/group, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 659A<br />

a risk for hepatic fibrosis, a concern in light of Phase III clinical<br />

trials of pharmacologic Mttp inhibitors.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elizabeth P. Newberry, Susan<br />

M. Kennedy, Yan Xie, Hui Jiang, Anping Chen, Daniel S. Ory, Nicholas O.<br />

Davidson<br />

912<br />

intrahepatic expression levels of bile acid transporters<br />

are inversely correlated with histological progression of<br />

non-alcoholic fatty liver disease<br />

Kazuya Okushin 1 , Takeya Tsutsumi 2 , Kenichiro Enooku 1 , Hidetaka<br />

Fujinaga 1 , Akira Kado 1 , Kyoji Moriya 3 , Hiroshi Yotsuyanagi 2 ,<br />

Kazuhiko Koike 1 ; 1 Department of Gastroenterology, Graduate<br />

School of Medicine, The University of Tokyo, Tokyo, Japan;<br />

2 Department of Infectious Diseases, Graduate School of Medicine,<br />

The University of Tokyo, Tokyo, Japan; 3 Department of Infection<br />

Control and Prevention, Graduate School of Medicine, The University<br />

of Tokyo, Tokyo, Japan<br />

Background & Aims Non-alcoholic fatty liver disease (NAFLD)<br />

is increasing in the world including Asian countries. NAFLD<br />

contains a disease spectrum ranging from simple steatosis to<br />

nonalcoholic steatohepatitis (NASH). The latter is considered<br />

as a progressive disease, of which pathogenesis remains<br />

largely unclear. Recently, bile acid (BA) metabolism is focused<br />

as a therapeutic target of NASH. The aim of this study was to<br />

identify changes of bile acid metabolism in NAFLD patients<br />

in terms of disease progression. Methods From November<br />

2011 to June 2014, we prospectively enrolled patients with<br />

clinically suspected NAFLD at The University of Tokyo Hospital.<br />

Patients taking ursodeoxycholic acid were excluded in this<br />

study. Disease progression was evaluated by NAFLD activity<br />

score (NAS). Intrahepatic expression levels of genes related<br />

to BA metabolism were determined by quantitative PCR and<br />

immunohistochemistry. This study was approved by the ethics<br />

committees of The University of Tokyo and all patients gave<br />

informed written consent for the study according to the Declaration<br />

of Helsinki. Results Seventy-eight patients (male: female<br />

= 49: 29) histologically diagnosed as NAFLD by significant<br />

steatosis, namely the percentage of hepatocytes showing fatty<br />

changes was ≥ 5% and coincidence with Matteoni classification,<br />

were analyzed. Expression levels of farnesoid X receptor<br />

(FXR) and liver receptor homolog 1 (LRH1), key molecules<br />

associated with BA metabolism, were significantly decreased<br />

along with NAS increase only in female. Small heterodimer<br />

partner (SHP), a downstream transcriptional repressor of FXR,<br />

was similarly changed but only in male. On the other hand,<br />

the level of cholesterol 7 alpha-hydroxylase (CYP7A1), a key<br />

enzyme of BA synthesis, was not changed with the elevation of<br />

NAS in both genders. However, expression levels of an export<br />

transporter, bile salt export pump (BSEP), and an uptake transporter,<br />

Na + /taurocholate cotransporter (NTCP), were significantly<br />

down-regulated as the NAS increases in both genders.<br />

Another export transporter, multidrug resistance-associated<br />

protein 2 (MRP2) was also significantly down-regulated but<br />

only in female. Decreases in protein levels were confirmed by<br />

immunohistochemistry on both BSEP and MRP2. Conclusions<br />

Expression levels of BA export transporters, BSEP and MRP2,<br />

were inversely correlated with NAS in the liver of NAFLD<br />

patients. This down-regulation are supposed to cause excessive<br />

increase of BA levels in hepatocyte, leading to hepatocyte injuries.<br />

Although the mechanism underlying the down-regulation<br />

remains to be elucidated, our findings in patients might lead to<br />

the development of new therapeutic options for NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kazuya Okushin, Takeya Tsutsumi,<br />

Kenichiro Enooku, Hidetaka Fujinaga, Akira Kado, Kyoji Moriya, Hiroshi<br />

Yotsuyanagi, Kazuhiko Koike<br />

913<br />

Non-invasive quantitative decline in liver fat content on<br />

MRI and histologic response in nonalcoholic steatohepatitis:<br />

A secondary analysis of MOZART trial<br />

Janki R. Patel 1 , Ricki Bettencourt 2,3 , Jeffrey Y. Cui 2 , Joanie<br />

Salotti 4,2 , Jonathan Hooker 5 , Archana Bhatt 2 , Carolyn Hernandez<br />

2 , Phirum Nguyen 2 , Hamed Aryafar 5 , William Haufe 5 , Catherine<br />

A. Hooker 5 , Lisa M. Richards 2,4 , Claude B. Sirlin 5 , Rohit<br />

Loomba 2,4 ; 1 Department of Internal Medicine, University of California,<br />

San Diego, La Jolla, CA; 2 NAFLD Translational Research<br />

Unit, University of California, San Diego, La Jolla, CA; 3 Division of<br />

Epidemiology, University of California, San Diego, La Jolla, CA;<br />

4 Division of Gastroenterology, Department of Medicine, University<br />

of California, San Diego, La Jolla, CA; 5 Liver Imaging Group,<br />

Department of Radiology, University of California, San Diego, La<br />

Jolla, CA<br />

Background: Magnetic resonance imaging-estimated proton<br />

density fat fraction (MRI-PDFF) has been shown to be a non-invasive,<br />

accurate and reproducible imaging-based biomarker<br />

for assessing steatosis change and treatment response in nonalcoholic<br />

steatohepatitis (NASH) clinical trials. However, there<br />

are no data on the amount of decline in liver fat shown on MRI-<br />

PDFF that corresponds to histologic response in the setting of a<br />

clinical trial in NASH. We aimed to quantitatively compare the<br />

amount of change in liver fat using MRI-PDFF between histologic<br />

responders vs. histologic non-responders. Methods: This study is<br />

a secondary analysis of the MOZART trial, a randomized, placebo-controlled,<br />

double-blind trial, which included 50 patients<br />

with biopsy-proven NASH to assess efficacy of ezetimibe 10<br />

mg orally daily in reducing liver fat as measured by MRI-PDFF.<br />

All patients enrolled in the trial underwent biochemical testing,<br />

liver biopsy, and MRI at baseline and of those who had both<br />

liver biopsy and MRI-PDFF at weeks 0 and 24 were included in<br />

this analysis. Patients were classified as histologic responders<br />

if they had ≥2 point reduction in NAFLD Activity Score (NAS)<br />

without any increase in fibrosis stage and as histologic non-responders<br />

if they did not meet this criterion. We performed a<br />

head-to-head comparative analysis of histologic responders<br />

and histologic non-responders, and associated quantitative<br />

changes in liver fat as measured via MRI-PDFF. Results: Of the<br />

35 patients who underwent paired liver biopsy and MRI-PDFF<br />

assessment, 10 demonstrated histologic response. Histologic<br />

responders and non-responders had similar baseline demographic<br />

and histological characteristics, with a median age<br />

of 60.5 years and 70% female vs. median age of 49 years<br />

and 64% female, respectively. Compared to histologic non-responders,<br />

histologic responders had a statistically significant<br />

reduction in net MRI-PDFF of -4.1% ± 4.9 vs. +0.6% ± 4.1 (P<br />

< 0.036) with a mean percent change of -29.3% ± 33.0 vs.<br />

+2.0% ± 24.0 (P < 0.004), respectively. Histologic responders<br />

had a significant decrease in hepatic steatosis (P = 0.007) and<br />

hepatocellular ballooning (P = 0.025) compared to histologic<br />

non-responders, and no significant changes were seen in lobular<br />

inflammation and fibrosis scores within or between groups.<br />

Conclusion: Utilizing paired MRI-PDFF and liver histology data<br />

from MOZART Trial, we demonstrate that a 29% reduction in<br />

liver fat on MRI-PDFF is associated with histologic response in<br />

NASH. These novel data can be incorporated into designing<br />

future NASH clinical trials, especially those utilizing change in<br />

liver fat quantified by MRI as an end-point.


660A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Lisa M. Richards - Advisory Committees or Review Panels: Merck; Speaking and<br />

Teaching: AbbVie, BMS, Gilead<br />

Claude B. Sirlin - Grant/Research Support: GE, Pfizer, Bayer, Guerbet, Siemens<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

The following authors have nothing to disclose: Janki R. Patel, Ricki Bettencourt,<br />

Jeffrey Y. Cui, Joanie Salotti, Jonathan Hooker, Archana Bhatt, Carolyn Hernandez,<br />

Phirum Nguyen, Hamed Aryafar, William Haufe, Catherine A. Hooker<br />

914<br />

Mast Cell Involvement in Non-Alcoholic Fatty Liver Disease<br />

in Humans<br />

Anna Christina Dela Cruz 1 , Mehmet M. Altintas 1 , Nestor De La<br />

Cruz-Munoz 2 , Gabriel S. Gaidosh 4 , Leopoldo Arosemena 1 , Mehrdad<br />

Nadji 3 , Monica Garcia-Buitrago 3 , Ali Nayer 1 ; 1 Department of<br />

Medicine, University of Miami Miller School of Medicine, Miami,<br />

FL; 2 Department of Surgery, University of Miami Miller School<br />

of Medicine, Miami, FL; 3 Department of Pathology, University of<br />

Miami Miller School of Medicine, Miami, FL; 4 Bascom Palmer Eye<br />

Institute, University of Miami Miller School of Medicine, Miami, FL<br />

Inflammation plays a critical role in the progression of non-alcoholic<br />

fatty liver disease (NAFLD). Mast cells accumulate in<br />

the liver of rodents with NAFLD and may contribute to the<br />

progression of liver disease. It is not known whether mast cells<br />

are involved in human NAFLD. We examined liver biopsies<br />

from humans with NAFLD (n=42) and controls (n=7). Hepatic<br />

steatosis, necroinflammatory activity, and fibrosis were graded<br />

according to the Batts-Ludwig classification. Mast cells were<br />

identified using c-Kit and tryptase immunostaining. Two pathologists<br />

independently determined the density of c-Kit + mast cells<br />

in the liver in a blinded fashion. To determine the effect of mast<br />

cell deficiency on the development of NAFLD, 6-week-old male<br />

Kit W /Kit W-v and congenic control mice were fed a high-fat diet<br />

(60% energy from fat) for 18 weeks. Data was expressed in<br />

mean±SEM. Student’s t-test and Pearson’s test were used to<br />

analyze the data (GraphPad Prism, version 5.0a). P values of<br />

less than 0.05 were considered statistically significant. There<br />

was significant inter-observer correlation with respect to mast<br />

cell quantification (r=0.78, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 661A<br />

916<br />

An Intact Fgf15 Signaling Pathway is Necessary for<br />

Hepatoprotection after Bariatric Surgery<br />

Andriy Myronovych 1,2 , Kenneth D. Setchell 3 , Brandon Tan 4 ,<br />

Rosa-Maria Salazar-Gonzalez 2 , Wujuan Zhang 3 , Karen K. Ryan 5 ,<br />

Darleen A. Sandoval 1 , Randy J. Seeley 1 , Rohit Kohli 2 ; 1 Surgery,<br />

University of Michigan, Ann Arbor, MI; 2 Gastroenterology, Hepatology<br />

and Nutrition, Cincinnati Children’s Hospital Medical<br />

Center, Cincinnati, OH; 3 Pathology and Laboratory Medicine,<br />

Cincinnati Children’s Hospital Medical Center, Cincinnati, OH;<br />

4 College of Medicine, University of Cincinnati, Cincinnati, OH;<br />

5 Department of Physiology & Membrane Biology, University of<br />

California, Davis, Davis, CA<br />

Introduction: Vertical sleeve gastrectomy (VSG) results in the<br />

reduction of body weight and nonalcoholic fatty liver disease<br />

(NAFLD). Serum bile acids, well-recognized signaling molecules,<br />

are increased after VSG and their nuclear receptor,<br />

farnesoid X receptor (Fxr) plays a role in regulating metabolic<br />

improvements seen following VSG. Intestinal fibroblast growth<br />

factor 15 (Fgf15) is the Fxr target which suppresses liver lipogenesis.<br />

We hypothesized that an intact Fgf15 is necessary for<br />

NAFLD resolution and hepatoprotection after VSG. Methods:<br />

Fgf15 knockout mice (Fgf15 KO) and their wild type (WT)<br />

littermates were fed a high fat diet (HFD, 60 kcal%) to induce<br />

obesity. Mice were randomized into groups, viz.; VSG surgery<br />

on Fgf15 KO (Fgf15 KO VSG) and WT (WT VSG) mice, Sham<br />

surgery on Fgf15 KO (Fgf15 KO Sham) and WT (WT Sham)<br />

mice. Animals received the same HFD thereafter. Mice were<br />

sacrificed 7 weeks post-surgery. Results: Mice post-VSG lost<br />

more weight than their respective Sham groups (p


662A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

918<br />

Wheat Bran Autolytic Peptides, Containing a Branchedchain<br />

Amino Acid, Ameliorate Nonalcoholic Steatohepatitis<br />

with Up-regulation of Anti-oxidant Potential in<br />

High-fat Diet-fed Mice<br />

Takumi Kawaguchi 4 , Takato Ueno 1,3 , Yoichi Nogata 2 , Masako<br />

Hayakawa 3 , Hironori Koga 4,3 , Takuji Torimura 4,3 ; 1 Asakura<br />

Medical Association Hospital, Asakura, Japan; 2 NARO Western<br />

Region Agricultural Research Center, Kagawa, Japan; 3 Liver<br />

Cancer Division, Research Center for Innovative Cancer Therapy,<br />

Kurume University, Kurume, Japan; 4 Division of Gastroenterology,<br />

Department of Medicine, Kurume University School of Medicine,<br />

Kurume, Japan<br />

Backgrounds and Aims: Intake of whole wheat is known to<br />

decrease risk of lifestyle-related diseases. Recently, we have<br />

developed a simple extracting method of antihypertensive peptides<br />

from wheat bran by autolysis reactions. The wheat bran<br />

autolytic peptides (WBAPs) contain leucine, a branched-chain<br />

amino acid, which ameliorates oxidative stress and insulin<br />

resistance. The aims of this study are to investigate therapeutic<br />

efficacy of WBAPs for nonalcoholic steatohepatitis (NASH) in<br />

a mouse model. In addition, we evaluate effects of WBAPs<br />

on oxidative stress and anti-oxidant potential. Material and<br />

Methods: Seven-week-old male C57BL/6 mice were fed with a<br />

high fat diet for 10 weeks to induce NASH, and were simultaneously<br />

administrated with either distilled water (CON; n=5),<br />

or WBAP-1 supplemented water (0.05% leucine-arginine-proline;<br />

n=5), or WBAP-2 supplemented water (0.05% leucine-glutamine-proline;<br />

n=5) ad libitum. Then, they were sacrificed,<br />

and body weight, nonalcoholic fatty liver disease activity score<br />

(NAS), and its components were evaluated. Moreover, severity<br />

of oxidative stress in liver tissues was evaluated by measurement<br />

of derivate of reactive oxygen metabolites (d-ROM) and<br />

biological antioxidant potential (BAP) using free radical elective<br />

evaluator (FREE cape diem, DIACRON INTERNATIONAL<br />

s.r.l.). Results: No significant difference was seen in body<br />

weight between the CON and WBAP-1 groups. While, body<br />

weight in the WBAP-2 group was significantly lower compared<br />

to the CON group (46.0±3.0 vs. 38.0±3.3 g/body, P=0.02).<br />

The grades of steatosis, lobular inflammation, and hepatocyte<br />

ballooning was significantly lower in the WBAP-1 and WBAP-2<br />

groups compared to the CON group. NAS was also significantly<br />

lower in the WBAP-1 and WBAP-2 groups compared to<br />

the CON group (CON; 6.6±0.5, WBAP-1; 1.8±2.0, WBAP-<br />

2; 3.8±2.4, P=0.02). There was no significant difference in<br />

hepatic d-ROM levels among the groups. Although no significant<br />

difference was seen in hepatic BAP levels between the<br />

CON and WABP-2 groups, hepatic BAP levels in the WBAP-1<br />

group were significantly higher compared to the CON group<br />

(3242± 338 vs. 2549±63 mmol/L, P=0.03). Conclusions: We<br />

demonstrated that WBAPs ameliorated NASH in a high-fat<br />

diet-induced mouse model. Furthermore, WBAPs up-regulated<br />

BAP. Since WBAPs can be obtained easily at a low cost,<br />

WBAPs may be clinically applicable agents in patients with<br />

NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takumi Kawaguchi, Takato Ueno,<br />

Yoichi Nogata, Masako Hayakawa, Hironori Koga, Takuji Torimura<br />

919<br />

Hepatic but not Intestinal Specific FXR Deficiency Led to<br />

Accelerated Non-alcoholic Steatohepatitis Development<br />

Bo Kong 1 , Runbin Sun 1,2 , Jianliang Shen 1 , Yang Pan 1 , Jia He 1 ,<br />

Quan Jin 1 , Justin D. Schumacher 1 , Le Zhan 1 , Haoyue Zhou 1 , Xinjie<br />

Qiu 1 , Yongping Wang 3 , Jiye Aa 2 , Jason R. Richardson 4 , Tracy<br />

Anthony 3 , Grace L. Guo 1 ; 1 Department of Pharmacology and Toxicology,<br />

Rutgers University, Piscataway, NJ; 2 Key Laboratory of<br />

Drug Metabolism and Pharmacokinetics, China Pharmaceutical<br />

University, Nanjing, China; 3 Department of Nutritional Sciences,<br />

Rutgers University, New Brunswick, NJ; 4 Department of Environmental<br />

& Occupational Medicine, Robert Wood Johnson Medical<br />

School, Piscataway, NJ<br />

Non-alcoholic fatty liver disease (NAFLD) is a spectrum of<br />

chronic liver diseases ranging from simple hepatic steatosis to<br />

non-alcoholic steatohepatitis (NASH), characterized by steatosis<br />

with inflammation. NAFLD can progress to fibrosis, cirrhosis,<br />

and liver tumor. The mechanisms underlying the transition from<br />

simple steatosis to NASH are still elusive and effective therapies<br />

to treat NASH are unavailable. Our previous study showed<br />

that FXR deficiency and/or altered bile acid homeostasis is<br />

involved in the development of NASH in mice, and FXR ligands<br />

emerge as novel treatment strategy for NASH in humans. However,<br />

the tissue specific roles of FXR in NASH development<br />

are not clear. In this study, wild-type (WT; C57BL/6J), wholebody<br />

FXR knockout (FXR KO), hepatocyte-specific FXR knockout<br />

(FXR Hep-/- ) and enterocyte-specific FXR knockout (FXR Int-/- ) mice<br />

were fed a high-fat diet (HFD) with 60% calories from fat for six<br />

months. Glucose tolerance test and serum liver functional tests,<br />

as well as body weight changes and lipid parameters were<br />

measured for monitoring metabolic syndrome development.<br />

H&E and Sirius red staining were performed to determine the<br />

degree of pathological changes and fibrosis. Hepatic levels of<br />

triglycerides and cholesterols were determined to assess the<br />

degree of liver steatosis. Changes in the expression of genes<br />

involved in lipid metabolism, bile acid homeostasis, inflammation,<br />

fibrogenesis and ER stress were determined using Q-PCR.<br />

Consistent with our previous report, the results showed that HFD<br />

fed WT mice developed steatosis, inflammation, and fibrosis,<br />

resembling human NASH characteristics. In detail, they had<br />

increased body weight, reduced insulin sensitivity, increased<br />

serum and hepatic levels of triglycerides, cholesterols and alanine<br />

aminotransferase (ALT) activities, enhanced expression<br />

of inflammation-related genes (IL-6, TNFα, ICAM-1), ER stress<br />

(Chop) and fibrogenesis-related genes (Collagen 1a1, TIMP-1).<br />

FXR KO mice showed exacerbated phenotype compared to<br />

WT mice on HFD. Furthermore, tissue-specific deletion of the<br />

FXR gene affects NASH severity. Specifically, on HFD FXR Hep-/-<br />

mice manifested severe NASH phenotype comparable to that<br />

in FXR KO mice, but the FXR Int-/- mice showed similar changes<br />

to WT mice. In conclusion, FXR in hepatocytes plays a critical<br />

role in the protection against NASH development. The study<br />

suggests that activation of FXR in hepatocytes may represent a<br />

better strategy for NASH prevention and treatment.<br />

Disclosures:<br />

Grace L. Guo - Consulting: NGM<br />

The following authors have nothing to disclose: Bo Kong, Runbin Sun, Jianliang<br />

Shen, Yang Pan, Jia He, Quan Jin, Justin D. Schumacher, Le Zhan, Haoyue Zhou,<br />

Xinjie Qiu, Yongping Wang, Jiye Aa, Jason R. Richardson, Tracy Anthony


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 663A<br />

920<br />

Deletion of orphan nuclear receptor small heterodimer<br />

partner prevents development of nonalcoholic steatohepatitis<br />

Chunki Kim, Mikang Lee, Yoonkwang Lee; Integrative Medical<br />

Sciences, Northeast Ohio Medical University, Rootstown, OH<br />

An orphan nuclear hormone receptor Small Heterodimer<br />

Partner (SHP) is involved in lipid and bile acid metabolism.<br />

Our earlier <strong>studies</strong> have shown that deletion of SHP protects<br />

the mice from diet-induced obesity and hepatic steatosis. In<br />

order to explore a potential role of SHP in the development<br />

of NASH, we quantified inflammatory gene expression in the<br />

mice fed a westernized diet (WestD) for 6 month. The longterm<br />

diet regimen induced mRNA levels of inflammatory genes<br />

such as TNFa, IL-1b, CXCL-1, and iNOS in wild type (WT)<br />

mice. However, the expression of these genes was significantly<br />

down-regulated in SHPKO counterparts. In order to find out<br />

whether the reduction of inflammatory responses is due to<br />

SHP gene deletion or lower hepatic lipid level, we generated<br />

ApoE/SHP double knockout mice (ApoE/SHPKO). ApoEKO<br />

mouse model has been shown to be susceptible to diet-induced<br />

inflammation but resistant to hepatic steatosis. One week-<br />

WestD regimen significantly induced the expression of the<br />

pro-inflammatory genes in ApoEKO mice but not in ApoE/<br />

SHPKO mice without evidence of hepatic TG accumulation in<br />

either genotype as expected, suggesting SHP, not hepatic TG<br />

level, is responsible for the observed inflammatory responses.<br />

To study NASH development in SHPKO mice further, the mice<br />

were challenged with methionine and choline-deficient (MCD)<br />

diet for 6 weeks. SHPKO mice were strongly protected from<br />

NASH development evidenced by reduced pro-inflammatory<br />

gene expression, reduced F4/80 positive macrophage infiltration,<br />

decreased mRNA levels of fibrosis-related genes such<br />

as collagen 1a1, TGF-b1, TIMP1, and a-SMA, which were all<br />

significantly induced in WT mice. In addition, hepatocyte specific-SHPKO<br />

(AlbCre/floxedSHP) mice challenged with MCD<br />

diet also showed significant protection in NASH development,<br />

as assessed with qPCR, immunofluorescence, Sirius red staining,<br />

compared to floxed SHP counterparts. The current study<br />

identifies the role of SHP as pro-inflammatory in the liver and<br />

will provide new insights into development of NASH and its<br />

possible therapeutic or preventative approaches.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chunki Kim, Mikang Lee, Yoonkwang<br />

Lee<br />

921<br />

Differential carbonylation of proteins in human fatty<br />

and nonfatty NASH<br />

Colin T. Shearn, David J. Orlicky, Dennis R. Petersen; Department<br />

of Pathology, University of Colorado Anschutz Medical Campus,<br />

Aurora, CO<br />

Objective: In the liver, a contributing factor in the pathogenesis<br />

of non-alcoholic fatty liver disease is oxidative stress leading<br />

to the accumulation of highly reactive electrophilic α/β unsaturated<br />

aldehydes. The objective of this study was to determine<br />

if significant differences were evident when evaluating carbonylation<br />

in human end-stage fatty nonalcoholic steatohepatitis<br />

(fNASH) compared to end stage non-fatty NASH (nfNASH).<br />

Methods: Using age-matched pooled hepatic tissue obtained<br />

from healthy humans and patients diagnosed with end stage<br />

nfNASH or fNASH, overall carbonylation was assessed by<br />

immunohistochemistry (IHC) and LC-MS/MS of streptavidin<br />

purified hepatic whole cell extracts treated with biotin hydrazide.<br />

Identified carbonylated proteins were further evaluated<br />

using bioinformatics analyses. Results: Picrosirius red staining<br />

revealed extensive fibrosis in both fNASH and nfNASH which<br />

corresponded with increased 4-HNE staining. Although significantly<br />

elevated when compared to normal hepatic tissue,<br />

no significant differences in overall carbonylation and fibrosis<br />

were evident when comparing fNASH with nfNASH. Mass<br />

spectrometric analysis revealed a total of 184 carbonylated<br />

proteins. Of these, 51 were unique to nfNASH, 32 unique to<br />

fNASH, 33 unique to normal hepatic tissue and 27 common to<br />

all groups. DAVID and Gene Ontology bioinformatic pathway<br />

analysis of hepatic carbonylated proteins revealed a propensity<br />

for increased carbonylation of carboxylic acid catabolic<br />

process, generation of precursor metabolites as well as energy<br />

and proteolysis in nfNASH. Whereas carbonylation of proteins<br />

implicated in protein complex biogenesis and regulation of<br />

apoptosis occurred in fNASH. Using LC-MS/MS analysis, in<br />

nfNASH a nonenaldehyde adduct was identified on Lys 235<br />

on the cytoskeletal protein vimentin. Conclusions: These results<br />

suggest that cellular factors regulating mechanisms of protein<br />

carbonylation may be different depending on pathological<br />

diagnosis of NASH. Furthermore these <strong>studies</strong> are the first to<br />

use LC-MS/MS analysis of carbonylated proteins in human<br />

NAFLD and begin exploring possible mechanistic links with<br />

end stage cirrhosis due to fatty liver disease and oxidative<br />

stress.<br />

Disclosures:<br />

The following authors have nothing to disclose: Colin T. Shearn, David J. Orlicky,<br />

Dennis R. Petersen<br />

922<br />

Lysosomal cholesterol in Kupffer cells, particularly when<br />

oxidized, contributes to murine steatohepatitis<br />

Sofie Walenbergh 1 , Tom Houben 1 , Tim Hendrikx 1 , Patrick van<br />

Gorp 1 , Mike Jeurissen 1 , Marie-Hélène Lenders 1 , Marion J. Gijbels<br />

1 , Jogchum Plat 1 , Marten H. Hofker 2 , Christoph J. Binder 3 ,<br />

Dieter Luetjohann 4 , Fons Verheyen 1 , Ger H. Koek 1 , Ronit Shiri-Sverdlov<br />

1 ; 1 Maastricht University, Maastricht, Netherlands;<br />

2 University Medical Center Groningen, Groningen, Netherlands;<br />

3 Research Center for Molecular Medicine of the Austrian Academy<br />

of Sciences, Vienna, Austria; 4 Institute of Clinical Chemistry and<br />

Clinical Pharmacology, Bonn, Germany<br />

Background and Aims: Recently, the importance of lysosomes<br />

within the metabolic syndrome, including fatty liver disease,<br />

is gaining increasing attention. It has been suggested that<br />

macrophages during atherosclerosis as well as Kupffer cells<br />

(KCs) during hepatic inflammation demonstrate properties of<br />

an acquired lysosomal storage disorder. So far, it is unclear<br />

whether there is a causal relationship between lysosomal cholesterol<br />

accumulation (LCA) in KCs and hepatic inflammation.<br />

Additionally, the specific contribution of the oxidized LDL<br />

(oxLDL) fraction to LCA, its concomitant effect on lysosomal<br />

function and on hepatic inflammation is unexplored. Methods:<br />

Hematopoietic deficiency of the mutant Niemann-Pick type C1<br />

(NPC1 mutant ) protein was used as a tool to induce LCA in KCs<br />

of hyperlipidemic mice. To induce high levels of anti-oxLDL antibodies,<br />

NPC1 mutant -transplanted mice were immunized every<br />

two weeks with heat-inactivated pneumococci until the end<br />

of the experiment. Results: Compared to wildtype, NPC1 mutant<br />

-transplanted mice displayed severe hepatic inflammation<br />

and fibrosis. Anti-oxLDL immunization of NPC1 mutant -transplanted<br />

mice improved cholesterol metabolism, lysosomal dysfunction<br />

and liver inflammation, in contrast to non-immunized<br />

NPC1 mutant -transplanted mice. Conclusions: A direct causal link<br />

exists between LCA in KCs and hepatic inflammation. Rather


664A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

than total cholesterol, we specifically show that oxLDL significantly<br />

contributes to lysosomal dysfunction, cholesterol homeostasis<br />

and the hepatic inflammatory response.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sofie Walenbergh, Tom Houben,<br />

Tim Hendrikx, Patrick van Gorp, Mike Jeurissen, Marie-Hélène Lenders, Marion J.<br />

Gijbels, Jogchum Plat, Marten H. Hofker, Christoph J. Binder, Dieter Luetjohann,<br />

Fons Verheyen, Ger H. Koek, Ronit Shiri-Sverdlov<br />

923<br />

Cholesterol crystallization within hepatocyte lipid droplets<br />

and its role in NASH<br />

George N. Ioannou 2,1 , Savitha Subramanian 1 , Alan Chait 1 , Matthew<br />

M. Yeh 1 , W. G. Haigh 2 , Christopher Savard 2,1 ; 1 University<br />

of Washington, Seattle, WA; 2 Veterans Affairs Puget Sound<br />

Healthcare System, Seattle, WA<br />

Background/Aims We recently reported that cholesterol crystals<br />

were present within the lipid droplets of hepatocytes in<br />

patients and mouse models with NASH. We sought to further<br />

characterize the process of cholesterol crystallization in hepatocyte<br />

lipid droplets and its role in the development of Kupffer<br />

cell (KC) crown-like structures (CLS). Methods C57BL/6J, wildtype<br />

mice were assigned to a high-fat (15%, w/w) diet for<br />

6 months supplemented with 0%, 0.25%, 0.5%, 0.75%, or<br />

1% dietary cholesterol (5 groups, n=12 mice/group). HepG2<br />

hepatoma cells were exposed to varying concentrations of LDL<br />

cholesterol, oleic acid, and ACAT inhibitor, in order to induce<br />

lipid droplet formation and cholesterol crystallization. Results<br />

Fibrosing steatohepatitis developed at a dietary cholesterol<br />

concentration ≥0.5%, whereas mice on a diet of 0% or 0.25%<br />

cholesterol developed only simple steatosis. Hepatic cholesterol<br />

crystals and CLSs were also only observed at a dietary cholesterol<br />

concentration ≥0.5%. CLSs consisted of activated KCs that<br />

surrounded and processed cholesterol-crystal containing remnant<br />

lipid droplets of dead hepatocytes and stained intensely<br />

positive for NLRP3 and activated caspase1. When HepG2<br />

cells were exposed to 2000 mg/ml LDL and 200 mM Oleic<br />

acid for >20 days, a sub-population of cells displayed birefringent<br />

cholesterol crystals around the periphery of large lipid<br />

droplets. Conclusion A specific threshold dietary cholesterol<br />

concentration that leads to cholesterol crystallization within<br />

hepatocyte lipid droplets also leads to CLSs and fibrosing<br />

NASH, suggesting a causative association. Exposure of KCs in<br />

CLSs to cholesterol crystals activates the NLRP3 inflammasome.<br />

We developed an in vitro cell culture model of steatosis and<br />

cholesterol crystallization in HepG2 cells.<br />

Disclosures:<br />

The following authors have nothing to disclose: George N. Ioannou, Savitha<br />

Subramanian, Alan Chait, Matthew M. Yeh, W. G. Haigh, Christopher Savard<br />

924<br />

Fenofibrate and LXRα agonist combination attenuated<br />

intrahepatic inflammation in non-alcoholic fatty liver.<br />

Eun Chul Jang 1 , Dae Won Jun 2 , Seung Min Lee 3 , Yong Kyun<br />

Cho 4 , Sang Bong Ahn 5 ; 1 Department of Occupational and Environmental<br />

Medicine, Soonchunhyang University Cheonan Hospital,<br />

Cheonan, Korea (the Republic of); 2 Department of Internal<br />

Medicine, Hanyang University College of Medicine, Seoul, Korea<br />

(the Republic of); 3 Department of Food and Nutrition, Sungshin<br />

Women’s University, Seoul, Korea (the Republic of); 4 Department<br />

of Internal Medicine, Kangbuk Samsung Hospital, Sungkyunkwan<br />

University, School of Medicine, Seoul, Korea (the Republic of);<br />

5 Department of Internal Medicine, Eulji University School of Medicine,<br />

Seoul, Korea (the Republic of)<br />

Background/Aims: Liver X receptors (LXR) is key transcription<br />

factor in the regulation of lipid and cholesterol metabolism.<br />

in addition, LXRα has been implicated as regulator of inflammation.<br />

LXRα activation is associated with hepatic steatosis<br />

and hyperlipidemia in mice. Fenofibrate, a PPARα agonist and<br />

omega-3, an antihypertriglyceride agent, lead to a reduction<br />

of hyperlipidemia. The aim of this study to investigate whether<br />

concurrent effect of LXRα and fenofibrate or omega-3 can produce<br />

synergic benefits in high-fat diet induced obese mice.<br />

Methods: Normal chow and high-fat diet mice treated with<br />

LXRα agonist (T0901317), or combined fenofibrate or omega-3<br />

for 4 weeks. Hematoxylin and eosin staining was performed<br />

on liver tissue extracts after animal sacrifice. SREBP1c, FAS,<br />

SCD-1, PPAR-α, and MCP-1 mRNA expressions were assessed<br />

with reverse transcription-polymerase chain reaction. Results:<br />

LXRα agonist increase intrahepatic fat amount in normal chow<br />

group. Degree of intrahepatic inflammation was not different<br />

among LXRα agonist, or combined with fenofibrate/omega-<br />

3. In NAFLD model, the combined treatment with fenofibrate<br />

did not increase hepatic steatosis. LXRα agonist and/or fenofibrate<br />

or omega-3 decreased intrahepatic inflammation. In<br />

high-fat diet groups, combined treatment with fenofibrate markedly<br />

reduced the expression of genes involved in lipogenesis,<br />

including srebp-1c, fas and scd1. Furthermore, LXRα agonist<br />

and/or fenofibrate or omega3 treatment decreased expression<br />

of abca-1, abcg5, and abcg8 genes, three vital genes for<br />

cholesterol efflux in diet induced fatty liver model, but those<br />

expressions were opposite in normal chow group. Intraheptic<br />

MCP-1, and TNF-α expression and markers of inflammasome<br />

were also decreased in NAFLD group. Conclusion: LXRα agonist<br />

and fenofibrate combination treatment attenuated hepatic<br />

inflammation in NAFLD model.<br />

Disclosures:<br />

The following authors have nothing to disclose: Eun Chul Jang, Dae Won Jun,<br />

Seung Min Lee, Yong Kyun Cho, Sang Bong Ahn


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 665A<br />

925<br />

A novel transcriptional repressor Hairy and Enhancer<br />

Split 6 mediates hepatic lipid homeostasis through inhibition<br />

of Pparg2 expression<br />

Chunki Kim, Mikang Lee, James P. Hardwick, Yoonkwang Lee;<br />

IMS, Northeastern Ohio Univ. College of Med, Rootstown, OH<br />

Hepatic steatosis is a required prelude for development of<br />

non-alcoholic fatty liver diseases. Pparg is a member of nuclear<br />

hormone receptor superfamily and a master regulator for white<br />

adipocyte differentiation and lipid storage. Uprising hepatic<br />

Pparg2 level reprograms liver for lipid storage instead of for<br />

VLDL secretion or fatty acid oxidation, which leads to hepatic<br />

fat accumulation in pathophysiological conditions such as<br />

obesity and diabetes. Our earlier study suggested that small<br />

heterodimer partner and retinoic acid receptor coordinately<br />

regulates Pparg2 gene expression via directly regulating a<br />

novel transcriptional repressor Hairy and Enhancer Split 6<br />

(Hes6) mRNA level. In the suggested regulatory paradigm,<br />

Hes6 represses Pparg2 gene expression by inhibiting DNAbound<br />

HNF4a transcriptional activity. To explore function of<br />

Hes6 in hepatic lipid mobilization, we overexpressed Hes6<br />

using adenovirus in the livers of mice fed a westernized diet<br />

for 2 months. The targeted overexpression reduced Pparg2<br />

mRNA level by 60% and hepatic triglyceride accumulation by<br />

30% compared to the levels obtained from mice injected with<br />

adenoviral empty vector. We also silenced hepatic Hes6 using<br />

adenoviral shRNA and fed the mice western diet for additional<br />

2 weeks. The adenoviral shHes6 effectively silenced Hes6<br />

gene expression by approximately 90% and increased hepatic<br />

fat accumulation and Pparg2 mRNA level by 50% and 6 fold,<br />

respectively. In subsequent experiments, we utilized transient<br />

transfection and gel mobility shift assays to identify a specific<br />

HNF4a binding element in the Pparg2 promoter. Indeed, an<br />

Hnf4a binding consensus sequence was identified at -903bp<br />

from transcription start site. Deletion or point mutation of the<br />

sequence in a luciferase reporter containing the Pparg2 promoter<br />

abolished Hes6-mediated repression in HepG2 cells.<br />

Chromatin immunoprecipitation and gel mobility shift assays<br />

further confirmed direct recruitment and binding of Hnf4a to<br />

the site. The data strongly prove that expression of Pparg2 is<br />

maintained low by coordinate repression by Hes6 and Hnf4a<br />

in normal liver and thus the Hes6-Hnf4a-Pparg2 transcriptional<br />

axis is considered as a critical determinant for hepatic lipid<br />

homeostasis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chunki Kim, Mikang Lee, James<br />

P. Hardwick, Yoonkwang Lee<br />

926<br />

Effects of Dietary Different Doses of Copper and High<br />

Fructose Feeding on Rat fecal Metabolome<br />

Ming Song 1 , Xiaoli Wei 2 , Xinmin Yin 2 , Dale Schuschke 3 , Imhoi<br />

Koo 2 , Craig J. McClain 1,4 , Xiang Zhang 2 ; 1 Department of Medicine/GI,<br />

University of Louisville, Louisville, KY; 2 Chemistry, University<br />

of Louisville, Louisville, KY; 3 Physiology, University of Louisville,<br />

Louisville, KY; 4 Robley Rex VAMC, Louisville, KY<br />

Background/Aims: The gut microbiota play a critical role in<br />

the pathogenesis of nonalcoholic fatty liver disease (NAFLD)<br />

through altering the gut metabolites and gut barrier function.<br />

Increased fructose consumption and inadequate copper intake<br />

are two critical risk factors in the development of NAFLD. The<br />

aim of this study was to determine the effect of different dietary<br />

doses of copper and high fructose feeding on fecal metabolites<br />

in a rat model. Methods: To gain insights into the role of gut<br />

microbiota, male weanling Sprague-Dawley rats were exposed<br />

to different dietary levels of copper (adequate copper, 6ppm;<br />

marginal copper, 1.5ppm; supplemental copper, 20ppm) with<br />

and without high fructose (30% fructose, w/v, in the drinking<br />

water was given ad lib) intake for 4 weeks. Fecal metabolites<br />

of each rat were then analyzed by comprehensive two-dimensional<br />

gas chromatography time-of-flight mass spectrometry<br />

(GC×GC-TOF MS). In parallel, liver tissues were assessed by<br />

histology and triglyceride assay. Results: Our data showed<br />

that high fructose feeding led to obvious hepatic steatosis in<br />

both marginal copper deficient rats and copper supplementation<br />

rats. Among the 38 metabolites detected with significant<br />

abundance alteration between groups, short chain fatty acids<br />

(SCFAs) were markedly decreased with excessive fructose<br />

intake, irrespective of copper levels. C15:0 and C17:0 long<br />

chain fatty acids (LCFAs), produced only by bacteria, were<br />

increased by either high copper level or high fructose intake.<br />

In addition, increased fecal urea and malic acid paralleled<br />

the increased hepatic fat accumulation. Conclusion: GC×GC-<br />

TOF MS analysis of rat fecal samples revealed distinct fecal<br />

metabolome profiles associated with the dietary high fructose<br />

and copper level, with some metabolites possibly serving as<br />

potential noninvasive biomarkers of fructose induced-NAFLD.<br />

Disclosures:<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Ming Song, Xiaoli Wei, Xinmin<br />

Yin, Dale Schuschke, Imhoi Koo, Xiang Zhang<br />

927<br />

Liver sinusoid endothelial cell derived bone morphogenetic<br />

protein binding endothelial regulator (BMPER)<br />

induces iron overload of high-fat diet induced non-alcoholic<br />

fatty liver mice<br />

Takumu Hasebe 1 , Koji Sawada 1 , Shunsuke Nakajima 1 , Hiroki<br />

Tanaka 2 , Takaaki Ohtake 3 , Mikihiro Fujiya 1 , Yutaka Kohgo 3 ;<br />

1 Medicine, Asahikawa Medical University, Asahikawa, Japan;<br />

2 Gastrointestinal Immunology and Regenerative Medicine, Asahikawa<br />

Medical University, Asahikawa, Japan; 3 Gastroenterology,<br />

International University of Health and Welfare Hospital,<br />

Otawara, Japan<br />

[Background] Excessive irons frequently coexist with non-alcoholic<br />

fatty liver (NAFL), which induces hepatic inflammation<br />

and fibrosis. Down regulation of iron regulatory protein<br />

hepcidin and its inducer bone morphogenetic protein (BMP)<br />

signals can be the central cause of iron overload, however<br />

the mechanism in NAFL is still controversial. To investigate<br />

the mechanism of iron overload in NAFL, we performed transcriptome<br />

analysis using high throughput sequencer. [Methods]<br />

Male C57BL/6 mice were fed on regular or high-fat diet for 16<br />

weeks. Internal iron was evaluated by plasma iron, ferritin or<br />

hepatic iron content. Whole RNA sequencing was performed<br />

as transcriptome analysis using Ion Proton (Life Technologies).<br />

Altered expressions of genes comparing between regular and<br />

high-fat diet mice were as follows; fold change > 1.5, p value<br />

< 0.05 (Student’s t test). Altered gene expressions in mice liver<br />

were confirmed by RT-PCR. Plasma hepcidin concentration was<br />

measured by LC-MS/MS. Localization of protein expression<br />

in mice liver was analyzed by immunofluorescence. Hepatocytes<br />

and liver sinusoid endothelial cells were isolated from<br />

regular mice by collagenase perfusion to assess expressions<br />

of iron regulating molecule by RT-PCR. [Results] High-fat diet<br />

mice showed significant obesity and fatty liver, however neither<br />

hepatic inflammation nor fibrosis. Plasma iron and ferritin were<br />

increased in high-fat diet mice, whereas hepatic iron content


666A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

was not changed. RNA sequencing showed 2314 expression<br />

altered genes of total 38114 analyzed. When we focus on<br />

iron regulatory proteins, we found decreased hepcidin and<br />

increased BMP-SMAD signal related genes: Bmp4, Bmper and<br />

Hfe2. In contrast, the other hepcidin inducing signal: transferrin<br />

receptor signal and IL-6 signal, related genes were not<br />

altered. Plasma hepcidin concentration in high-fat mice was<br />

decreased and the hepcidin inducer SMAD phosphorylation<br />

was decreased. BMP binding endothelial regulator (BMPER), a<br />

BMP inhibiting protein, is increased in high-fat diet mice. Immunofluorescense<br />

of frozen liver section showed BMPER expression<br />

localized on sinusoid lumen. Cell isolation also showed<br />

higher expression of BMPER from liver sinusoid endothelial<br />

cells. [Discussions] Excessive irons in high-fat diet fed mice<br />

were caused by decreased hepcidin. We identified that BMPER<br />

is down-regulating hepcidin via BMP-SMAD signaling inhibition.<br />

We also discovered BMPER is expressed by liver sinusoid<br />

endothelial cells to work on hepcidin producer hepatocyte<br />

extracellularly. This study shows importance of interaction with<br />

hepatocytes and non-parenchymal cells in NAFL pathogenesis.<br />

Disclosures:<br />

Yutaka Kohgo - Grant/Research Support: Chugai Pharm., Novartis Japan, Asahikasei<br />

Medical, Sapporo Beer Co.<br />

The following authors have nothing to disclose: Takumu Hasebe, Koji Sawada,<br />

Shunsuke Nakajima, Hiroki Tanaka, Takaaki Ohtake, Mikihiro Fujiya<br />

928<br />

Analysis of T and myeloid cells in patients with NASH<br />

and diabetes mellitus<br />

Luisa Vonghia 1 , Denis Mogilenko 2,3 , An Verrijken 4,5 , Luc van<br />

Gaal 4,5 , Bart Staels 2,3 , Sven M. Francque 1,5 , David Dombrowicz<br />

2,3 ; 1 Department of Gastroenterology and hepatology, University<br />

Hospital, Antwerp, Belgium; 2 Inserm U1011, Lille, France;<br />

3 Institut Pasteur, University of Lille, Lille, France; 4 Department of<br />

Endocrinology, Diabetology and Metabolism, University Hospital<br />

of Antwerp, Antwerp, Belgium; 5 Laboratory of Experimental Medicine<br />

and Paediatrics, Faculty of Medicine and Health Sciences,<br />

University of Antwerp, Antwerp, Belgium<br />

BACKGROUND: The immune system potentially plays a pivotal<br />

role in the onset of Non-Alcoholic Steatohepatitis (NASH) and<br />

of the associated metabolic disturbances, including diabetes<br />

mellitus (DM). AIM: To study the immune cells in peripheral<br />

blood and differentially expressed genes in the liver of preselected<br />

patients according to presence or absence of NASH and<br />

DM. METHODS: We enrolled 32 patients who underwent liver<br />

biopsy because of suspected NASH. The patients were divided<br />

in 4 predefined groups according to liver biopsy and glucose<br />

parameters: 1) control (NO NASH/NO DM), 2) NASH/NO<br />

DM, 3) NASH/DM, 4) NO NASH/DM. A multicolour flow<br />

cytometry analysis was performed in order to investigate T,<br />

natural killer T cells (NKT), natural killer (NK) cells, monocytes<br />

and dendritic cells. Gene expression in the liver was investigated<br />

by microarray analysis in patients with similar clinical<br />

characteristics. RESULTS: Significant differences between<br />

groups of patients were found among T lymphocyte populations<br />

but not in subsets of myeloid cells. Proportions of CD8 +<br />

cells with enhanced production of pro-inflammatory cytokines<br />

(TNF and INFγ) and cytotoxic molecules (granzyme A and B,<br />

perforin) were increased in patients with NASH and/or DM.<br />

The central memory CD8 + cells (CCR7 + CD45RA - ) were significantly<br />

increased in NASH/DM compared to NASH/NO DM,<br />

whereas the effector (CCR7 - CD45RA + ) and effector memory<br />

(CCR7 - CD45RA - ) CD8 + cells were more potent to produce cytotoxic<br />

molecules in all groups of patients with NASH and DM.<br />

Gene set enrichment analysis evenso revealed dysregulation of<br />

genes associated with activation of CD8 + cells in the liver in<br />

NASH/DM compared to NASH/NO DM. We found a trend<br />

to increased proportion of NKT cells in patients with NASH.<br />

CD4 + T regulatory (Treg) and Th22 cells were increased in<br />

NASH/DM compared to NASH/NO DM patients, while Tregs<br />

where decreased in NASH patients without DM. CONCLU-<br />

SION: T-cell, bot not myeloid cell, alterations are associated<br />

with the presence of NASH regardless of the presence of diabetes.<br />

Both flow cytometry and gene expression demonstrate<br />

that increased effector functions of cytotoxic CD8 + T cells and<br />

CD4 + Th22 and Treg cells are found when NASH is associated<br />

with DM, suggesting that these cells are specifically involved<br />

in the crosstalk between liver inflammation and metabolic dysfunctions<br />

and contribute both to hepatotoxic and hepatoprotective<br />

mechanisms.<br />

Disclosures:<br />

Bart Staels - Advisory Committees or Review Panels: MSD; Consulting: Genfit<br />

The following authors have nothing to disclose: Luisa Vonghia, Denis Mogilenko,<br />

An Verrijken, Luc van Gaal, Sven M. Francque, David Dombrowicz<br />

929<br />

Regulation of hepatocellular senescence by melatonin<br />

during alcoholic liver injury<br />

Jessica S. Garner 1 , Yuyan Han 3,2 , Tiaohao Zhou 3,2 , Kelly McDaniel<br />

2,3 , Ying Wan 1,2 , Tami Annable 2,1 , Nan Wu 3,2 , Julie Venter 3,2 ,<br />

Haibo Bai 1,2 , Shannon S. Glaser 3,2 , Heather L. Francis 1,2 , Fanyin<br />

Meng 2,1 , Gianfranco Alpini 3,2 ; 1 Scott & White Hospital, Texas<br />

A&M HSC College of Medicine, Temple, TX; 2 Central Texas Veterans<br />

Healthcare System, Temple, TX; 3 Texas A&M HSC College of<br />

Medicine, Temple, TX<br />

Background: Chronic alcohol consumption leads to hepatic<br />

DNA damage, cellular senescence and permanent cell cycle<br />

arrest. Melatonin, an endogenously produced neurohormone<br />

secreted by the pineal gland as well as the liver, has a variety<br />

of protective effects during organ injury. However, its effect<br />

and mechanism on the hepatic tissues and cells during alcoholic<br />

liver injury remains to be explored. The objective of this<br />

study was to evaluate the role of melatonin regulated hepatic<br />

senescence phenotype during alcoholic liver injury. Methods:<br />

The mRNA expression of melatonin, its downstream miRNA<br />

(miR-34a) and its upstream enzyme serotonin N-acetyltransferase<br />

(AANAT), was assessed in ethanol and LPS-treated<br />

human hepatocytes (N-Heps), as well as in 5 weeks chronic<br />

alcohol feeding mouse liver specimen (with or without antimiR-34a<br />

Vivo-Morpholino treatment) and control liver tissue<br />

by PCR array and real-time PCR assay. The secretion of melatonin<br />

was verified by ELISA assay. Cellular senescence and<br />

proliferation was evaluated by β-gal activity and MTS assays.<br />

The hepatic expressions of the senescence genes (EGR1, PAI-1<br />

and CCL2), clock circadian genes (PER1, BMAL1 and CRY1),<br />

and miR-34a were also determined by real-time PCR assay.<br />

Results: The total liver histopathology score, beta-gal activity<br />

and miR-34a expression increased after 5 weeks chronic ethanol<br />

feeding relative to control mice, along with the significant<br />

reduction of melatonin and AANAT in isolated liver tissues.<br />

Enhanced expression of the senescence markers EGR1, PAI-1<br />

and CCL2 was demonstrated by senescence PCR array in vivo.<br />

Treatment with ethanol (20 mM) and LPS (20 μg/ml) for 7<br />

days induced significant increases of beta-gal staining, along<br />

with the enhanced expression of cellular senescence markers<br />

EGR1, PAI-1 and CCL2 in cultured human hepatocytes. Treatment<br />

of N-Heps with melatonin (10 -11 M for 7 days) also prevented<br />

alcohol-induced cellular senescence and death, and<br />

subsequently reduced senescence markers EGR1 and CCL2,<br />

as well as miR-34a expression. Furthermore, the expression


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 667A<br />

of melatonin regulated senescence and clock genes, including<br />

EGR1, PAI-1, CCL2, CLOCK, PER1, BMAL1 and CRY1,<br />

was significantly altered in chronic ethanol feeding mice livers<br />

after anti-miR-34a Morpholino treatment relative to the controls.<br />

Conclusion: The finding that melatonin plays a significant<br />

role in the regulation of hepatocellular senescence phenotype<br />

provides the basis for an exciting field in which the melatonin<br />

related upstream enzyme (AANAT), the downstream miRNAs<br />

(miR-34a) and clock circadian genes may be manipulated with<br />

potential therapeutic benefits for alcoholic liver injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jessica S. Garner, Yuyan Han,<br />

Tiaohao Zhou, Kelly McDaniel, Ying Wan, Tami Annable, Nan Wu, Julie Venter,<br />

Haibo Bai, Shannon S. Glaser, Heather L. Francis, Fanyin Meng, Gianfranco<br />

Alpini<br />

930<br />

Hypoxia accelerates fatty acid uptake leading to<br />

increased fat accumulation and inflammation in mice<br />

and in cultured human hepatocyte-derived cells<br />

Agueda Gonzalez-Rodriguez, Gloria Mateo, Ines Soro-Arnaiz,<br />

Mar Torres-Capelli, Ainara Elorza, Julian Aragones, Carmelo<br />

García-Monzón; Liver Research Unit, University Hospital Santa<br />

Cristina, Madrid, Spain<br />

Nonalcoholic fatty liver disease (NAFLD) is the most common<br />

cause of chronic liver disease in Western countries. NAFLD is<br />

strongly associated with overweight/obesity, insulin resistance,<br />

type 2 diabetes (T2DM) and cardiovascular complications;<br />

therefore, it is considered the hepatic component of the metabolic<br />

syndrome. NAFLD is characterized by the progression<br />

from a benign steatosis to more severe liver injuries directly<br />

associated with lipotoxicity such as nonalcoholic steatohepatitis<br />

(NASH), cirrhosis and hepatocellular carcinoma in 10% of<br />

NAFLD patients. Recent evidence indicates that NAFLD severity<br />

is affected by obstructive sleep apnoea syndrome (OSAS), a<br />

recurrent upper-airway obstruction during sleep, leading to<br />

periods of chronic intermittent hypoxia (CIH). In this regard,<br />

dysregulation of the normal oxygen gradient in the liver that<br />

promotes the stabilization of the hypoxia-inducible factors<br />

(HIFs) can induce liver steatosis and inflammation. However,<br />

the pathogenic mechanisms underlying the progression of<br />

NAFLD to NASH in the context of CIH featuring OSAS are not<br />

fully understood. As working hypothesis, the more pronounced<br />

the CIH is, the more pronounced the progression to NASH and<br />

fibrosis. AIM: The purpose of this study was focused on the<br />

molecular mechanisms linking hypoxia to NAFLD/NASH setup.<br />

METHODS: HIF system, CD36 content and liver damage markers<br />

were analyzed in livers from conditional Von Hippel-Lindau<br />

knockout (VHL-KO) mice, which display an overexpression of<br />

HIFs after VHL gene deletion induced by tamoxifen, and in<br />

HepG2 human hepatocytes loaded with palmitic acid submitted<br />

to an hypoxic environment (1% O 2<br />

). RESULTS: As expected,<br />

HIF1 and HIF2 were overexpressed in livers from VHL-KO mice<br />

compared to control mice. Remarkably, hepatic features of<br />

NAFLD and NASH were found in livers from VHL-KO mice<br />

together with increased lipid content. Accordingly, CD36 levels<br />

were upregulated in these mice after VHL deletion. In human<br />

hepatic cells, HIFs were stabilized under hypoxic conditions.<br />

Interestingly, hypoxia itself enhanced fatty acid uptake monitored<br />

by Nile Red staining due to the increase of CD36 translocation<br />

to the plasma membrane. This event was parallel to<br />

an increase of inflammatory markers. Noteworthy, palmitic-induced<br />

lipotoxicity was more pronounced under hypoxic conditions.<br />

CONCLUSIONS: Hypoxia accelerates fatty acid uptake,<br />

largely due to CD36 translocation to the plasma membrane of<br />

hepatocytes, leading to increased fat accumulation and inflammation<br />

in mice and in cultured human hepatocyte-derived cells.<br />

These results suggest that hypoxia could have a key pathogenic<br />

role in the progression of NAFLD to NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Agueda Gonzalez-Rodriguez,<br />

Gloria Mateo, Ines Soro-Arnaiz, Mar Torres-Capelli, Ainara Elorza, Julian Aragones,<br />

Carmelo García-Monzón<br />

931<br />

Predicted prevalence of NAFLD and NASH in a large<br />

population using non-invasive multiparametric MRI<br />

Catherine Kelly, Rajarshi Banerjee; Perspectum Diagnostics Ltd,<br />

Oxford, United Kingdom<br />

Aim: To determine the suitability of multiparametric MRI of<br />

the liver for the assessment of NAFLD in large populations.<br />

Methods: 1000 people, aged 40-69, were recruited from the<br />

electoral register for multiparametric MRI, according to the<br />

LiverMultiScan protocol. Liver disease status was unknown.<br />

Estimates of liver fat fraction (PDFF %) and fibroinflammatory<br />

disease (LIF score) were calculated using LiverMultiScan<br />

software. Population statistics were compared to those in a<br />

previously-reported cohort (Banerjee et al. 2014) with biopsy-proven<br />

NASH (NAS ≥ 5). Results: In the unselected population<br />

sample, 19.4% had steatosis (>5% PDFF), in agreement<br />

with UK population estimates (Preiss & Sattar, 2008). In the reference<br />

NASH cohort, 100% had steatosis (mean PDFF 29.6%).<br />

The mean LIF score in the NASH cohort was 2.85 (s.d. 0.75).<br />

A LIF score > 2 is associated with a higher likelihood of liver-related<br />

clinical events (Pavlides, et al., 2014). In the unselected<br />

group, LIF ranged from 0 to 3.2 with a mean of 0.89, (s.d.<br />

0.3). In a plot of LIF vs PDFF, the NASH cohort can be used<br />

to define the upper right quadrant (PDFF > 5% and LIF > 1.1).<br />

Approximately 4% of the unselected cohort occupy the NASH<br />

quadrant, in agreement with previous estimates of NASH prevalence<br />

(Vernon et al. 2011). 9 of these individuals had LIF > 2.<br />

Conclusion: Estimates of steatosis and NASH in an unselected<br />

population agree with previous prevalence <strong>studies</strong>, suggesting<br />

that multiparametric MRI is a promising technique for the<br />

assessment of NAFLD. These findings encourage the use of this<br />

methodology in other population <strong>studies</strong>, including screening.


668A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

LIF vs. PDFF in unselected (dots) and confirmed NASH (squares)<br />

populations.<br />

932<br />

Akkermansia Muciniphila Is Decreased In Patients With<br />

Non-Alcoholic Fatty Liver Disease<br />

Tarkan Karakan 1 , Ceren Gundogdu 2 , Meltem Yalinaycirak 2 , Mehmet<br />

Ibis 1 ; 1 Gastroenterology, Gazi University, Ankara, Turkey;<br />

2 Medical Microbiology, Gazi University, Ankara, Turkey<br />

Objective The bacterial overgrowth in the intestine, disruption<br />

of the balance in the gut microbiota (dysbiosis) may have an<br />

effect on NAFLD pathogenesis. Previous reports suggest a beneficial<br />

role of Akkermansia muciniphila in obesity and metabolic<br />

syndrome. The purpose of the present study is to compare<br />

the gut microbiota of the patients with NAFLD and the healthy<br />

controls by quantitative Real Time PCR (qPCR) analysis. In<br />

order to understood the potential role of gut dysbiosis and subsequent<br />

translocation of bacterial products, serum endotoxin<br />

levels were also been analyzed. Methods The stool and serum<br />

samples from 52 NAFLD patients and 38 healthy controls have<br />

been collected. qPCR analysis of Akkermansia muciniphila,<br />

Faecalibacterium prausnitzii, Lactobacillus spp., Bifidobacterium<br />

spp., Bacteroides fragilis group was performed. Serum<br />

endotoxin levels were also determined by Chromogenic LAL<br />

Assay. Results Akkermansia muciniphila and Bacteroides fragilis<br />

group were significantly lower in patients with NAFLD as<br />

compared with the healthy control (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 669A<br />

bined with low doses of metformin and sildenafil in the treatment<br />

of NAFLD and NASH.<br />

Disclosures:<br />

Michael Zemel - Board Membership: NuSirt Biopharma; Management Position:<br />

NuSirt Biopharma; Patent Held/Filed: NuSirt Biopharma; Stock Shareholder:<br />

Nusirt Biopharma<br />

Antje Bruckbauer - Employment: NuSirt Biopharma; Patent Held/Filed: NuSirt<br />

Biopharma; Stock Shareholder: NuSirt Biopharma<br />

The following authors have nothing to disclose: Bingzhong Xue, Hang Shi<br />

934<br />

Role of fibroblast growth factor 15 in the development<br />

of high fat diet induced NASH<br />

Justin D. Schumacher 1 , Bo Kong 1 , Yang Pan 2 , Le Zhan 1 , Runbin<br />

Sun 2 , Jiye Aa 2 , Jason R. Richardson 1 , Debra Laskin 1 , Grace L.<br />

Guo 1 ; 1 Toxicology, Rutgers University, Smithtown, NY; 2 Key Laboratory<br />

of Drug Metabolism and Pharmacokinetics, Chinese Pharmaceutical<br />

University, Nanjing, China<br />

With the rise in obesity in Western civilizations, the prevalence<br />

of non-alcoholic steatohepatitis (NASH) is increasing with<br />

an estimated 5-10% of the population affected. The primary<br />

features of NASH include steatosis, inflammation, and fibrosis<br />

often accompanied with metabolic syndrome. Fibroblast<br />

growth factor 15 (Fgf15), an endocrine factor mainly produced<br />

in the distal part of small intestine, emerges to be a critical<br />

factor in regulating bile acid homeostasis, energy metabolism,<br />

and liver regeneration. In order to investigate the effects of<br />

Fgf15 on the development of each the listed features of NASH,<br />

Fgf15-/- mice were bred into a 75% A129 and 25% C57BL/6<br />

background. Four-week old Fgf15-/- and wild-type (WT) control<br />

mice were fed either a high fat diet (HFD) or control diet for<br />

six months. Body weight changes and food consumption were<br />

recorded regularly and a glucose tolerance test was administered<br />

during the fifth month. At the end of feeding, liver, intestine,<br />

and blood samples were collected for gene expression<br />

analysis, determination of serum and tissue lipid composition,<br />

identification of biomarker concentrations, and histology. The<br />

results showed that HFD-fed Fgf15-/- mice had metabolic syndrome<br />

presenting with increased body weight, decreased insulin<br />

sensitivity and increased basal total serum cholesterol levels.<br />

Knockout animals had altered expression of lipid metabolic<br />

enzymes with basal Mtp down-regulated and Acss2 up-regulated<br />

while on a HFD compared to controls. Fgf15 deficiency<br />

had no observed effects on steatosis as WT and Fgf15-/- mice<br />

were found to have comparable hepatic levels of triglycerides<br />

and total cholesterol. However, the Fgf15-/- mice were protected<br />

from the development of hepatic fibrosis revealed by<br />

histology analysis and expression of the genes involved in<br />

fibrosis. No changes were observed in hepatic expression of<br />

inflammatory genes such as Tnf-α or Icam. Lastly, Fgf15-/- mice<br />

fed the HFD had increased bile acid pools and up-regulated<br />

gene expression of ileal Ibabp and hepatic Cyp7a1, Cyp8b1,<br />

and Bsep compared to WT mice. In summary, during NASH<br />

development, Fgf15 deficiency was found to have no effects<br />

on liver steatosis or inflammation, however, led to decreased<br />

insulin tolerance, increased basal serum total cholesterol levels,<br />

altered expression of lipid metabolic enzymes, disrupted bile<br />

homeostasis, and decreased liver fibrosis.<br />

Disclosures:<br />

Grace L. Guo - Consulting: NGM<br />

The following authors have nothing to disclose: Justin D. Schumacher, Bo Kong,<br />

Yang Pan, Le Zhan, Runbin Sun, Jiye Aa, Jason R. Richardson, Debra Laskin<br />

935<br />

Fatty Acid-Stimulated Inflammasome Activation Contributes<br />

to Hepatic Stellate Cell Activation in High Fat/<br />

Clarrie Diet-Fed Mice<br />

Xue-Jing Liu 1 , Na-Na Duan 1 , Ning-Ping Zhang 2,3 , Jia Ding 1,4 ,<br />

Jian Wu 1,2 ; 1 Dept. of Pathogenic Biology/Key Lab of Molecular<br />

Virology, Fudan University Shanghai Medical College, Shanghai,<br />

China; 2 Shanghai Institute of Liver Diseases, Fudan University<br />

Shanghai Medical College, Shanghai, China; 3 Department<br />

of Gastroenterology & Hepatology, Fudan University Affiliated<br />

Zhongshan Hospital, Shanghai, China; 4 Department of Gastroenterology,<br />

Shanghai Jing’an District Central Hospital, Shanghai,<br />

China<br />

BACKGROUND: There is a significant increase in free fatty<br />

acids, especially saturated and oxidized fatty acids, which<br />

contributes to lipotoxicity, oxidant stress, and subsequent<br />

hepatic fibrogenesis in non-alcoholic steatohepatitis (NASH).<br />

However, how fatty acids lead to hepatic stellate cell activation<br />

is lack of experimental evidence. The present study aims to<br />

investigate the role of fatty-acid-mediated inflammasone activation<br />

in a NASH model of high fat/Clarie diet (HFC)-fed mice.<br />

METHODS: Mice were fed HFC diet for 13 weeks, and the<br />

development of NASH was assessed by liver histology. Expression<br />

of inflammasome molecules, NLRP1 and NLRP3 was colocalized<br />

in hepatocytes, Kupffer cells and hepatic stellate cells<br />

(HSCs) in liver section by immunohistochemistry. Palmitic acid<br />

(PA, an oxidized fatty acid)-stimulated activation of inflammasome<br />

molecules and subsequent activation in two human<br />

HSC lines were determined by quantitative RT-PCR. RESULTS:<br />

The body weight and liver weight were significantly increased<br />

in HFC diet-fed mice compared to controls (48.0±1.40 vs.<br />

30.1±0.60g, p


670A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

936<br />

Ancestral starvation-activated caspase-2 promotes obesity,<br />

the metabolic syndrome and Nonalcoholic fatty<br />

liver disease<br />

Mariana V. Machado 1,2 , Gregory A. Michelotti 1 , Guanhua Xie 1 ,<br />

Thiago A. Pereira 1 , Mark Jewell 1 , Richard T. Premont 1 , Anna Mae<br />

Diehl 1 ; 1 Division of Gastroenterology, Duke University Medical<br />

Center, Durham, NC; 2 Gastrenterologia e Hepatologia, Hospital<br />

de Santa Maria, CHLN, Lisbon, Portugal<br />

Background: Obesity is an increasingly prevalent risk factor for<br />

common metabolic disorders such as the metabolic syndrome<br />

(MetS). The MetS promotes nonalcoholic fatty liver disease<br />

(NAFLD). One hypothesis for obesity-related systemic dysfunction<br />

is that overtaxed adipocytes release fatty acids and<br />

pro-inflammatory factors into the circulation to trigger insulin<br />

resistance, changes in pancreatic beta cells, and other manifestations<br />

of the MetS, including diabetes mellitus (DM) and<br />

NAFLD. Caspase-2 is an ancestral caspase that is activated<br />

by fatty acid accumulation to eliminate dispensible cells during<br />

starvation, thereby reducing net energy requirements. Aim: To<br />

evaluate the hypothesis that caspase-2 activation is a critical<br />

link in diet-induced MetS that leads to DM and NAFLD Methods:<br />

Caspase-2 deficient mice and congenic wild type mice<br />

were fed a Western diet (high fat diet enriched with saturated<br />

fatty acids and supplemented with 0.2% cholesterol + fructose<br />

and glucose in the drinking water) for 16 weeks. Metabolic<br />

and hepatic outcomes were evaluated. Results: Caspase-2 deficient<br />

mice fed an obesogenic Western diet were protected from<br />

visceral fat deposition, glucose intolerance/insulin resistance,<br />

pancreatic beta cell hyperplasia, dyslipidemia, and NAFLD.<br />

Adipose tissue in caspase-2 deficient mice was more proliferative,<br />

up-regulated mitochondrial uncoupling proteins, and<br />

was resistant to cell hypertrophy and cell death, suggesting<br />

that caspase 2 deficiency caused “browning” of white adipose<br />

tissue. Compared to congenic controls fed the Western diet,<br />

hepatic accumulation of nonesterified fatty acids was reduced,<br />

de novo lipogenesis was decreased, fatty acid b-oxidation and<br />

lipoprotein export were increased in caspase-2 deficient mice<br />

fed Western diets. Conclusion: Caspase-2 promotes lipoapoptosis<br />

in adipocytes and hepatocytes, resulting in the MetS,<br />

DM, and NAFLD. Given that genetic deficiency of caspase-2<br />

protects mice from the MetS and MetS-related tissue damage,<br />

caspase-2 is a potential therapeutic target for the treatment of<br />

NAFLD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mariana V. Machado, Gregory<br />

A. Michelotti, Guanhua Xie, Thiago A. Pereira, Mark Jewell, Richard T. Premont,<br />

Anna Mae Diehl<br />

937<br />

Antihypertensive therapy improves insulin resistance<br />

and imbalances of interleukin-6 and -10 in spontaneously<br />

hypertensive rats with steatohepatitis<br />

Masaya Kozono 1 , Hirofumi Uto 2,1 , Rie Ibusuki 3,1 , Shiho Arima 1 ,<br />

Kohei Oda 1 , Sho Ijuin 1 , Hiroka Onishi 1 , Haruka Sakae 1 , Kaori<br />

Muromachi 1 , Akihiko Oshige 1 , Seiichi Mawatari 1 , Tsutomu<br />

Tamai 1 , Akihiro Moriuchi 4 , Hirohito Tsubouchi 5 , Akio Ido 1 ; 1 Digestive<br />

and Lifestyle Diseases, Department of Human and Environmental<br />

Sciences, Kagoshima University Graduate School of Medical<br />

and Dental Sciences, Kagoshima, Japan; 2 Center for Digestive<br />

and Liver Diseases, Miyazaki Medical Center Hospital, Miyazaki,<br />

Japan; 3 Department of International Island and Community Medicine,<br />

Kagoshima University Graduate School of Medical and<br />

Dental Sciences, Kagoshima, Japan; 4 Department of HGF Tissue<br />

Repair and Regenerative Medicine, Kagoshima University Graduate<br />

School of Medical and Dental Sciences, Kagoshima, Japan;<br />

5 Kagoshima City Hospital, Kagoshima, Japan<br />

Objective: Nonalcoholic steatohepatitis (NASH) is a hepatic<br />

manifestation of the metabolic syndrome. Features of the metabolic<br />

syndrome such as insulin resistance (IR) and hypertension<br />

are risk factors for both advanced liver disease and cardiovascular<br />

disease in patients with NASH. Immune factors also play<br />

a key role in the pathogenesis of NASH, IR, and hypertension.<br />

However, the molecular mechanisms of these disorders and<br />

the association between them have not been fully elucidated.<br />

In this study, we investigated the effects of severe hypertension<br />

induced by a high-salt diet on the pathological condition of<br />

spontaneously hypertensive rats (SHRs) with steatohepatitis.<br />

Methods: Steatohepatitis was induced by a choline-deficient,<br />

L-amino acid–defined (CDAA) diet. Seven-week-old male SHRs<br />

were randomly divided into five groups: those receiving six<br />

weeks of standard chow with normal salt concentration, followed<br />

by an additional 8 weeks of a standard chow or CDAA<br />

diet with normal salt concentration (control and CDAA groups,<br />

respectively); and those receiving six weeks of standard chow<br />

with high-salt (HS) concentration, followed by a CDAA diet<br />

containing high salt for an additional 8 weeks with or without<br />

the anti-hypertensive agents amlodipine (Aml) or hydralazine<br />

(Hyd) (CDAA+HS, CDAA+HS+Aml, and CDAA+HS+Hyd<br />

groups, respectively). Results: In the CDAA and CDAA+HS<br />

groups, blood pressure was significantly correlated with serum<br />

levels of alanine aminotransferase (ALT), alkaline phosphatase<br />

(ALP), fasting blood glucose, serum insulin, and homeostasis<br />

model assessment-IR (HOMA-IR). Anti-hypertensive therapy<br />

significantly ameliorated elevated ALP, glucose, insulin, and<br />

HOMA-IR. Furthermore, increased serum interleukin (IL)-6<br />

levels following the high-salt CDAA diet were attenuated by<br />

anti-hypertensive therapy. Additionally, serum IL-10 levels were<br />

increased by anti-hypertensive therapy, and the decreased proportion<br />

of CD4+CD25+ T cells and CD4+CD25+Foxp3+ T<br />

cells observed following a high-salt CDAA diet tended to be<br />

restored by amlodipine. Conclusion: These results demonstrate<br />

that anti-hypertensive therapy improved glucose metabolism<br />

and imbalances of cytokine expression in a rat model of hypertension<br />

with steatohepatitis, suggesting that anti-hypertensive<br />

therapy acting through immunological factors may be beneficial<br />

for patients with hypertension–associated NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Masaya Kozono, Hirofumi Uto,<br />

Rie Ibusuki, Shiho Arima, Kohei Oda, Sho Ijuin, Hiroka Onishi, Haruka Sakae,<br />

Kaori Muromachi, Akihiko Oshige, Seiichi Mawatari, Tsutomu Tamai, Akihiro<br />

Moriuchi, Hirohito Tsubouchi, Akio Ido


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 671A<br />

938<br />

Alterations in NK cell phenotype with disease progression<br />

in a murine model of non-alcoholic fatty liver disease<br />

(NAFLD)<br />

Rachel McMahan, Cara Porsche, Lucy Golden-Mason, Hugo R.<br />

Rosen; Gastroenterology, University of Colorado, Aurora, CO<br />

Background: Nonalcoholic fatty liver disease (NAFLD) is associated<br />

with a spectrum ranging from benign accumulation of triglycerides<br />

within the cytoplasm of liver hepatocytes (steatosis)<br />

to a more severe disease, nonalcoholic steatohepatitis (NASH).<br />

While the function of NKT cells in NAFLD progression has been<br />

extensively investigated, there are limited data on the role of<br />

NK cells in this disease. To further clarify the role of these cells<br />

in NAFLD we analyzed NK cell phenotype and function over<br />

time in a murine model of NAFLD. Methods: To investigate the<br />

role of NK cells during NAFLD disease progression we established<br />

a murine model of NAFLD/NASH demonstrating the full<br />

disease spectrum. Mice were fed high fat, high cholesterol, high<br />

sucrose diet containing trans-fats from 8 weeks to 20 weeks.<br />

NAFLD severity at 8, 12 and 20 weeks was evaluated by<br />

liver histology and intrahepatic gene expression. NK cells were<br />

isolated from livers at the indicated time points and stained<br />

for cell surface markers and degranulation. Results: We have<br />

established a murine progressive NAFLD model mimicking the<br />

disease spectrum observed in NASH/NAFLD. After 8 weeks of<br />

feeding the mice had increased hepatic steatosis without significant<br />

inflammation while at the 12 week time point inflammatory<br />

foci were found along with increased macrovesicular steatosis.<br />

By 20 weeks, the mice had also developed significant fibrosis<br />

along with inflammation and steatosis mimicking NASH. As<br />

disease progressed from steatosis to steatohepatitis intrahepatic<br />

NK cells significantly increased expression of activation<br />

markers (CD69, p


672A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

C57BL6 wild-type or RIP3-deficient (RIP3 -/- ) mice were fed highfat<br />

choline-deficient (HFCD) or methionine and choline-deficient<br />

(MCD) diets, with subsequent histological and biochemical<br />

analysis of hepatic damage. In primary murine hepatocytes,<br />

necroptosis and oxidative stress were also assessed after necrostatin-1<br />

treatment or RIP3 silencing. We show that in chronic<br />

liver disease patients, RIP3 levels were significantly increased<br />

and correlated with steatohepatitis histological severity (at least<br />

p < 0.05). Circulating markers of necrosis and TNF-α, as well<br />

as liver MLKL phosphorylation were increased in NAFLD (at<br />

least p < 0.05). Likewise, RIP3 and MLKL protein levels and<br />

TNF-α expression were increased in the liver of HFCD and<br />

MCD diet-fed mice (at least p < 0.05). Moreover, RIP3 and<br />

MLKL sequestration in the insoluble protein fraction of NASH<br />

mice liver lysates, strongly suggesting necroptosis activation,<br />

represented an early event during stetatohepatitis progression<br />

(at least p < 0.05). Functional <strong>studies</strong> in primary murine hepatocytes<br />

established the association between TNF-α-induced RIP3<br />

expression, activation of necroptosis and oxidative stress (p <<br />

0.05). Strikingly, RIP3 deficiency attenuated MCD diet-induced<br />

liver injury, steatosis, inflammation, fibrosis and oxidative stress<br />

(at least p < 0.05). In conclusion, necroptosis is increased in<br />

the liver of NAFLD patients and in experimental models of<br />

NASH. Further, TNF-α triggers RIP3-dependent oxidative stress<br />

during hepatocyte necroptosis. As such, targeting necroptosis<br />

appears to arrest or at least impair NAFLD progression.<br />

Disclosures:<br />

Helena Cortez-Pinto - Advisory Committees or Review Panels: Norgine, Lundbeck;<br />

Speaking and Teaching: Janssen, Gilead Janssen<br />

The following authors have nothing to disclose: Marta B. Afonso, Pedro M.<br />

Rodrigues, Tânia Carvalho, Marta Caridade, Paula Borralho Nunes, Rui E. Castro,<br />

Cecília M. Rodrigues<br />

941<br />

Over-expression of HNF4α attenuated lipotoxicity in<br />

non-alcoholic fatty liver disease model<br />

Jai Sun Lee 1 , Dae Won Jun 2 , Seung Min Lee 3 , Yong Kyun Cho 4 ,<br />

Eun Chul Jang 5 , Sang Bong Ahn 6 ; 1 Department translational medicine,<br />

Hanyang University Graduate School of Biomedical Science<br />

and Engineering, Seoul, Korea (the Republic of); 2 Department<br />

of Internal Medicine, Hanyang University College of Medicine,<br />

Seoul, Korea (the Republic of); 3 Department of Food and Nutrition,<br />

Sungshin Women’s University, Seoul, Korea (the Republic<br />

of); 4 Department of Internal Medicine, Kangbuk Samsung Hospital,<br />

Sungkyunkwan University, School of Medicine, Seoul, Korea<br />

(the Republic of); 5 Department of Occupational and Environmental<br />

Medicine, Soonchunhyang University Cheonan Hospital, Cheonan,<br />

Korea (the Republic of); 6 Department of Internal Medicine, Eulji<br />

University School of Medicine, Seoul, Korea (the Republic of)<br />

Background: Hepatocyte nuclear factor 4α (HNF4α) is known<br />

as a master regulator of liver-specific gene expression. The<br />

effects of HNF4α on non-alcohol fatty liver disease (NAFLD)<br />

at transcriptional level are largely unknown. In this study, we<br />

evaluated the role of HNF4α in NAFLD model. Methods: High<br />

fat (HF) diet was fed to rats for 20 weeks to induce NAFLD. To<br />

make vitro NAFLD model, palmatic acid (PA) treated in HepG2<br />

cells for 24 hours. Later, liver-specific gene expressions were<br />

evaluated. Using microporator, a transcription factor, HNF4α<br />

was over-expressed in HepG2 cells. After transfection, MTT<br />

assay, Tunnel assay, FACS, Nile-red staining, and qPCR gene<br />

evaluation were performed. Results: Hepatic gene expression<br />

of Foxa2 and GATA4 were up regulated in HF diet fed rats<br />

compared to control group. But HNF4α mRNA expression was<br />

decreased in diet induced fatty disease model. In HepG2 cells,<br />

palmatic acid treatment also decreased HNF4α mRNA expression;<br />

however, interestingly, the oleic acid treatment did not<br />

affect the HNF4α mRNA expression. HepG2 cells over-expressing<br />

HNF4α through transfection showed a protective effect<br />

against palmatic acid-induced apoptosis. HNF4α over-expression<br />

was also associated with increased fatty acid oxidation<br />

and VLDL secretion. The HepG2 cells over-expressing HNF4α<br />

also exhibited increased the mRNA expression of enzymes<br />

related to VLDL secretion (MTTP and ApoB) and fatty acid oxidation<br />

(ACOX and CYP4A1). Moreover HNF4α over-expression<br />

was also involved in decreased fatty acid uptake. The<br />

HepG2 cells over-expressing HNF4α also exhibited decreased<br />

the mRNA expression of fatty acid uptake mediator (CD36<br />

and FATP2). Conclusions: HNF4α attenuated lipotoxicity by<br />

increased VLDL secretion and fatty acid oxidation in NASH<br />

model.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jai Sun Lee, Dae Won Jun, Seung<br />

Min Lee, Yong Kyun Cho, Eun Chul Jang, Sang Bong Ahn<br />

942<br />

Thioesterase Superfamily Member 2 (Them2) Reduces<br />

Membrane Fluidity and Promotes Endoplasmic Reticulum<br />

(ER) Calcium Loss in Response to Saturated Free<br />

Fatty Acids: Pathogenic Role in Non-Alcoholic Fatty Liver<br />

Disease (NAFLD)<br />

Baran A. Ersoy, Yingxia Li, David E. Cohen; Medicine, Brigham<br />

and Women’s Hospital and Harvard Medical School, Boston, MA<br />

Background: NAFLD is associated with maladaptive increases<br />

in hepatic glucose production in the setting of excessive nutrition<br />

that are due in part to insulin resistance. Saturated free<br />

fatty acids (FFA) contribute by reducing ER membrane fluidity.<br />

This promotes insulin resistance, which is attributable in part to<br />

depletion of ER calcium and ER stress. In addition, increased<br />

cytosolic calcium levels promote hepatic glucose production<br />

by activating calcium/calmodulin-dependent protein kinase 2<br />

(CaMKII) and amp-activated protein kinase (AMPK). Genetic<br />

ablation of Them2, an acyl-CoA thioesterase that preferentially<br />

converts saturated fatty acyl-CoAs to FFA, protects against<br />

depletion of ER calcium and ER stress in cell culture systems and<br />

reduces hepatic glucose production in high fat fed mice. Aim:<br />

This study was designed to determine the molecular mechanism<br />

by which Them2 expression regulates ER calcium homeostasis<br />

and hepatic glucose production. Methods: HEK 293E cells<br />

were exposed for 6 h to 0.5 mM palmitic acid after treatment<br />

with Them2 or scrambled (control) siRNA. ER calcium release<br />

and reuptake were measured using Fluo-4 following treatment<br />

of cells with the re-uptake inhibitor thapsigargin (2 mM). Total<br />

ER calcium stores were measured following treatment of cells<br />

with 5 mM ionomycin, an ionophore that creates calcium-permeable<br />

membrane pores. Membrane fluidity was measured by<br />

polarization anisotropy of diphenylhexatriene, as well as by<br />

formation of pyrenedecanoic acid eximers using fluorescence<br />

spectrometry. Activation of calcium-sensitive phosphoproteins<br />

that promote hepatic glucose production was assessed by<br />

immunoblot analysis. Results: In the absence of palmitic acid<br />

treatment, Them2 knockdown reduced thapsigargin-mediated<br />

loss of ER calcium by 40%. This was not attributable to reduced<br />

ER calcium stores because knockdown of Them2 expression<br />

did not alter ionomycin-mediated calcium release. Palmitic acid<br />

treatment reduced ER membrane fluidity by 58%, reduced ER<br />

calcium uptake by 11% and increased cytosolic calcium by<br />

2-fold. These effects were abrogated by knockdown of Them2.<br />

Moreover, reductions in cytosolic calcium following Them2<br />

knockdown correlated with decreases in activation of both<br />

CaMKII and AMPK. Conclusions: Them2 mediates the effects of<br />

saturated FFA on ER membrane fluidity and calcium depletion,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 673A<br />

which in turn lead to activation of CaMKII and AMPK. These<br />

findings identify a likely mechanism by which Them2 promotes<br />

hepatic glucose production in response to high fat feeding and<br />

provide a rationale for targeting Them2 in the management of<br />

NAFLD.<br />

Disclosures:<br />

David E. Cohen - Advisory Committees or Review Panels: Merck, Aegerion,<br />

Genzyme; Consulting: Intercept<br />

The following authors have nothing to disclose: Baran A. Ersoy, Yingxia Li<br />

943<br />

Genetic analysis of NAFLD within a Caribbean Hispanic<br />

Population<br />

Devorah Edelman 2 , Bernice Morrow 2 , Maria Delio 2 , Mortadha<br />

Abd 4 , Adam Auton 2 , Tao Wang 3 , Allan W. Wolkoff 1,4 , Harmit S.<br />

Kalia 1,4 ; 1 Medicine, Albert Einstein College of Medicine, Bronx,<br />

NY; 2 Genetics, Albert Einstein College of Medicine, Bronx, NY;<br />

3 Epidemiology, Albert Einstein College of Medicine, Bronx, NY;<br />

4 Gastroenterology & Liver Diseases, Montefiore Medical Center,<br />

Bronx, NY<br />

We performed the first genetic investigation of Non-alcoholic<br />

fatty liver disease (NAFLD) in Hispanic patients of majority<br />

Caribbean descent. A total of 316 individuals including 40<br />

subjects with biopsy-proven NAFLD, 24 ethnically matched<br />

non-NAFLD controls, and a 252 ethnically-mixed random sampling<br />

of Bronx County, New York were analyzed. Genotype<br />

analysis was performed to determine allelic frequencies of 234<br />

known single nucleotide polymorphisms (SNPs) associated with<br />

NAFLD risk based on previous genome-wide association study<br />

(GWAS) and candidate-gene <strong>studies</strong>. Additionally, the entire<br />

coding region of PNPLA3, a gene showing the strongest association<br />

to NAFLD was subjected to Sanger sequencing to identify<br />

all exonic variants in our NAFLD samples and controls. The<br />

highly NAFLD-correlated rs738409 polymorphism in PNPLA3<br />

occurred more frequently in the case population than controls<br />

(OR 2.95, p=0.003), the general population (OR 1.97,<br />

p=0.0001), and would be expected in a random Puerto Rican<br />

population (OR 1.56, p= 0.05) based on data from the 1000<br />

Genomes Project. Results suggest that both rare and common<br />

DNA variations in PNPLA3 and SAMM50 may be correlated<br />

with NAFLD in this small population study, while common DNA<br />

variations in CHUK and ERLIN1, may have a protective interaction.<br />

Common SNPs in ENPP1 and ABCC2 have suggestive<br />

association with fatty liver, but with less compelling significance.<br />

In conclusion, Hispanic patients of Caribbean ancestry<br />

may have different interactions with NAFLD genetic modifiers;<br />

therefore, further investigation with a larger sample size, into<br />

this Caribbean Hispanic population is warranted.<br />

Genotyping Results<br />

Disclosures:<br />

Allan W. Wolkoff - Consulting: Synageva; Grant/Research Support: Merck<br />

The following authors have nothing to disclose: Devorah Edelman, Bernice Morrow,<br />

Maria Delio, Mortadha Abd, Adam Auton, Tao Wang, Harmit S. Kalia<br />

944<br />

Inhibitor Of DNA Binding 2 Gene Induced By High Levels<br />

Of Insulin Contributes To Adipogenesis And Fatty<br />

Liver In A Model Of Type 2 Diabetes Mellitus<br />

Tomoyuki Nemoto, Hidetaka Matsuda, Katsushi Hiramatsu, Yoshihiko<br />

Ozaki, Tatsushi Naito, Kazuto Takahashi, Kazuya Ofuji,<br />

Masahiro Ohtani, Hiroyuki Suto, Yasunari Nakamoto; University<br />

of Fukui, Fukui, Japan<br />

BACKGROUND: Hyperinsulinemia is well known to be associated<br />

with type 2 diabetic patients. It remains unclear how<br />

hyperinsulinemia contributes to progression of obesity, pathogenesis<br />

of fatty liver and development of non-alcoholic steatohepatitis<br />

(NASH). In the previous <strong>studies</strong>, inhibitor of DNA<br />

binding 2 (Id2) gene, a dominant-negative transcriptional<br />

repressor, has been shown to function as a promoting factor<br />

for obesity. Here, we examined, in vitro and in vivo, whether<br />

hyperinsulinemia causes promotion of obesity and pathogenesis<br />

of fatty liver mediated by Id2 using type 2 diabetes<br />

mellitus models. METHODS: Mouse fibroblast NIH-3T3 cells<br />

and human hepatoma HepG2 cells were treated with various<br />

concentrations of insulin, and Id2 mRNA and protein expression<br />

were analyzed by Northern and Western blots. Mouse<br />

preadipocyte 3T3-L1 cells were differentiated to adipocytes in<br />

differentiation cocktail (insulin, 3-isobutyl-1-methylxanthine and<br />

dexamethasone). KK-A y mice were used as a type 2 diabetes<br />

mellitus model. Blood glucose was measured using a glucose<br />

monitor, and serum insulin was measured by ELISA. Histological<br />

examination was performed on liver specimens. KK-A y mice<br />

were crossed with Id2 KO mice, and KK-A y -Id2 KO mice were<br />

obtained. Mice had body weight recorded weekly. RESULTS:<br />

Id2 expression at mRNA and protein levels in NIH-3T3 cells<br />

and HepG2 cells was induced by insulin in a dose-dependent<br />

manner. Id2 induction by insulin was partially inhibited by PI3<br />

kinase inhibitor. 3T3-L1 cells expressed high amount of Id2<br />

protein by addition of differentiation cocktail containing insulin<br />

and differentiated to adipocytes. In contrast, the cells did<br />

not differentiate by the cocktail lacking insulin. In KK-A y mice,<br />

blood glucose and serum insulin (3.0 ng/ml) was significantly<br />

higher compared to non-diabetic 129/Sv mice (0.58 ng/ml)<br />

(P


674A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

945<br />

Human amniotic epithelial cells reduce liver fibrosis<br />

and the incidence of hepatocellular carcinoma in a ‘fast<br />

food’ NAFLD model<br />

Alexander Hodge 1 , Nathan T. Kuk 1 , Ying S. Sun 1 , Kristy Nguyen 1 ,<br />

Rebecca Lim 2 , William Sievert 1 ; 1 Monash Medical Centre,Monash<br />

University, Clayton, VIC, Australia; 2 Hudson Institute of Medical<br />

Research, Melbourne, VIC, Australia<br />

Non-alcoholic fatty liver disease (NAFLD) is associated with<br />

obesity and can lead to liver cirrhosis and hepatocellular carcinoma<br />

(HCC). Currently there are no effective treatments except<br />

lifestyle modification. Sourced from human placentas, human<br />

amniotic epithelial stem cells (hAEC) and hAEC-conditioned<br />

media (hAEC-CM) have anti-inflammatory and anti-fibrotic<br />

properties. We examined the efficacy of these cells in a NAFLD<br />

model. Methods: C57BL/6J mice received a ‘fast food’ diet<br />

(FFD) (21% fat, 2% cholesterol with 42g/L fructose drinking<br />

water). At week 34, mice were given one intraperitoneal (IP)<br />

injection of 2x10 6 hAEC (group 1); two injections of 2x10 6<br />

hAEC four weeks apart (group 2) or thrice weekly IP 400mL<br />

hAEC-CM (group 3). Controls were untreated. Mice were<br />

culled at week 42. Liver fibrosis area was determined by sirius<br />

red staining. Hepatic stellate cell activation was determined by<br />

qPCR for alpha-smooth muscle actin (α-SMA). Hepatic inflammation<br />

was measured by serum ALT and staining for liver macrophages<br />

(F4/80 + ). Livers were evaluated macroscopically<br />

for the presence of HCC and confirmed histologically. Liver<br />

progenitor cell (LPC) numbers were assessed by pan-CK staining.<br />

Computer assisted morphometry was used to quantify the<br />

percentage of positively stained liver (expressed as % area<br />

of liver tissue). Results: Controls showed typical pericellular<br />

hepatic fibrosis and steatosis. Liver fibrosis area was reduced<br />

by 40% in all treatment groups (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 675A<br />

947<br />

Adipose tissue-derived stromal cells preventatively suppressed<br />

pathological progression of murine non-alcoholic<br />

steatohepatitis model.<br />

Masatoshi Yamato 1,2 , Yoshio Sakai 1 , Hatsune Mochida 2 , Akihiro<br />

Seki 2 , Kosuke Ishida 2 , Masao Honda 2,1 , Shuichi Kaneko 2,1 ; 1 Gastoroenterology,<br />

Kanazawa University, Kanazawa, Japan; 2 Disease<br />

Control and Homeostasis, Kanazawa University, Kanazawa,<br />

Japan<br />

Non-alcoholic steatohepatitis (NASH) is characterized by steatosis<br />

in hepatocytes accompanied with persisting inflammation,<br />

developing fibrosis to ultimate cirrhosis. The pathogenesis<br />

of NASH is not fully elucidated, and the fundamental treatment<br />

of NASH is yet to be established. We previously reported that<br />

ADSCs are therapeutically beneficial as being functionally and<br />

histologically repairable of the end-stage cirrhotic mice caused<br />

by NASH. In this study, we analyzed pathological progression<br />

of NASH mice model and succeedingly assessed how<br />

ADSC administration affected on early stage NASH disease<br />

progression.[Materials and Methods] C57Bl/6J mice (female,<br />

10 weeks old) were initiated to be fed with high fat atherogenic<br />

(HF-AT) diet to develop steatohepatitis. On week 1, 2,<br />

3, 4, 8, and 12, we obtained liver tissue samples. ADSCs<br />

were isolated from adipose tissue of C57BL/6J mice by collagenase<br />

I digestion, cultured and expanded. On week 4 and 8,<br />

ADSCs (1×10^5 ) were injected twice every 4 weeks into the<br />

splenic subcapsule, and on week 12, the liver tissue samples<br />

were obtained. These liver tissues were analyzed for gene<br />

expressions by real-time detection PCR (RTD-PCR) and immunohistochemistry.[Results]Hepatocyte<br />

steatosis was observed<br />

from 1 week after initiation of HF-AT feeding, not accompanied<br />

with infiltration of inflammatory cells until 4 week. On week 4,<br />

Gr-1+ inflammatory cells were initially found to be infiltrated<br />

into the portal area and had continuously increased with time<br />

until week 12. TNF-α gene expression of the liver tissue was<br />

also increased consistently with increase of hepatic infiltration<br />

of Gr-1+ cells. Obvious fibrosis had been observed since<br />

week 8. After 8 week, liver fibrosis was extended, and gene<br />

expressions of albumin and type IVa collagen was decreased<br />

and increased, respectively. When ADSCs were injected into<br />

NASH mice, the fibrosis area was less extended on week 12<br />

than that of PBS injection mice group. RTD-PCR analysis of<br />

hepatic tissue showed the substantial maintained albumin gene<br />

expression (p=0.04) and the significant decrease level of Collagena<br />

(p=0.04) and TNFα(p=0.04) gene expressions, which<br />

suggested that ADSC improved protein production, suppressing<br />

fibrosis progression, and hepatic inflammation.[Conclusion]<br />

ADSC treatment prevented pathological progression of liver<br />

fibrosis with persisting hepatic inflammation, and prevented the<br />

decrease of albumin production of the liver during early phase<br />

of NASH development. ADSC treatment is considered to be<br />

preventively and therapeutically efficacious in part to NASH.<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Masatoshi Yamato, Yoshio Sakai,<br />

Hatsune Mochida, Akihiro Seki, Kosuke Ishida, Masao Honda<br />

948<br />

High and Low Capacity Runner (HCR and LCR) rat lines<br />

do not display significant differences in NASH phenotype<br />

when fed a fast food diet<br />

Howard C. Masuoka 1,3 , Oscar Cummings 2 , Naga P. Chalasani 1,3 ,<br />

Lawrence Lumeng 1 ; 1 Medicine, Division of Digestive and Liver DIsorders,<br />

Indiana University, Indianapolis, IN; 2 Pathology and Laboratory<br />

Medicine, Indiana University, Indianapolis, IN; 3 Cellular<br />

and Integrative Physiology, Indiana University, Indianapolis, IN<br />

Exercise and dietary changes are cornerstones of treatment<br />

of non-alcoholic fatty liver disease (NAFLD) including non-alcoholic<br />

steatohepatitis (NASH). Sedentary lifestyle is felt to<br />

predispose individuals to NAFLD/NASH, and improved cardiorespiratory<br />

fitness is associated with lower hepatic steatosis.<br />

High and Low Capacity Runner (HCR and LCR) are established,<br />

extensively phenotyped rat lines generated by selective breeding<br />

for contrasting extremes in maximum aerobic-exercise<br />

capacity (MAC). Relative to the HCR, LCR rats exhibit greater<br />

visceral adiposity, insulin resistance, and levels of circulating<br />

pro-inflammatory hormones but lower home cage activity, and<br />

non-exercise activity thermogenesis. We hypothesized that LCR<br />

rats will display increased features of NASH relatively to HCR<br />

rats when fed a “fast food” (FF) diet high in saturated fat,<br />

cholesterol and fructose. In mice this diet better recapitulates<br />

features of human NASH compared to a high fat diet. We had<br />

custom formulated a FF diet containing a caloric density of 4.61<br />

kcal/gram with a caloric composition of 44.9% fat, 15.1%<br />

protein, and 39.9% carbohydrates. The diet was nutritionally<br />

complete, and contained 30% fructose, 14% lard, 8% coconut<br />

oil, and 2% cholesterol. For 16 weeks 10 LCR and 10 HCR<br />

male rats were fed the FF diet ad libitum. Fasting weights were<br />

obtained weekly. Fasting serum ALT, glucose and insulin were<br />

measured every 4 weeks. MRI images were obtained monthly<br />

to determine adipose, lean, and fluid mass. As expected, LCR<br />

rats gained significantly more weight than HCR rats (P


676A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

949<br />

High calories intake and particulate matter exposition<br />

promote the progression from steatosis to steatohepatitis<br />

and create a permissive enviroment for hepatocellular<br />

carcinoma development<br />

Mirko Tarocchi, Giada Marroncini, Simone Polvani, Sara Tempesti,<br />

Tommaso Mello, Francesca Zanieri, Elisabetta Ceni, Andrea<br />

Galli; Department of Experimental and Clinical Biomedical Sciences,<br />

University of Florence, Florence, Italy<br />

Introduction: Nonalcoholic fatty liver disease (NAFLD) is<br />

becoming the most common chronic liver disease, and the prevalence<br />

is rapidly increasing in developed countries. Nonalcoholic<br />

steatohepatitis (NASH), the severe form of NAFLD, can<br />

also progress to liver cirrhosis and hepatocellular carcinoma.<br />

Recent evidences suggest that environmental factors can trigger<br />

hepatic inflammation and progression of steatosis to NASH.<br />

Aim: We evaluate if a western style diet in association with<br />

chronic urban particulate matter exposition can modifies the<br />

pathogenesis and progression of NASH, and support hepatocellular<br />

carcinoma development. Materials and methods: The<br />

experimental model was created to reproduce urban lifestyle:<br />

C57Bl/6 mice were fed with a western style diet (HFD), and<br />

treated with particular matter (PM) collected from the urban<br />

area of Florence (Italy). After 4 and 8 weeks were performed<br />

the morphologic analysis of liver tissues, the evaluation of<br />

inflammatory cell infiltrate and the collagen deposition; we<br />

evaluate also the effects of PM on cytokine production and oxidative<br />

stress related cellular damage. After 8 weeks part of the<br />

animals was also injected with murine hepatoma cells (2x10 6<br />

Hepa1-6) and sacrificed after additional 2 weeks. Results: Both<br />

the HFD groups developed fat accumulation in the liver, and<br />

at 8 weeks, the NASH score was significantly increased in<br />

HFD-PM group (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 677A<br />

951<br />

Fermented soymilk prevent free fatty acid-induced lipogenesis<br />

and production of reactive oxygen species in<br />

hepatocellular steatosis model<br />

Sang Bong Ahn 1 , Byoung Kwan Son 1 , Dae Won Jun 2 , Yong Kyun<br />

Cho 3 ; 1 Gastroenterology, Eulji University, School of Medicine,<br />

Seoul, Korea (the Republic of); 2 Internal Medicine, Hanyang<br />

University, College of Medicine, Seoul, Korea (the Republic of);<br />

3 Internal Medicine, Kangbuk Samsung Hospital, Sungkyunkwan<br />

University, School of Medicine, Seoul, Korea (the Republic of)<br />

INTRODUCTION: Ingredients of soy and fermented products<br />

have been widely utilized as food supplement for health-enhancing<br />

properties, such as reducing the risk of osteoporosis,<br />

protection of cardiovascular diseases, and prevention of prostate<br />

and breast cancer. This study was carried out to examine<br />

the effects of fermented soymilk (FSM) on the free fatty acid-induced<br />

lipogenesis in an in vitro model of hepatocellular steatosis<br />

model. MATERIALS AND METHODS: HepG2 cells were<br />

incubated with 0.2 mM of palmitic acid (PA) for 24 h to induce<br />

lipogenesis and to accumulate the intracellular lipid accumulation,<br />

which was observed by oil red O and Nile red staining.<br />

The PA treated cells were co-incubated with 0.04~1.0% of<br />

lyophilized FSM, 0.05 mM of genistein, and 50 nM of estrogen,<br />

respectively. Western blot analysis of sterol regulatory element-binding<br />

protein-1 (SREBP-1) and nuclear factor erythroid<br />

2-related factor-2 (NRF-2) were performed to examine the lipogenesis<br />

related extracellular signal-regulated kinase (ERK) pathway.<br />

Cellular reactive oxygen species (ROS) was measured<br />

by the DCFDA assay kit. RESULTS: Lipid accumulations in the<br />

PA and FSM co-incubated cells were significantly decreased<br />

by 0.5% and 1.0% of FSM without cytotoxicity. Treatments of<br />

PA and combining with genistein and estrogen significantly<br />

increased the expressions of SREBP-1. However, FSM co-incubation<br />

significantly attenuated the expression of SREBP-1<br />

in the PA treated cells. In addition, expression of NRF-2 and<br />

phosphorylation of ERK were significantly increased in the PA<br />

and FSM co-incubated cells. PA induced ROS production was<br />

significantly reduced by 1.0% of FSM. Meanwhile, genistein<br />

or estrogen alone did not lead to significant differences in ROS<br />

production. CONCLUSION: Our results show that bioactive<br />

components, except genistein and phytoestrogen, in fermented<br />

soymilk protect hepatocytes against lipid accumulation and<br />

ROS production induced by free fatty acid. These effects may<br />

be mediated by inhibition of SREBP-1 and activation of NFR-2<br />

via ERK pathway in hepatocytes.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sang Bong Ahn, Byoung Kwan<br />

Son, Dae Won Jun, Yong Kyun Cho<br />

952<br />

Increased 53-binding protein 1 nuclear focus expression<br />

in patients with non-alcoholic steatohepatitis<br />

Koji Okamoto 2 , Yuko Akazawa 4 , Hisamitsu Miyaaki 4 , Satoshi<br />

Miuma 4 , Naota Taura 4 , Hisayoshi Kondo 3 , Masahiro Nakashima 1 ,<br />

Kazuhiko Nakao 4 ; 1 Department of tumor and diagnostic pathology,<br />

Atomic bomb disease institute, Nagasaki, Japan; 2 Nagasaki<br />

University School of Medicine, Nagasaki, Japan; 3 Biostatistics Section,<br />

Division of Scientific Data Registry, Department of Radioisotope<br />

Medicine, Atomic Bomb Disease Institute, Nagasaki, Japan;<br />

4 Gastroenterology and Hepatology, Nagasaki University Hospital,<br />

Nagasaki, Japan<br />

Background : Pathogenesis of Non-alcoholic fatty liver disease<br />

(NAFLD) and the mechanisms of HCC development remain<br />

largely unclear. Elevated DNA damage response can result<br />

in genomic instability, which may lead to transformation to<br />

cancer. The p53-binding protein 1 (53-BP1) localizes at the<br />

sites of DNA double-strand breaks by irradiation and rapidly<br />

forms nuclear foci. We have reported that increased number<br />

and size of 53-BP1 nuclear foci expression represents the level<br />

of sporadic genomic instability during cervical carcinogenesis<br />

(Matsuda K, Histopathology. 2011,441-51). However, the<br />

presence of genomic instability in the NAFLD liver, especially<br />

the expression pattern of 53-BP1, is largely uninvestigated.<br />

Our aim was to assess the double-strand breaks based on<br />

53-BP1 expression in hepatocytes of NAFLD liver and free fatty<br />

acid-treated hepatocytes. Methods: Isolated rat primary hepatocytes<br />

were exposed to saturated free fatty acid palmitate (200<br />

μM) for 16 hour. In addition, 23 human liver biopsy samples,<br />

including 5 from normal livers, 5 from simple steatotic livers,<br />

and 13 from livers with NASH, were examined. Cells and<br />

biopsy samples were studied by using co-immunofluorescence<br />

with the anti-53-BP1 and hepatocyte marker (hepatocyte). The<br />

number of 53-BP1 nuclear foci in hepatocytes and its association<br />

with clinical features including levels of aspartate aminotransferase<br />

(AST), alanine aminotransferase (ALT), platelet (PLT),<br />

serum collagen type 4, and serum hyaluronic acid were then<br />

examined. Results: The palmitate treatment in the rat primary<br />

hepatocytes resulted in a significant increase in the number of<br />

53-BP1 nuclear foci (6.01% vs 24.5%, control vs palmitate,<br />

p< 0.05). Furthermore, in the human liver biopsy tissue samples,<br />

the number of 53-BP1 nuclear foci in the hepatocytes was<br />

increased by approximately 6 folds in the NASH livers compared<br />

to the simple steatotic livers, and approximately 10 folds<br />

compared to normal controls (p < 0.01). 53-BP1 nuclear foci<br />

were evenly distributed throughout zone 1 (portal area), zone<br />

2, and zone 3 (hepatic vein area) in the NASH liver. The rate<br />

of hepatocytes with more than two 53-BP1 foci/ nucleus was<br />

positively associated with age (p


678A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

bution. Mat1a encodes the enzyme mainly expressed in normal<br />

liver. Mat1a ablation in mice results in the spontaneous<br />

development of non-alcoholic steatohepatitis (NASH). We<br />

observed that SAMe depletion in Mat1a KO mice had three<br />

main effects on hepatic lipid metabolism: 1) impaired TG (triglyceride)<br />

export via VLDL; 2) impaired mitochondrial FA (fatty<br />

acid) oxidation (as evidenced by membrane depolarization,<br />

downregulation of Phb1 (prohibitin 1, a mitochondrial chaperone<br />

protein) and Mcj/Dnajc15 (endogenous mitochondrial<br />

repressor of respiratory chain), and accumulation of long-chain<br />

acylcarnitines); and 3) increased FA uptake. The convergence<br />

of these three factors induced TG accumulation in LD (lipid<br />

droplets). LD expansion confronts hepatocytes with a high<br />

demand of PC (phosphatidylcholine) molecules to cover the LD<br />

surface since other phospholipids, such as PE (phosphatidylethanolamine),<br />

cannot stabilize LD and prevent coalescence. In<br />

Mat1a KO this situation is aggravated, since SAMe-dependent<br />

PC synthesis via PE methylation is decreased, the PC/PE ratio<br />

reduced and mitochondrial FA oxidation impaired. To put a<br />

brake to this drain of PC molecules to LD, FA are rerouted in<br />

Mat1a KO mice liver to other catabolic (endoplasmic reticulum<br />

and peroxisome oxidation) and biosynthetic (ceramides<br />

synthesis) pathways, causing oxidative stress, inflammation<br />

and fibrosis. SAMe treatment for two months in 8-9 month<br />

old Mat1a KO mice ameliorated mitochondrial dysfunction<br />

(reduces membrane depolarization, improves Phb1 and Mcj<br />

expression, and increases SAMe transport to mitochondria)<br />

improving FA oxidation efficiency (FA and acylcarnitine levels<br />

decrease), which results in a drastic reduction in TG accumulation.<br />

SAMe treatment in Mat1a KO mice resulted in more PC<br />

available for proper membrane function, improving liver lipid<br />

homeostasis, histology (H&E, Sudan red, Sirius red) and liver<br />

injury (ALT, AST).<br />

Disclosures:<br />

The following authors have nothing to disclose: David Fernández, Cristina<br />

Alonso, Sebastiaan van Liempd, Marta Varela, Nicolás Navasa, Ibon Martinez-Arranz,<br />

Patricia Aspichueta, Ana M. Aransay, Juan Anguita, Juan M. Falcon-Perez,<br />

Maria L. Martinez-Chantar, Shelly C. Lu, Jose M. Mato<br />

954<br />

Angiogenesis in the pathogenesis of non-alcoholic fatty<br />

liver disease: a differential gene expression study<br />

Savneet Kaur 1 , Sebastian Vlaic 2 , Reinhard Guthke 2 , Rania Dayoub<br />

3 , Mohsin Hasan 4 , Hamda Siddiqui 1 , Thomas S. Weiss 3 , Shiv<br />

K. Sarin 4 ; 1 Gautam Buddha University, Greater Noida, India;<br />

2 Leibniz Institute for Natural Product Research and Infection Biology,<br />

Jena, Germany; 3 University children hospital (KUNO), University<br />

of Regensburg, Regensburg, Germany; 4 Institute of liver and<br />

biliary sciences, New Delhi, India<br />

Background: Nonalcoholic fatty liver disease (NAFLD) is a progressive<br />

liver disease that ranges from simple steatosis to the<br />

most severe form, nonalcoholic steatohepatitis (NASH). The<br />

onset of a chronic inflammatory reaction marks the progression<br />

from simple steatosis to NASH and the expansion of adipose<br />

tissue is strongly associated with neo-angiogenesis. Henceforth,<br />

in the current study, we investigated correlations between<br />

changes in the expression of lipid metabolism associated genes<br />

and genes which are associated with angiogenesis. Methods:<br />

Liver tissues were obtained from healthy controls, patients with<br />

simple liver steatosis, and with non-alcoholic steatohepatitis<br />

(NASH). Using Affymetrix Gene chips, differentially expressed<br />

genes (DEGs) were identified in 21 human liver tissue samples<br />

including healthy controls (n= 7), patients with simple liver<br />

steatosis (n= 7), and with NASH (n= 8). DEGs were identified<br />

using the R-package limma with a false discovery rate (FDR)-adjusted<br />

p-value of 0.1. Among all the DEGs, a subset of the<br />

DEGs associated with lipid metabolism and angiogenesis were<br />

selected. Further, potential key regulatory factors associated<br />

with both lipid metabolism and angiogenesis were analyzed<br />

to study relationships between these two pathways using a<br />

network inference approach. Results: Among the angiogenic<br />

genes, the genes associated with the chemokine signaling<br />

pathway and extracellular matrix proteins (ECM) were significantly<br />

upregulated in patients with NASH as compared to the<br />

patients with simple liver steatosis and healthy controls. Among<br />

the genes associated with lipid metabolism, genes associated<br />

with arachidonic acid metabolism and lipid biosynthesis were<br />

differentially expressed in NASH patients as compared to the<br />

patients with simple steatosis and controls. The network inference<br />

approach identified two genes, endothelin 1 (EDN1) and<br />

nuclear receptor, NR2F2 as key regulatory genes associated<br />

with both lipid metabolism and neo-vessel formation. The major<br />

targets of these genes included cyclooxygenase-2 (COX-2),<br />

chemokine receptor 4 (CXCR4), matrix metalloproteinases 9<br />

and 2 (MMP-9 and MMP-2). The analysis also inferred a novel<br />

positive correlation between COX-2 and MMP2, suggesting<br />

an important link between accumulation of MMPs and prostaglandin<br />

synthesis by COX-2. Conclusion: The study highlights<br />

the relationships between neovessel formation and lipid metabolism<br />

in pathogenesis of NASH. Further analysis on a larger<br />

cohort of patients and animal <strong>studies</strong> will shed light on how<br />

does angiogenesis leads to dysregulated lipid metabolism or<br />

vice-a-versa in NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Savneet Kaur, Sebastian Vlaic,<br />

Reinhard Guthke, Rania Dayoub, Mohsin Hasan, Hamda Siddiqui, Thomas S.<br />

Weiss, Shiv K. Sarin<br />

955<br />

Glucagon-like peptide-1 potently attenuates steatohepatitis<br />

and liver fibrosis by regulating liver macrophage<br />

infiltration, activation and polarization<br />

Xiaoyu Wang 1 , Shih-Yen Weng 1 , Yong Ook Kim 1 , Olena Molokanova<br />

1 , Thomas Klein 2 , Detlef Schuppan 1,3 ; 1 Institute of Translational<br />

Immunology, Mainz, Germany; 2 Boehringer Ingelheim<br />

Pharma, Biberach an der Riss, Germany; 3 Division of Gastroenterology,<br />

Boston, MA<br />

Background and aims: Glucagon-like peptide-1 (GLP-1)<br />

improves insulin sensitivity via enhanced glucose-dependent<br />

insulin secretion, inhibition of glucagon release, and delayed<br />

gastric emptying following its release into the circulation from<br />

the gut. We aimed to explore the utility of the long-acting<br />

GLP-1 receptor agonist Bydureon (BY) to address both inflammation<br />

and fibrosis in models of NASH (nonalcoholic steatohepatitis)<br />

and biliary fibrosis. Methods: BY was administered<br />

twice weekly by subcutaneous injection of 0.4 or 2 mg/kg<br />

to Mdr2KO mice, and to C57BL/6 mice fed a methionine<br />

and choline deficient (MCD) diet for 4 weeks. Inflammation,<br />

fibrosis, steatohepatitis and especially macrophage function<br />

and polarization were assessed. Results: In both the MCD<br />

and Mdr2KO model, BY significantly decreased serum liver<br />

enzymes, liver collagen content, and fibrosis and inflammation<br />

related transcripts and protein levels (e.g., Col1a1, α-Sma,<br />

CD68, CCL3, and TNFα), while it increased the (anti-inflammatory)<br />

macrophage markers Arg1 and Ym1. BY treatment also<br />

reduced the IHC expression of collagen type III, CD68, F4/80<br />

and caspase3. In FACS analysis MCD livers showed no difference<br />

in CD11c + cells in all the groups, while the total percentage<br />

of F4/80 + macrophages was significantly decreased on<br />

BY treatment, indicating a direct effect of BY on macrophages<br />

but no significant impact on hepatic DC populations. More-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 679A<br />

over, BY induced a significant decrease of liver CD11b + Ly6C hi<br />

inflammatory/fibrogenic myeloid cells and accompanied by<br />

a massive increase of hepatic CD11b + Ly6C lo cells. BY also<br />

reduced the CD68 + total macrophages and increased anti-inflammatory<br />

Ym1 + macrophages in Epididymal tissue and<br />

liver. Overall BY suppressed JNK phosphorylation and Erk1/2<br />

expression to inhibited fibrosis and inflammation. Conclusions:<br />

We show that GLP-1 exerts a direct beneficial effect on liver<br />

macrophages which is accompanied by significant effects on<br />

inflammation fibrosis and steatosis in models of biliary fibrosis<br />

and hepatic lipoapoptosis/NASH. Here it interferes with<br />

JNK signaling. Our data support further clinical evaluation of<br />

the utility of GLP-1R agonists for the treatment of patients with<br />

NASH and liver fibrosis.<br />

Disclosures:<br />

Thomas Klein - Employment: Boehringer Ingelheim<br />

Detlef Schuppan - Consulting: Boehringer Ingelheim, Conatus, GLG, Merck,<br />

Mitsubishi-Tanabe, Takeda, Silence, Glenmark, Isis; Grant/Research Support:<br />

Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Xiaoyu Wang, Shih-Yen Weng,<br />

Yong Ook Kim, Olena Molokanova<br />

Disclosures:<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Rohini Mehta, Michael J. Estep,<br />

Gary Bratthauer, Fanny Monge, Thomas Jeffers<br />

956<br />

Hepatic Estrogen Receptor Negatively Correlates with<br />

Fibrosis Only in Male Patients with Non-Alcoholic Steatohepatitis<br />

(NASH)<br />

Rohini Mehta 2 , Michael J. Estep 2 , Gary Bratthauer 2 , Fanny<br />

Monge 1 , Thomas Jeffers 2 , Zachary D. Goodman 1 , Zobair M.<br />

Younossi 2,1 ; 1 Center For Liver Disease, Department of Medicine,<br />

Inova Fairfax Hospital, Falls Church, VA; 2 Betty and Guy<br />

Beatty Center for Integrated Research, Inova Health Systems, Falls<br />

Church, VA<br />

Background and Aim: Evidence suggests that the pathogenesis<br />

of hepatic fibrosis may be influenced by gender and could<br />

potentially be related to estrogen. Our aim was to examine the<br />

relationship of estrogen receptor alpha (ERa) positive hepatocyte<br />

nuclei to the degree of hepatic fibrosis among male and<br />

female NASH patients. Methods: After informed consent, liver<br />

biopsies and clinical data were obtained from patients with<br />

biopsy-proven NASH (N=32; BMI=45.59 ± 14.65; Male<br />

Gender=50%; Age=45.16 ± 11.03). For immunohistochemistry,<br />

paraffin-embedded liver tissue (5uM) was de-paraffinized,<br />

rehydrated. The Anti-Estrogen Receptor alpha (clone ID5;<br />

AbCam) primary antibody (1:100 dilution) was applied to<br />

the sections. Visual staining was obtained using DAKO EnVision<br />

Dual link system. Peroxidase activity was visualized with<br />

3,3’-diaminobenzidine (DAB). Sections were lightly counterstained<br />

with hematoxylin. Ratio of positive hepatocyte nuclei to<br />

total hepatocyte nuclei were counted (Cell Sens,Olympus software)<br />

in three microscopic areas (60X) for each sample. Group<br />

comparisons by Mann-whitney analysis and Spearman’s correlation<br />

analysis were carried out. A p-value of ≤0.05 was<br />

considered to be statistically significant. Results: ER expression<br />

was detected in 75% (N=24) of patients. ER staining ratio<br />

(positive hepatocyte nuclei to total hepatocyte nuclei) ranged<br />

from 0%-85% in the entire population. Females had a wider<br />

range and a broader central tendency of ER positive hepatocytes<br />

(mean=31.62±27.88%, range: 0%-85.22%,) compared<br />

to males (mean=43.67±19.18, 1.62% - 68.57%, p>0.05).<br />

Males with an ER staining ratio >50% were associated with<br />

lower pericellular fibrosis (r=-0.51, p=0.04). However, this<br />

relationship was not detected in women. Conclusion: Hepatic<br />

ERa expression is negatively associated with degree of pericellular<br />

fibrosis in men but not women with NASH. Further <strong>studies</strong><br />

are needed to establish differential role of ERa in males and<br />

females.<br />

957<br />

Liver-Directed Allosteric Inhibitors of Acetyl-CoA Carboxylase<br />

Reduce Hepatic Steatosis and Improve Dyslipidemia<br />

in Diet-Induced Obese Rat Models and Reduce<br />

Inflammation and Fibrosis in a Cirrhotic Rat Model<br />

Geraldine Harriman 1 , Danielle K. DePeralta 2 , Omeed Moaven 2 ,<br />

Lan Wei 2 , Jeremy R. Greenwood 3 , Sathesh P. Bhat 3 , Kenneth K.<br />

Tanabe 2 , Bryan C. Fuchs 2 , William Westlin 1 , H. James Harwood 1 ,<br />

Rosana Kapeller 1 ; 1 Nimbus Therapeutics, Cambridge, MA; 2 Massachusets<br />

General Hospital, Boston, MA; 3 Schrodinger, New York,<br />

NY<br />

Background: Non-alcoholic steatohepatitis (NASH) affects 16<br />

million people in the US and is projected to surpass hepatitis<br />

C as the leading cause of liver transplant by 2020. To date<br />

there are no approved drugs to treat NASH. It has been proposed<br />

that increased hepatic steatosis can lead to lipotoxicity<br />

and inflammation that drive development of NASH, fibrosis,<br />

cirrhosis, and HCC. The rate-limiting enzyme in fatty acid (FA)<br />

synthesis and oxidation is Acetyl-CoA Carboxylase (ACC).<br />

Pharmacologic inhibition of ACC presents an attractive mode<br />

for reducing de novo lipogenesis and enhancing FA oxidation.<br />

Using state-of-the-art structure-based drug design, we identified<br />

a unique series of allosteric ACC inhibitors that bind to its ‘BC’<br />

domain, blocking enzymatic activity. Here we present data<br />

on the impact of liver-directed, selective ACC inhibition on<br />

the pathological processes involved in the progression from<br />

hepatic steatosis to NASH and cirrhosis. Methods: Liver-selective<br />

allosteric ACC inhibitors, ND-630 and ND-654, were<br />

evaluated to determine the effects of ACC inhibition on the<br />

specific pathologies associated with NASH and HCC including<br />

hepatic de novo lipogenesis, inflammation, and fibrosis.<br />

ND-630 and ND-654 inhibit HepG2 FASyn with EC 50<br />

of<br />

8.7nM and 4nM, and hepatic FASyn in vivo with ED 50<br />

of<br />

0.14mg/kg and 0.3mg/kg, respectively. We tested the efficacy<br />

of these compounds in a high sucrose diet-induce obesity<br />

(HS DIO) rat model and in a diethylnitrosamine (DEN)-induced<br />

rat model of cirrhosis and HCC. Results: Inhibition of ACC<br />

with ND-630 in a rat model of HS DIO led to potent and dose<br />

responsive decrease in plasma triglycerides, cholesterol, free<br />

fatty acids, and basal insulin and reduced hepatic triglyceride<br />

levels with maximal inhibition on Day 14 at 10 mg/kg. Inhibition<br />

of hepatic ACC by ND-654 produced a dose-dependent<br />

decrease in liver fibrosis, as determined by Sirius Red stain,


680A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and a decrease in mRNA for markers of inflammation and stellate<br />

cell activation CD-68 and α-SMA, respectively, in a DEN<br />

model of cirrhosis. Furthermore, ND-654 markedly decreased,<br />

in a dose dependent manner, DEN-induced protein expression<br />

of inflammatory cytokines including MCP1, TNFα and LIF, and<br />

fibrotic markers such as TIMP1 and FGF21 and other markers<br />

including Serpin E1, Leptin and Resistin. Conclusions: These<br />

data demonstrate that liver-directed inhibition of ACC can<br />

impact multiple steps in the pathogenesis of fatty liver disease<br />

to NASH and cirrhosis .<br />

Disclosures:<br />

Geraldine Harriman - Employment: Nimbus Discovery<br />

Jeremy R. Greenwood - Employment: Schrodinger Inc.; Patent Held/Filed: Nimbus<br />

Therapeutics; Stock Shareholder: Schrodinger Inc.<br />

Kenneth K. Tanabe - Patent Held/Filed: EGF SNP and risk for HCC, EGFR inhibition<br />

and HCC, gene signature for prognosis in cirrhosis<br />

Bryan C. Fuchs - Consulting: Collagen Medical; Grant/Research Support: Nimbus<br />

Therapeutics<br />

Rosana Kapeller - Employment: Nimbus Therapeutics; Stock Shareholder: Aileron<br />

Therapeutics<br />

The following authors have nothing to disclose: Danielle K. DePeralta, Omeed<br />

Moaven, Lan Wei, Sathesh P. Bhat, William Westlin, H. James Harwood<br />

958<br />

Metformin targets a phosphoSTAT3-miRNAs pathway<br />

to inhibit lipid droplets accumulation and intracellular<br />

inflammation in vitro and in vivo<br />

Natalia Pediconi 1 , Silvia Di Cocco 3 , Silvia Piconese 3 , Federica<br />

Mori 4 , Laura Belloni 2 , Abigail D. Nunn 2 , Tullio Scopigno 2 , Vincenzo<br />

Barnaba 3 , Giovanni Blandino 4 , Sabrina Strano 4 , Massimo<br />

Levrero 3,2 ; 1 Dept of Molecular Medicine, Sapienza University,<br />

Rome, Italy; 2 Center for Life NanoScience (CNLS), IIT-Sapienza,<br />

Rome, Italy; 3 Dept of Internal Medicine - DMISM, Sapienza University,<br />

Rome, Italy; 4 Translational Oncogenomic Unit, Regina Elena<br />

Cancer Institute, Rome, Italy<br />

Background and aim: Accumulation of triglyceride-containing<br />

lipid droplets (LDs) within hepatocytes in NAFLD patients is a<br />

potentially reversible process, although sustained activation of<br />

inflammatory signaling pathways leads to non-alcoholic steatohepatitis<br />

(NASH) that can evolve into cirrhosis and HCC.<br />

The prevalence of NAFLD is increased with aging, however<br />

the molecular mechanism for aging-induced fatty liver remains<br />

poorly understood. Here we investigated the role of a phosphoSTAT3-miRNAs<br />

pathway in LD accumulation and the<br />

activation of intracellular inflammatory pathways in vescicular<br />

steatosis. Methods: DMSO-differentiated human non-transformed<br />

HepaRG cells treated with oleic acid were used as a<br />

cellular model of vescicolar steatosis. C57BL/6 mice treated or<br />

not with metformin for 78 weeks were used as an in vivo model<br />

of aging-induced fatty liver. Results: dHepaRG cells treated<br />

with oleic acid show: a) an accumulation of intracellular lipids<br />

with an increase of both total cellular lipid volume and lipid<br />

droplets number, as evaluated by picosecond CARS (coherent<br />

anti-stoke Raman spettroscopy) mycroscopy; b) the generation<br />

of reactive oxygen species (ROS); c) deregulated lipid metabolism<br />

and liver-specific genes, including PDK4, PLIN4, SLC2A1,<br />

ALB and ALDOB; d) activation of an intracellular inflammatory<br />

response, with the upregulation of NFkB and IFN-a regulated<br />

genes and the activation of the phosphoSTAT3/IL6 pathway.<br />

Treatment of oleate-exposed HepaRG with the S3I STAT3<br />

inhibitor reduces bothe tryglicerides and ROS accumulation<br />

and inhibits phosphoSTAT3 accumulation. We also found that<br />

several STAT3/IL6 responsive miRNAs, including miR21 and<br />

miR24, are upregulated after lipid overload, paralleling STAT3<br />

activation. Chromatin immuno-precipitation (ChIP) experiments<br />

showed that the oleate-dependent transcriptional deregulation<br />

of these miRNAs correlates with phosphoSTAT3 bound to their<br />

promoters. Treatment with S3I inhibits both oleate-dependent<br />

miR21 and miR24 upregulation and phosphoSTAT3 binding to<br />

their promoter. Treatment with metformin, an AMPK activator<br />

widely used as anti-diabetic drug and known to reduce liver<br />

steatosis in vivo, reduces phosphoSTAT3 accumulation, inhibits<br />

miRNAs upregulation and reduces trygliceride accumulation in<br />

fatty HepaRG. Importantly, a chronic 78 weeks treatment with<br />

metformin prevents aging-induced steatosis in C57BL/6 mice<br />

and reduces both liver and serum miR21 levels. Conclusions:<br />

We show that a phosphoSTAT3-miRNAs intracellular inflammatory<br />

pathway amenable to therapeutic targeting by metformin<br />

is activated in response to lipid accumulation in differentiated<br />

hepatocytes in vitro and in vivo.<br />

Disclosures:<br />

Massimo Levrero - Advisory Committees or Review Panels: BMS, Jansen, Gilead,<br />

Tekmira, Galapagos, Medimmune; Speaking and Teaching: MSD, Roche<br />

The following authors have nothing to disclose: Natalia Pediconi, Silvia Di<br />

Cocco, Silvia Piconese, Federica Mori, Laura Belloni, Abigail D. Nunn, Tullio<br />

Scopigno, Vincenzo Barnaba, Giovanni Blandino, Sabrina Strano<br />

959<br />

Exacerbation of non-alcoholic fatty liver disease<br />

(NAFLD) in mice with myeloid cell-specific deletion of<br />

FXR<br />

Rachel McMahan 1 , Cara Porsche 1 , Moshe Levi 2 , Hugo R. Rosen 1 ;<br />

1 Gastroenterology, University of Colorado, Aurora, CO; 2 Renal<br />

Diseases & Hypertension, University of Colorado, Aurora, CO<br />

Background: Nonalcoholic fatty liver disease (NAFLD) is estimated<br />

to affect a third of the U.S. population. The progression<br />

from benign steatosis to steatohepatitis in NAFLD has been<br />

linked to increased recruitment of pro-inflammatory monocytes<br />

and activation of liver macrophages. We have previously<br />

shown that activation of the nuclear bile acid receptor FXR can<br />

improve disease in murine models of NALFD and that in vitro<br />

activation of FXR in macrophages inhibits pro-inflammatory<br />

cytokine production. However, FXR is also expressed in other<br />

cell types including hepatocytes, and the protective effects of<br />

FXR may be through non-macrophage FXR signaling. Therefore,<br />

we aimed to determine the specific role of macrophage<br />

FXR expression in NALFD. Methods: To evaluate to role of<br />

macrophage expression of FXR in NAFLD, we used the Cre-<br />

Lox system to create myeloid-cell specific targeted mutants of<br />

the FXR gene (FXR fl/fl LyzMCre). FXR fl/fl LyzMCre or littermate<br />

controls were fed a western or control diet for 3 months and<br />

NAFLD severity was determined by histological analysis,<br />

hepatic gene expression and flow cytometric analysis of intrahepatic<br />

immune cells isolated by collagenase and mechanical<br />

digestion. Results: Loss of FXR expression in myeloid cells led to<br />

a more severe histological phenotype with increased macrovesicular<br />

steatosis and inflammatory foci compared to littermate<br />

controls. In addition, western diet feeding increased hepatic<br />

expression of pro-inflammatory genes (CCL2, TNF-alpha) and<br />

pro-fibrotic genes (TIMP-1) in FXR fl/fl LyzMCre mice. Myeloid<br />

cell deletion of FXR also led to increased recruitment of intrahepatic<br />

pro-inflammatory monocytes and an increase in IL-17<br />

producing T cells. Conclusions: Myeloid cell-specific deletion of<br />

FXR leads to increased liver inflammation in a murine model of<br />

NAFLD. Targeting FXR in macrophages may be important for<br />

inhibiting hepatic inflammation and NAFLD development.<br />

Disclosures:<br />

Moshe Levi - Grant/Research Support: Intercept, Genzyme-Sanofi, Daiichi Sankyo,<br />

Merck, Novartis<br />

The following authors have nothing to disclose: Rachel McMahan, Cara Porsche,<br />

Hugo R. Rosen


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 681A<br />

960<br />

miRNA-339-3p is increased in human and experimental<br />

models of NAFLD and targeted by TUDCA in insulin-resistant<br />

muscle cells<br />

André L. Simão 2 , Marta B. Afonso 2 , Pedro M. Rodrigues 2 , Andreia<br />

J. Amaral 5 , Margarida Gama-Carvalho 5 , Pedro M. Borralho 1 , Mariana<br />

V. Machado 3 , Helena Cortez-Pinto 3,4 , Cecília M. Rodrigues 1 ,<br />

Rui E. Castro 1 ; 1 iMed.ULisboa & Dep. of Biochemistry and Human<br />

Biology, Faculty of Pharmacy, University of Lisbon, Lisboa, Portugal;<br />

2 iMed.ULisboa, Faculty of Pharmacy, University of Lisbon,<br />

Lisbon, Portugal; 3 Gastrenterology, Hospital Santa Maria, Lisbon,<br />

Portugal; 4 Faculty of Medicine, University of Lisbon, Lisbon, Portugal;<br />

5 BioISI– Biosystems & Integrative Sciences Institute, Faculty of<br />

Sciences, University of Lisbon, Lisbon, Portugal<br />

Intramyocellular lipid deposition associates with insulin resistance<br />

(IR), constituting a key pathophysiological event in human<br />

non-alcoholic fatty liver disease (NAFLD). microRNAs (miRNA/<br />

miRs) play also a functional role in NAFLD both in the liver<br />

and in other metabolic tissues, including the muscle. Finally,<br />

tauroursodeoxycholic acid (TUDCA) is cytoprotective in both<br />

liver and muscle cells, in part by modulating insulin-signaling<br />

pathways. Our aims were to evaluate the functional role of<br />

muscle miR-339-3p in human and animal models of NAFLD, as<br />

well as in insulin-resistant C2C12 muscle cells, in the presence<br />

or absence of TUDCA. Skeletal muscle biopsies were obtained<br />

from morbid obese NAFLD patients undergoing bariatric surgery.<br />

Muscle and plasma RNA were extracted with Trizol (Life<br />

Technologies) and the miRCURY kit (Exiqon), respectively.<br />

Human muscle RNA was run in TaqMan MicroRNA arrays<br />

for global miRNA expression analysis. C57BL6 mice were fed<br />

either a standard or a fast food (FF) diet for 25 weeks. C2C12<br />

muscle cells were incubated with or without 200 or 400 mM<br />

palmitic acid (PA), +- 100 mM of TUDCA, for evaluation of the<br />

insulin-signaling pathway, mitochondrial function and overall<br />

cellular toxicity. Six miRNAs, including miR-339-3p, a regulator<br />

of glucose synthesis, were found to progressively and<br />

significantly increase from steatosis to NASH (at least p


682A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Christian Trautwein - Grant/Research Support: BMS, Novartis, BMS, Novartis;<br />

Speaking and Teaching: Roche, BMS, Roche, BMS<br />

Frank Tacke - Advisory Committees or Review Panels: Tobira; Grant/Research<br />

Support: Novartis, Noxxon; Speaking and Teaching: BMS, Gilead, Falk, MSD,<br />

Janssen, Abbvie<br />

The following authors have nothing to disclose: Robert Schierwagen, Lara Maybüchen,<br />

Sebastian Zimmer, Kanishka Hittatiya, Christer Baeck, Sabine Klein,<br />

Frank E. Uschner, Winfried Reul, Peter Boor, Georg Nickenig, Jogchum Plat,<br />

Dieter Luetjohann, Jonel Trebicka<br />

962<br />

TRPV4 deficiency enhances TLR4 recruitment to lipid<br />

rafts exhibiting exacerbated stellate cell activation and<br />

fibrosis in experimental nonalcoholic steatohepatitis<br />

Suvarthi Das 1 , Ratanesh K. Seth 1 , Varun Chandrashekaran 1 , Firas<br />

Alhasson 1 , Diptadip Dattaroy 1 , Gregory A. Michelotti 2 , Prakash<br />

Nagarkatti 3 , Phillip D. Bell 4 , Wolfgang B. Liedtke 2 , Anna Diehl 2 ,<br />

Saurabh Chatterjee 1 , Mitzi Nagarkatti 3 ; 1 Environmental Health<br />

Sciences, University of South Carolina, Columbia, SC; 2 Duke University,<br />

Durham, NC; 3 University of South Carolina, Columbia,<br />

NC; 4 Medical University of South Carolina, Charleston, SC<br />

Transient receptor potential vanilloid channel 4 (TRPV4) is a<br />

nonselective cation channel that confers sensitivity to extracellular<br />

osmolarity and regulates calcium inflow. We have observed<br />

previously that TRPV4 protein levels are significantly higher in<br />

NASH rodent models and human NASH patients. For the present<br />

study we aimed to investigate the role of TRPV4 in stellate<br />

cell activation and fibrosis in a methionine choline deficient<br />

(MCD) diet-induced mouse model of NASH. Liver homogenates<br />

from mice fed with MCD diet for 8 weeks (8 to 16 weeks<br />

of age), were used for qRTPCR and Western Blots. Paraffin<br />

embedded liver sections from the same mice were used for<br />

immunostaining and histopathological analysis. The Results<br />

showed that TRPV4 deficiency in mice fed with an MCD diet<br />

for 8 weeks showed significantly higher -SMA immunoreactivity<br />

in livers followed by significantly higher macro and micro<br />

vesicular fibrosis (Picrosirius Red staining) when compared to<br />

MCD wild type mice. TRPV4 deficiency led to inflammasome<br />

activation as evidenced by significantly high Apoptosis-associated<br />

speck-like protein containing a CARD (ASC II), cleaved<br />

IL-1β and cleaved Caspase-1 immunoreactivity when assessed<br />

by immunoblots as compared to wild type mice. Mice deficient<br />

in TRPV4 showed higher mRNA expressions of MCP-1 and<br />

IL-1β in the livers as compared to wildtype mice, an observation<br />

that correlated well with histopathology where increased<br />

infiltrating leukocytes were present in the liver. Since NASH<br />

progression and fibrosis were associated with increased lipid<br />

raft recruitment of TLR4 primarily mediated by NADPH oxidase,<br />

we studied whether TRPV4 deficiency could lead to<br />

enhanced NADPH oxidase driven TLR4 recruitment to lipid<br />

rafts in these mice. Results showed that there was a significant<br />

increase in NADPH oxidase activation as evidenced by colocalization<br />

of membrane gp91phox and p47phox subunits in<br />

TRPV4 knockout MCD mice as compared to wild type MCD<br />

mice. This correlated well with a significant increase in the TLR4<br />

colocalization in flotillin containing rafts in the cell membranes.<br />

These dual labeled liver sections were imaged using immunofluorescence<br />

microscopy. The TLR4 raft recruitment was also<br />

accompanied by significant elevation in TLR4 protein in the<br />

TRPV4 deficient mice. Finally, we conclude and propose that<br />

MCD diet feeding might generate severe oxidative stress which<br />

possibly leads to TRPV4 led calcium influx and might aid in<br />

preservation of membrane integrity while lack of TRPV4 causes<br />

severe loss of membrane structure causing leakage of damage<br />

associated molecular patterns into cells causing enhanced<br />

inflammation, stellate cell activation and fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Suvarthi Das, Ratanesh K. Seth,<br />

Varun Chandrashekaran, Firas Alhasson, Diptadip Dattaroy, Gregory A. Michelotti,<br />

Prakash Nagarkatti, Phillip D. Bell, Wolfgang B. Liedtke, Anna Diehl,<br />

Saurabh Chatterjee, Mitzi Nagarkatti<br />

963<br />

β 7<br />

-Integrins and MAdCAM-1 have opposing functions in<br />

development and progression of Non-alcoholic steatohepatitis<br />

(NASH)<br />

Hannah K. Drescher 1 , Angela Schippers 2 , Thomas Clahsen 2 ,<br />

Hacer Sahin 1 , Norbert Wagner 2 , Christian Trautwein 1 , Konrad L.<br />

Streetz 1 , Daniela C. Kroy 1 ; 1 Department of Internal Medicine III,<br />

University Hospital, RWTH Aachen, Germany, Aachen, Germany;<br />

2 Department of Pediatrics, University Hospital RWTH Aachen,<br />

Aachen, Germany<br />

Introduction: Non-alcoholic steatohepatitis (NASH) is the third<br />

most common reason for liver transplantations in developing<br />

countries and one of the fastest growing medical problems.<br />

Aim: The significance of infiltrating leukocytes and how they<br />

affect NASH development and progression remains unclear.<br />

Therefore, this study investigates the role of the homing receptor<br />

pair MAdCAM-1/β 7<br />

in two different mouse NASH-models.<br />

Methods: Constitutive β 7<br />

-Integrin and MAdCAM-1 knockout<br />

mice were fed either MCDdiet (methionine and choline deficient)<br />

for 4 weeks or HF-diet (high fat) for 26 weeks. Results:<br />

After 4 weeks of MCD treatment β 7<br />

-deficient animals displayed<br />

earlier and faster progressing steatosis reflected by significant<br />

histomorphological changes while MAdCAM-1-KO mice<br />

showed an intact liver architecture, both compared to wildtype<br />

controls. Consistent with severe disease progression β 7<br />

-Integrin<br />

deficient animals showed an increased oxidative stress<br />

response (DHE (Dihydroethidium)-staining) with simultaneous<br />

down regulation of the expression of Treg markers (FoxP3,<br />

IL-2, TGFβ). In contrast, MAdCAM-1-KO mice displayed an<br />

up-regulation of anti-oxidative stress proteins involved and an<br />

enhanced number of infiltrating macrophages tending to provide<br />

a strengthening of the anti-inflammatory response. The<br />

β 7<br />

-KO group exhibited a stronger hepatic infiltration of inflammatory<br />

cells reflecting an earlier onset of NASH. Especially<br />

Th17 positive T cells were increased leading to elevated numbers<br />

of infiltrating neutrophils. Those changes finally resulted<br />

in an earlier and stronger collagen accumulation in these animals<br />

while MAdCAM-1-KO mice were protected from fibrosis<br />

progression. The results suggest a direct effect of the homing<br />

receptor pair MAdCAM-1/β 7<br />

on the gut-liver-axis in NASH<br />

development by an infiltration blockade of cells which have<br />

a positive immuno-modulatory effect on the liver. In a second<br />

model, better reflecting the metabolic changes during NASH<br />

development, HF-diet was fed to mice for 26 weeks. Preliminary<br />

results show comparable results to those in the MCD<br />

model, namely a significantly increased bodyweight and a<br />

worsened glucose tolerance in β 7<br />

-KO animals indicating the<br />

occurrence of the metabolic syndrome in this group compared<br />

to MAdCAM-1-KO and WT. Conclusion: MAdCAM-1 and β 7<br />

Integrin<br />

deficiency leads to opposing effects in the MCD and the<br />

HF model, with protection in MAdCAM-1-KO animals and a<br />

more severe phenotype and significantly stronger fibrosis progression<br />

in β 7<br />

-Integrin-KO mice. Therefore, the interaction of<br />

β 7<br />

Integrins and their receptor MAdCAM-1 provide a potential<br />

novel target for therapeutic interventions during NASH development.<br />

Disclosures:<br />

Christian Trautwein - Grant/Research Support: BMS, Novartis, BMS, Novartis;<br />

Speaking and Teaching: Roche, BMS, Roche, BMS


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 683A<br />

The following authors have nothing to disclose: Hannah K. Drescher, Angela<br />

Schippers, Thomas Clahsen, Hacer Sahin, Norbert Wagner, Konrad L. Streetz,<br />

Daniela C. Kroy<br />

964<br />

Association between nonalcoholic fatty liver disease<br />

and bone mineral density: a systematic review and<br />

meta-analysis<br />

Sikarin Upala 1,3 , Veeravich Jaruvongvanich 4,5 , Karn Wijarnpreecha<br />

1,2 , Anawin Sanguankeo 1,3 ; 1 Department of Internal Medicine,<br />

Bassett Medical Center and Columbia University College of Physicians<br />

and Surgeons, Cooperstown, NY; 2 Department of Physiology,<br />

Cardiac Electrophysiology Research and Training Center,<br />

Faculty of Medicine, Chiang Mai University, Chiang Mai, Thailand;<br />

3 Department of Preventive and Social Medicine, Faculty of<br />

Medicine Siriraj Hospital, Mahidol University, Bangkok, Thailand;<br />

4 Department of Internal Medicine, University of Hawaii, Honolulu,<br />

HI; 5 Department of Internal Medicine, Faculty of Medicine, Chulalongkorn<br />

University, Bangkok, Thailand<br />

Purpose: Several major risk factors for osteoporosis have been<br />

identified. One of these risk factors is chronic inflammation.<br />

Several recent <strong>studies</strong> have supported the association between<br />

low bone mineral density (BMD) and nonalcoholic fatty liver<br />

disease (NAFLD), which comprises a spectrum of disorders<br />

involving liver inflammation. However, conflicting evidence<br />

regarding this association has been obtained thus far. We<br />

therefore conducted a meta-analysis of observational <strong>studies</strong><br />

to show the association between NAFLD and BMD. Methods:<br />

The Cochrane Central Register of Controlled Trials, Cochrane<br />

Library, Medline, and Embase were searched from database<br />

inception to November 2014 for all observational <strong>studies</strong> evaluating<br />

the association between NAFLD or nonalcoholic steatohepatitis<br />

(NASH) and bone mass, BMD, or osteoporosis. All<br />

patients were ≥18 years of age and had no other cause of liver<br />

disease, osteoporosis, or pathological bone disease at baseline.<br />

Risk factors were NAFLD and NASH; control subjects were<br />

individuals without NAFLD. Results: Eleven articles underwent<br />

full-length review. Data were extracted from 5 cross-sectional<br />

<strong>studies</strong> involving 1,276 participants; 638 had NAFLD. The<br />

main meta-analysis showed no significant difference in BMD<br />

between patients with fatty liver disease and controls. This may<br />

be explained by the confounding effects of body mass index on<br />

BMD in patients with NAFLD. Among all variables analyzed,<br />

body mass index had the strongest and most significant predictive<br />

effect on the difference in BMD. Conclusions: Controversy<br />

exists regarding the effect of BMD on NAFLD. Further <strong>studies</strong><br />

are required to fully show this relationship.<br />

965<br />

Aliskiren reduces liver and epididymal fat steatosis and<br />

increases skeletal muscle insulin sensitivity in high-fat<br />

diet-fed mice<br />

Kuei-Chuan Lee, Yun-Cheng Hsieh, Ying-Ying Yang, Che-Chang<br />

Chan, Yi-Hsiang Huang, Han-Chieh Lin; Taipei Veterans General<br />

hospital, Taipei, Taiwan<br />

Aliskiren has been found to reduce chronic injury and steatosis<br />

in the liver of methionine-choline-deficient (MCD) diet-fed<br />

mice. This study investigated whether aliskiren has an anti-steatotic<br />

effect in HFD-fed mice, which are more relevant to human<br />

patients with non-alcoholic fatty liver disease than MCD mice.<br />

Mice fed with 4-week normal chow or HFD randomly received<br />

aliskiren (50 mg/kg/day) or vehicle via osmotic minipumps for<br />

further 4 weeks. Aliskiren reduced systemic insulin resistance,<br />

hepatic steatosis, epididymal fat mass and increased gastrocnemius<br />

muscle glucose transporter type 4 levels with lower<br />

tissue angiotensin II levels in the HFD-fed mice. In addition, aliskiren<br />

lowered nuclear peroxisome proliferator-activated receptor<br />

gamma and its down-signaling molecules and increased<br />

cytochrome P450 4A14 and carnitine palmitoyltransferase 1A<br />

(CPT1a) in liver. In epididymal fat, aliskiren inhibited expressions<br />

of lipogenic genes, leading to decrease in fat mass, body<br />

weight, and serum levels of leptin and free fatty acid. Notably,<br />

in the gastrocnemius muscle, aliskiren increased phosphorylation<br />

of insulin receptor substrate 1 and Akt. Based on these<br />

beneficial effects on liver, peripheral fat and skeletal muscle,<br />

aliskiren is a promising therapeutic agent for patients with<br />

NAFLD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kuei-Chuan Lee, Yun-Cheng<br />

Hsieh, Ying-Ying Yang, Che-Chang Chan, Yi-Hsiang Huang, Han-Chieh Lin<br />

966<br />

A DPP-4 inhibitor, sitagliptin, can be preventative<br />

drug for the acceleration of the development in NAFLD<br />

caused by lipopolysaccharides through inhibiting<br />

expression of MyD88 and NFκB.<br />

Keisuke Yokohama 2,1 , Shinya Fukunishi 1 , Sanomura Makoto 2 ,<br />

Akira Asai 1 , Yasuhiro Tsuds 1 , Kazuhide Higuchi 1 ; 1 2nd Department<br />

of internal medicine, Osaka Medical College, Osaka, Japan;<br />

2 Gastroenterology, Hokusetsu general hospital, Osaka, Japan<br />

Disclosures:<br />

The following authors have nothing to disclose: Sikarin Upala, Veeravich Jaruvongvanich,<br />

Karn Wijarnpreecha, Anawin Sanguankeo<br />

Aims: Nonalcoholic fatty liver disease (NAFLD) is a chronic<br />

liver disease, and commonly observed in patients with obesity<br />

or type 2 diabetes mellitus (T2DM). Characterized by metabolic<br />

syndrome, hepatic steatosis, and liver inflammation,<br />

nonalcoholic steatohepatitis(NASH)is believed to be under<br />

the influence of the gut microflora. Some reports showed that


684A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

lipopolysaccharides (LPS) contained in normal portal blood<br />

are one of the factors about the development of NASH. Two<br />

major pathway for LPS have been suggested in some <strong>studies</strong>,<br />

which are referred to as MyD88-dependent and –independent<br />

pathways. We aimed to determine the effects of sitagliptin on<br />

the development of LPS-induced NAFLD in ob/ob mice with<br />

diet-induced obesity and T2DM. Methods: Five-week-old male<br />

ob/ob mice, which develop T2DM and NAFLD by taking a<br />

high-carbohydrate diet (HCD), were divided into a group in<br />

which a HCD was given for 8 weeks and treated with LPS<br />

(n=6) as controls (Control group), and another in which a HCD<br />

added with 0.0018% sitagliptin was given for 8 weeks and<br />

treated with LPS (n=6) (Sitagliptin group). Results: Histological<br />

examination demonstrated that this protocol induced steatosis<br />

in the livers of two group mice, however severe fatty droplets<br />

were significantly observed in Control group, compared<br />

with Sitagliptin group. Sitagliptin group showed a decrease<br />

in serum ALT and triglyceride, in addition, hepatic mRNA<br />

expression levels of SREBP-1c, FAS and peroxisome PPARγ<br />

were decreased in Sitagliptin group. Furthermore, Sitagliptin<br />

inhibits the mRNA expression levels of proinflammatory gene<br />

such as TNF-α. Interestingly, hepatic mRNA expression levels of<br />

MyD88 and NFκB were decreased in Sitagliptin group. These<br />

finding coincided with histological findings and the results of<br />

biochemical examination. Conclusion: Our findings indicated<br />

that sitagliptin could be preventative drug for the development<br />

of LPS-induced NAFLD and MyD88-dependent pathway.<br />

Disclosures:<br />

The following authors have nothing to disclose: Keisuke Yokohama, Shinya<br />

Fukunishi, Sanomura Makoto, Akira Asai, Yasuhiro Tsuds, Kazuhide Higuchi<br />

967<br />

DSS colitis has tumorinogenesis and fibrogenesis in choline-deficiency<br />

and high-fat diet induced NASH mouse<br />

model<br />

Koichi Achiwa, Masatoshi Ishigami, Yoji Ishizu, Teiji Kuzuya,<br />

Takashi Honda, Kazuhiko Hayashi, Yoshiki Hirooka, Hidemi Goto;<br />

Gastroenterology and Hepatology, Nagoya Univercity Graduate<br />

School of Medicine, Nagoya, Japan<br />

Background and Aim Non-alcoholic steatohepatitis (NASH)<br />

patients progress to liver cirrhosis and even hepatocellular carcinoma(HCC).<br />

Several line of evidence indicated that accumulation<br />

of lipopolysaccharide (LPS) plays a contributory role in<br />

the pathogenesis HCC by eliciting proinflammatory response in<br />

the liver. Moreover, it has been reported that dextran sodium<br />

sulfate (DSS) induced colitis mouse model fed high fat diet<br />

increased portal LPS and developed hepatic inflammation and<br />

fibrosis. However, diet induced NASH model with carcinogenesis<br />

require over 50 weeks or more. The purpose of this study<br />

was to create a novel tumorigenesis in diet induced NASH<br />

with DSS administration. Methods 20-week-age C57BL/6<br />

mice were fed choline-deficiency and high-fat diet (CDHF) for<br />

4 and 12 weeks. DSS group was fed CDHF and 1% DSS in<br />

the drinking water intermittently. On the other hand control<br />

group was fed CDHF and free water supply. We evaluated the<br />

liver tumorigenesis. Hepatic lipid was measured and inflammatory<br />

and fibrosis factors and histological changes were evaluated.<br />

Results Exposure of DSS led to increase mucosal change<br />

such as crypt loss and increase of inflammatory cells in colon.<br />

Accordingly, the histological score significantly increased in<br />

DSS group than control group. But a few colon tissues were<br />

increased inflammatory cells at 12 week in control group.<br />

CDHF led to slightly cause intestinal inflammation. Portal LPS<br />

levels were increased in DSS group comparing with control<br />

group at 4 week, whereas there was no significant difference<br />

in both groups at 12week. Hepatic triglyceride, total and free<br />

cholesterol were increased in the DSS group comparing with<br />

control group at 4 week, whereas there was no significant<br />

difference in both groups at 12 week. Both groups showed<br />

severe macrovesicular steatosis in liver histology. Serum total<br />

bilirubin levels were significantly higher in DSS group than that<br />

in control group. Serum transaminase levels are higher in DSS<br />

group than that in control group. A large number of lobular<br />

inflammatory cells were seen in DSS group. The image analysis<br />

showed significant elevation of F4/80 staining positive area in<br />

DSS group comparing with control group. Fibrosis is increased<br />

in DSS group compared to control group. Liver tumors were<br />

developed in the DSS group of 12-week, while tumor was not<br />

found in the DSS group of 4-week and in the control group.<br />

One of tumors showed highly differentiated hepatocellular carcinoma.<br />

Conclusions DSS administration developed liver tumors<br />

in CDHF induced NASH mouse at the short term. Induction of<br />

intestinal inflammation in NASH may promote hepatic tumorigenesis.<br />

Disclosures:<br />

Hidemi Goto - Grant/Research Support: MSD, Roche, Bayer, Bristol-Myers, Eisai,<br />

Ajinomoto, Otsuka, Astra, Tanabe, Takeda<br />

The following authors have nothing to disclose: Koichi Achiwa, Masatoshi Ishigami,<br />

Yoji Ishizu, Teiji Kuzuya, Takashi Honda, Kazuhiko Hayashi, Yoshiki<br />

Hirooka<br />

968<br />

Inhibition of Mitophagy in Non-alcoholic Steatohepatitis<br />

lead to Apoptosis and Inflammation<br />

Muhammad Sohail 1 , Seong-Jun Kim 2 , Aleem Siddiqui 2 ; 1 GASTRO-<br />

ENTEROLOGY, university of California, San Diego, CA; 2 Infectious<br />

Disease, University of California, San Diego, CA<br />

Background: Fatty liver is benign but it can progress to Non-Alcoholic<br />

Steatohepatitis (NASH) and liver cirrhosis. Mitochondria<br />

are dynamic organelles that constantly undergo fission,<br />

fusion, and mitophagy to facilitate mitochondrial quality control,<br />

which is crucial for maintaining cell viability and bioenergetics.<br />

Mitochondrial fission/fragmentation is mediated<br />

by recruitment of cytosolic Drp1 to the mitochondria forming<br />

spirals that constrict both the inner and outer mitochondrial<br />

membranes. Drp1 recruitment to mitochondria is regulated by<br />

phosphorylation and dephosphorylation of respective serine<br />

residues by putative kinases and phosphatases. Mitochondrial<br />

dynamics is tightly regulated in response to alterations<br />

in cellular physiology and its dysfunction occurs in NASH.<br />

Methods Palmitic acid (PA) is able to reproduce in vitro model<br />

for fatty liver disease. To investigate the mitochondrial fission<br />

and mitophagy, the human hepatoma Huh7 cells cultured in<br />

the presence or absence of PA were observed by confocal<br />

microscope. PA-induced stimulation and mitochondrial translocation<br />

of dynamin-related protein 1 (Drp1) and Parkin, two<br />

key factors of mitochondrial fission and mitophagy, were analyzed<br />

by Western blot using purified mitochondria fraction<br />

and confocal microscope. The liver tissues extracted from highfat<br />

diet-fed C57BL/6 mice and biopsy samples from Human<br />

NASH patients were also used for analysis of Drp1 and Parkin<br />

activation. Small interfering RNA-mediated Drp1 and Parkin<br />

silencing strategy was further employed for further analysis of<br />

PA-induced mitochondrial fission and mitophagy and alteration<br />

in inflammatory response. Results: PA stimulated Drp1 gene<br />

expression and induced mitochondrial translocation of Drp1<br />

via promoting both its phosphorylation (Ser616) and dephosphorylation<br />

(Ser637) followed by mitochondrial fission. These<br />

results were confirmed in liver tissue extracted from high-fat<br />

diet-fed C57BL/6 mice and also in human NASH liver biopsy<br />

tissue. While PA induced Parkin gene expression, mitophagy


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 685A<br />

was not observed. PA also induce PINK1 cleavage, implying<br />

the possible inflammatory and apoptosis signaling. Using specific<br />

protease inhibitor that cleaves PINK1 in inner mitochondrial<br />

membrane, we were able to block its cleavage which<br />

also leads to the inhibition of apoptosis. As mitophagy does not<br />

occur which is important mitochondrial quality control, this lead<br />

to surplus damaged mitochondria and subsequent apoptosis<br />

of cell further triggering inflammation. Conclusion: Fatty acids<br />

induce aberrant mitochondrial dynamics involving mitochondrial<br />

fission, caused apoptosis but does not lead to mitophagy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Muhammad Sohail, Seong-Jun<br />

Kim, Aleem Siddiqui<br />

Figure: NFκB inhibition by AP inhibits IL-1β expression induced by<br />

LPS. (A) AP treatment inhibited NFκB activation by LPS in HepG2<br />

cells, evaluated through a reporter gene assay. (B) NFκB inhibition<br />

by AP treatment inhibited IL1β induction by LPS in Palmitic<br />

acid-treated HepG2 cells. (**p≤0.01, ***p ≤ 0.001).<br />

969<br />

NFkB inhibition by Andrographolide ameliorates<br />

inflammation and fibrogenesis through inflammasome<br />

substrate depletion in experimental Non-Alcoholic Steatohepatitis<br />

(NASH)<br />

Daniel Cabrera 1,2 , Alexander Wree 3 , Davide Povero 3 , Nancy<br />

Solis 1 , Margarita Pizarro 1 , Pamela Rojas de Santiago 1 , Javiera<br />

Torres 7 , Han Moshage 5 , Claudio Cabello-Verrugio 4 , Ariel E.<br />

Feldstein 3 , Enrique Brandan 6 , Marco Arrese 1 ; 1 Departamento de<br />

Gastroenterología, Facultad de Medicina Pontificia Universidad<br />

Católica de Chile, Santiago, Chile; 2 Departamento de Ciencias<br />

Químico-Biológicas, Universidad Bernardo O Higgins, Santiago,<br />

Chile; 3 Department of Pediatrics, University of California San<br />

Diego, San Diego, CA; 4 Departamento de Ciencias Biológicas,<br />

Universidad Andres Bello, Santiago, Chile; 5 Department of Gastroenterology<br />

and Hepatology, University of Groningen, Santiago,<br />

Chile; 6 Departamenti de Biologia Celular y Molecular, Pontificia<br />

Universidad Católica de Chile, Santiago, Chile; 7 Departamento<br />

de Anatomia Patologica, Pontificia Universidad Católica de Chile,<br />

Santiago, Chile<br />

Background: Options to treat NASH are limited. Andrographolide<br />

(AP), the bioactive component of Andrographis paniculata,<br />

has potent anti-inflammatory activity related to NFκB<br />

inhibition. NFkB is a crucial factor in Interleukin-1β (IL-1β)<br />

expression, the main substrate for the inflammasome. Aim: To<br />

evaluate the effects of AP in experimental NASH and its influence<br />

in inflammasome activity. Methods: C57bl6 mice were<br />

fed a choline-deficient-amino-acid–defined (CDAA) diet (22<br />

weeks) with or without AP (1 mg/Kg, 3 times/week, intraperitoneally).<br />

Serum levels of alanine aminotransferase (ALT) and<br />

hepatic steatosis, inflammation, and fibrosis were assessed.<br />

Hepatic triglyceride content (HTC) and hepatic mRNA levels<br />

of selected pro-inflammatory and pro-fibrotic genes as well as<br />

those related to inflammasome were also measured. Direct AP<br />

effect on IL-1β expression and inflammasome activity was evaluated<br />

in HepG2 cells loaded with a mixture of FFAs during 24h.<br />

Results: AP decreased HTC and attenuated hepatic inflammation<br />

and fibrosis in CDAA-fed mice reducing serum ALT activity<br />

and hepatic collagen deposition. AP treatment was associated<br />

with a strong reduction in hepatic macrophage infiltration<br />

and of hepatic mRNA levels of both pro-inflammatory and<br />

pro-fibrotic genes. In addition, mice treated with AP showed<br />

reduced expression of inflammasome genes (-60% reduction in<br />

inflammasome adaptor ASC hepatic mRNA levels, and -40%<br />

reduction in NLRP3 hepatic mRNA levels). Finally, AP inhibited<br />

IL-1β expression induced by LPS through NFκB inhibition in fat<br />

laden HepG2 cells (figure). Conclusion: AP administration prevented<br />

liver damage in NASH. Inflammasome inactivation by<br />

NFκB-dependent mechanism involving IL-1β expression inhibition<br />

is likely involved in the therapeutic effects of AP. (Fondecyt<br />

1150327 to M.A. and FONDECYT PD3140396 to D.C.)<br />

Disclosures:<br />

The following authors have nothing to disclose: Daniel Cabrera, Alexander<br />

Wree, Davide Povero, Nancy Solis, Margarita Pizarro, Pamela Rojas de Santiago,<br />

Javiera Torres, Han Moshage, Claudio Cabello-Verrugio, Ariel E. Feldstein,<br />

Enrique Brandan, Marco Arrese<br />

970<br />

Reduction of Hepatic 27-Hydroxycholesterol in Steatohepatitis<br />

Model Mice with Insulin Resistance<br />

Sho-ichiro Yara 1 , Tadashi Ikegami 1 , Akira Honda 1,2 , Teruo<br />

Miyazaki 2 , Masashi Murakami 1 , Junichi Iwamoto 1 , Yasushi Matsuzaki<br />

1 ; 1 Division of Gastroenterology and Hepatology, Tokyo<br />

Medical University Ibaraki Medical Center, Ami, Inashiki, Japan;<br />

2 Collaborative Research Center, Tokyo Medical University Ibaraki<br />

Medical Center, Ami, Inashiki, Japan<br />

BACKGROUND: 27-hydroxycholesterol (HC), one of the major<br />

oxysterol found in the human circulation, is produced by the<br />

mitochondrial enzyme CYP27A1. Previous report demonstrated<br />

that hepatic inflammation was increased by genetic deletion of<br />

CYP27A1 and decreased by 27HC treatment in high fat diet<br />

(HFD) fed steatotic mice, however, direct cause of dysregulation<br />

of CYP27A1 and 27HC in steatohepatitis is unknown.<br />

AIM: The current study was undertaken to examine the change<br />

of hepatic OS levels in the Stelic Animal Model (STAM) mice,<br />

a validated and widely used as NASH-derived HCC model.<br />

METHODS: Extracted sterols in the liver tissue homogenates<br />

were determined by LC-MS/MS. RESULTS: In STAM mice, IR<br />

was induced by injection of streptozotocin at the birth period.<br />

Four weeks after injection, feeding of high-fat diet was started,<br />

and given over time. At week 20, STAM mice developed hepatocellular<br />

carcinoma (HCC) whereas HCCs were not found in<br />

control (CTL) as well as mice only fed with HFD. (1) Concentration<br />

of hepatic cholesterol (CHOL): Hepatic CHOL concentrations<br />

were kept in the narrow range (2.5 to 2.95 mg/mg wet<br />

weight liver (ww)) by 12 weeks old either in CTL, HFD fed, or<br />

STAM mice, but significantly elevated at 20 weeks in HFD fed<br />

mice (4.7 mg/mg ww) and STAM mice (4.2 mg/mg ww) compared<br />

to CTL (2.2 mg/mg ww). (2) Concentration of hepatic<br />

bile acids (BAs): In HFD fed mice, hepatic BA concentrations<br />

(111.1 to 120.8 nmol/g ww) were significantly lower than CTL<br />

(177.4 to 205.6 nmol/g ww) over time, suggesting enhanced<br />

BA excretion from liver under the condition with higher hepatic<br />

CHOL level in the absence with IR. In STAM mice, hepatic BA<br />

concentrations were maintained in a similar level of HFD at<br />

earlier phase (4 to 8 wk old), but those were significantly elevated<br />

at later phase (12 to 20wk old)(215.0 nmol/g ww at 12<br />

wk, 289.0 nmol/g ww at 20wk). (3) Concentration of hepatic<br />

OS: Among major OS tested (24S, 25epo-HC, 22RHC, 24HC,<br />

25HC, 27HC, 4βHC), only 27HC was significantly lower in<br />

STAM mice (0.388 to 0.579 ng/mg ww) compared to either


686A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of CTL (0.606 to 0.942 ng/mg ww) or HFD (0.628 to 0.755<br />

ng/mg ww) over time. CONCLUSION: Reduced hepatic 27HC<br />

concentration observed in this model, is supportive the hypothesis<br />

proposed by previous <strong>studies</strong>. Enhancement of the reduction<br />

of hepatic 27HC concentration in the presence of IR suggested<br />

a linkage between IR and dysregulation of CYP27A1.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sho-ichiro Yara, Tadashi Ikegami,<br />

Akira Honda, Teruo Miyazaki, Masashi Murakami, Junichi Iwamoto, Yasushi<br />

Matsuzaki<br />

971<br />

Role of Fcgamma receptors in experimental Non-Alcoholic<br />

Steatohepatitis<br />

Daniel Cabrera 1,3 , Evelyn Jara 2 , Ricardo Cruz 1 , Natalia<br />

Muñoz-Durango 2 , Pamela Rojas de Santiago 1 , Alexis Kalergis 2 ,<br />

Marco Arrese 1 ; 1 Departamento de Gastroenterología, Facultad<br />

de Medicina, Pontificia Universidad Católica de Chile, Santiago,<br />

Chile; 2 Departamento de Genética Molecular y Microbiología,<br />

Facultad de Ciencias Biológicas, Pontificia Universidad Católica<br />

de Chile, Santiago, Chile; 3 Departamento de Ciencias Químico-Biológicas,<br />

Universidad Bernardo O Higgins, Santiago, Chile<br />

Background: The role of innate immunity in Non-alcoholic steatohepatitis<br />

(NASH) development is emergent. Different populations<br />

of innate immune cells are present in the liver controlling<br />

tissue cytokine production through Fc gamma receptors (FcγRs),<br />

which are receptors for the Fc region of IgG antibodies. FcγRs<br />

cell-type specifically interact with various other receptors for<br />

selective amplification or inhibition of particular cytokines.<br />

Aim: To evaluate the role of the inhibitory (FcγRIIB) and activatory<br />

(FcγRIII) receptors in NASH pathogenesis. Methods:<br />

Wild Type, FcγRIIb(-/-) and FcγRIII(-/-) mice were fed a methionine-choline<br />

deficient (MCD) diet for 5 weeks. Liver injury was<br />

assessed by measuring serum levels of alanine aminotransferase<br />

(ALT) and histologically. Hepatic triglyceride content (HTC)<br />

and hepatic mRNA levels of selected pro-inflammatory (TNF-α,<br />

IFN-γ, IL-1β, etc.), pro-fibrotic (TGF-β, CTGF, Collagen-1, etc.)<br />

and inflammasome (ASC, NLRP3, Caspase-1, etc.) genes were<br />

also assessed. Serum pro-inflammatory cytokine levels (TNF-α,<br />

IFN-γ, etc.) were determined by cromatographic bead assay<br />

(CBA). By flow cytometry was evaluated different populations<br />

of inflammatory cells infiltrated in the liver such as dendritic,<br />

neutrophils, macrophages and lymphocytes. Results: FcγRIIb(-/-)<br />

MCD-fed mice developed a more robust hepatic inflammatory<br />

(decreased hepatic expression of TNFα and other cytokines)<br />

and fibrotic response (decreased hepatic expression of collagen<br />

I) in comparison with WT MCD-fed mice with no differences<br />

in HTC (Figure 1 A and B). No differences were found<br />

in liver cell populations of lymphoid and myeloid lineages.<br />

The main finding in FcγRIII(-/-) MCD-fed mice was a significant<br />

reduction in histological steatosis and HTC. (see figure 1C,<br />

below) likely related to reduced interleukin-10 production. Liver<br />

lymphoid and myeloid cell populations remained unchanged in<br />

these mice. Conclusion: Our results suggest and important role<br />

of the FcγRs in NASH development. While the absence of Fcγ−<br />

RIIb seems to promote NASH induction, the absence of FcγRIII<br />

strongly reduces liver steatosis. This is the first report that shows<br />

a direct role of FcγRs in the pathogenesis of NASH.(Grant<br />

support: FONDECYT 1150327 to M.A., PD3140396 to D.A.)<br />

Disclosures:<br />

The following authors have nothing to disclose: Daniel Cabrera, Evelyn Jara,<br />

Ricardo Cruz, Natalia Muñoz-Durango, Pamela Rojas de Santiago, Alexis Kalergis,<br />

Marco Arrese<br />

972<br />

Regulation of Hepatocellular Fatty Acid Uptake in<br />

Mouse Models of Fatty Liver Disease With and Without<br />

Functional Leptin Signalling: Roles of NfKB and<br />

Srebp-1c<br />

Fengxia Ge, Jose L. Walewski, Paul D. Berk; Medicine, Columbia<br />

University Medical Center, New York, NY<br />

The specific processes leading to increased hepatic triglycerides<br />

(TGs) in mouse models of hepatic steatosis (HS) due<br />

to EtOH consumption, high fat diets (HFDs), or obesity mutations<br />

remain uncertain. This study focuses on regulation by 5<br />

transcription factors (Nfb, Srebp-1c, AMPK, PPARα, PPARγ) of<br />

the 7 most-studied hepatic LCFA transporters (FABPpm, CD36,<br />

FATP1, FATP2, FATP4, FATP5, & Caveolin-1 [CAV-1]) and<br />

enzymes of LCFA synthesis (SCD-1, FASN) in mice with HS<br />

from various causes. Methods: Seven groups of 20 wk old<br />

male mice (n=8/group; Jackson Labs) were studied: C57BL/6J<br />

controls (C); similar mice with HS from 12 wks of a high fat diet<br />

(HFD) or 10, 14, or 18% EtOH in drinking water (E10, E14,<br />

E18); & genetically obese mice lacking leptin (ob/ob) or its<br />

receptor (db/db). Hepatocyte [ 3 H]-oleate uptake was assayed<br />

by rapid filtration; Vmax for saturable uptake was computed<br />

with SAAM II. Gene expression for the 5 transcription factors,<br />

7 transporters, & for key enzymes of LCFA synthesis was<br />

assayed by qRT-PCR. Hepatic TG was measured biochemically.<br />

Expression ratios for transcription factors and transporter<br />

genes were compared with one another, with Vmax, & with<br />

hepatic TG. Results: [1] LCFA uptake Vmax was increased &<br />

highly correlated with hepatic TG in all groups except ob/ob &<br />

db/db. [2] Increased Vmax & hepatic TG in the EtOH & HFD<br />

groups correlated best with increased expression of genes for<br />

at least one & often multiple transporters. [3] Despite variable<br />

expression of single transporter genes in individual E & HFD<br />

groups, the mean expression ratio for FABPpm, FATPs 1,2,4,&<br />

5, & CD36 in each group was highly correlated with Vmax,<br />

hepatic TG, and expression of transcription factor genes. [4]<br />

Of the transcription factors, SREBP-1c (r = 0.99) and NfKB (r<br />

= 0.94) were by far the most closely correlated with Vmax. [5]<br />

Increased hepatic TGs in ob/ob & db/db mice did not relate<br />

to hepatic LCFA uptake, but instead were highly correlated<br />

with increased expression of LCFA synthetic enzymes (SCD-1,<br />

FASN). Conclusions: [1] The processes underlying increased<br />

hepatic TG are different in mice with vs without functional leptin<br />

signaling. Increased LCFA uptake is the principal cause of HS<br />

in the former, but increased LCFA synthesis predominates in<br />

the latter. [2] Regulation of LCFA transporter expression & participation<br />

of individual transporters in LCFA uptake are more<br />

complex than previously believed. [3] Correlations between<br />

transcription factor expression & mean expression of multiple<br />

transporter genes suggests possible regulatory interaction, and<br />

may support reports postulating that a complex of FABPpm,<br />

CAV-1, CD36 & FATP4 mediates hepatic LCFA uptake.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 687A<br />

Disclosures:<br />

The following authors have nothing to disclose: Fengxia Ge, Jose L. Walewski,<br />

Paul D. Berk<br />

973<br />

Role of the receptor tyrosine kinase Mer in the development<br />

of fibrosis in NAFLD<br />

Giovanni Di Maira 1 , Salvatore Petta 2 , Andrea Cappon 1 , Elisa<br />

Vivoli 1 , Luca Valenti 3 , Paola Dongiovanni 3 , Elisabetta Bugianesi 4 ,<br />

Vito Di Marco 2 , Fabio Marra 1 ; 1 University of Florence, Florence,<br />

Italy; 2 University of Palermo, Palermo, Italy; 3 University of Milan,<br />

Milan, Italy; 4 University of Turin, Turin, Italy<br />

Background/aims: Non-alcoholic fatty liver disease (NAFLD)<br />

describes a spectrum of conditions ranging from steatosis to<br />

non-alcoholic steatohepatitis (NASH), which may progress<br />

to fibrosis, cirrhosis and hepatocellular carcinoma. Several<br />

genetic factors involved in NAFLD severity and progression<br />

have been identified through genome-wide association <strong>studies</strong><br />

(GWAS) but their biologic significance and sites of action is<br />

still unclear. Mer (MerTK) is a receptor tyrosine kinase with<br />

oncogenic properties, over-expressed or activated in various<br />

malignancies. In a recent GWAS, a SNP of the non-coding<br />

region of MerTK (rs4374383 G>A) has been associated with<br />

fibrosis severity in patients with chronic hepatitis C, but currently<br />

there is no evidence of the involvement of this factor<br />

in the progression of NAFLD. Aim of this study was to assess<br />

the possible role of MerTK in the fibrogenic progression of<br />

NAFLD. Methods: Genetic analyses were performed on specimens<br />

from patients who underwent liver biopsy for suspected<br />

NASH without severe obesity. For murine models of fibrogenesis,<br />

C57BL6/J mice were treated with CCl4 for 6 weeks, and<br />

Balb/C mice were fed with a methionine and choline deficient<br />

(MCD) diet for 8 weeks. Biopsy samples from patients with<br />

NAFLD and different stages of fibrosis were also analyzed.<br />

Intrahepatic gene expression was measured by qPCR. Human<br />

HSC were isolated from normal liver tissue and cultured on<br />

plastic. UNC569 was used to inhibit MerTK activity. Results:<br />

The MerTK rs4374383 AA polymorphism, which is linked with<br />

a decreased expression of MerTK, was associated with a lower<br />

prevalence of clinically significant fibrosis in patients with<br />

NAFLD. In murine models of hepatic fibrogenesis, an increased<br />

hepatic expression of MerTK was observed. In human hepatic<br />

stellate cells (HSC) Mertk was found to be highly expressed<br />

at the gene and protein levels. Stimulation of MerTK with the<br />

cognate ligand, GAS6, resulted in time-dependent activation of<br />

ERK1/2 and in stimulation of cell migration. Moreover, pharmacological<br />

inhibition of MerTK with UNC569 reduced HSC<br />

survival activating apoptotic pathways. Finally, patients with<br />

NAFLD and severe fibrosis had significantly higher intrahepatic<br />

expression of MerTK than those with mild or no fibrosis.<br />

Conclusion: These data indicate the novel role of MerTK as a<br />

modulator of fibrogenesis in NAFLD and identify HSC as a<br />

cellular target of this kinase.<br />

Disclosures:<br />

Fabio Marra - Advisory Committees or Review Panels: Abbvie; Consulting: Bayer<br />

Healthcare; Grant/Research Support: ViiV<br />

The following authors have nothing to disclose: Giovanni Di Maira, Salvatore<br />

Petta, Andrea Cappon, Elisa Vivoli, Luca Valenti, Paola Dongiovanni, Elisabetta<br />

Bugianesi, Vito Di Marco<br />

974<br />

GFT505 (ELAFIBRANOR) prevents nonalcoholic steatohepatitis<br />

(NASH), hepatic fibrosis and hepatocarcinoma<br />

in a new disease model<br />

Benoit Noel 1 , Geraldine Rigou 1 , Nathalie Degallaix 1 , Valérie<br />

Daix 1 , Linda Cambula 1 , Alice Roudot 1 , Rémy Hanf 1 , Dean W.<br />

Hum 1 , Bart Staels 2 , Robert Walczak 1 ; 1 Genfit SA, Loos, France;<br />

2 INSERM UMR1011, Institut Pasteur, Lille, France<br />

Background: NAFLD is present in 20-30% of the adult population<br />

in developed countries. NASH, cirrhosis and hepatocellular<br />

carcinoma will develop in a substantial proportion of<br />

people with NAFLD. Recently, the phase 2B GOLDEN-505 trial<br />

has demonstrated the efficacy of GFT505 in reversing NASH<br />

in patients with advanced disease. In the present study we<br />

have assessed the efficacy of GFT505 in preventing NASH,<br />

liver fibrosis and hepatocarcinoma development in a new rat<br />

model. Methods: Steatohepatitis, NASH and hepatocarcinoma<br />

were induced by feeding Wistar rats with a choline-deficient<br />

L-amino-acid-defined-diet (CDAA) that was supplemented with<br />

1% cholesterol. In the intervention group, animals also received<br />

GFT505 (10mg/kg/day PO) for the entire study period.<br />

NASH, fibrosis and hepatocarcinoma development were evaluated<br />

by histology. Additional biochemical and molecular<br />

analyses were also performed on different relevant biomarkers.<br />

Results: CDAA diet administration in Wistar rats causes<br />

liver transaminase elevation and fibrosing steatohepatitis that is<br />

similar to human NASH pathology. In this study the supplementation<br />

of the CDAA diet with 1% cholesterol led to increased<br />

NASH incidence (92% vs 58%) and to fibrosis aggravation as<br />

evidenced by higher prevalence of cirrhotic animals (42% vs.<br />

8%). NASH development was significantly attenuated in rats on<br />

CDAA diet that were also administered GFT505. Consistently,<br />

hepatic inflammation, hepatocyte ballooning, oxidative stress<br />

and plasma transaminase levels were decreased, whereas<br />

plasma levels of FGF-21 and β-hydroxybutyrate, as well as<br />

the expression of fatty acid oxidation genes in the liver were<br />

increased in rats that received GFT505. Hepatic collagen was<br />

lower, fibrosis and cirrhosis development were all significantly<br />

prevented by GFT505 administration on CDAA diet. Finally,<br />

GFT505 administration completely prevented pre-neoplastic<br />

lesion development in rats exposed to CDAA diets as revealed<br />

by the GST-P staining. Conclusion: GFT505 prevented NASH,<br />

fibrosis, cirrhosis and pre-neoplastic lesion development in this<br />

novel NASH model. Since the effect of GFT505 on hepatic<br />

steatosis was rather modest in this particular model, this study<br />

demonstrates that the antifibrotic activity of GFT505 can be<br />

dissociated from its hypolipidemic efficacy.<br />

Disclosures:<br />

Benoit Noel - Employment: Genfit SA<br />

Nathalie Degallaix - Employment: GENFIT SA<br />

Linda Cambula - Employment: GENFIT SA<br />

Alice Roudot - Employment: GENFIT<br />

Rémy Hanf - Management Position: GENFIT<br />

Dean W. Hum - Management Position: Genfit<br />

Bart Staels - Advisory Committees or Review Panels: MSD; Consulting: Genfit<br />

Robert Walczak - Management Position: Genfit SA<br />

The following authors have nothing to disclose: Geraldine Rigou, Valérie Daix


688A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

975<br />

Free fatty acid treatment reduces BMP6 signalling in<br />

AML12 hepatocytes<br />

Nishreen Santrampurwala 1,2 , Kim Bridle 1,2 , Janske Reiling 1,3 ,<br />

Laurence J. Britton 1,2 , Lesley-Anne Jaskowski 1,2 , Nathan Subramaniam<br />

1,4 , Darrell H. Crawford 1,2 ; 1 School of Medicine, The University<br />

of Queensland, Greenslopes, QLD, Australia; 2 Gallipoli<br />

Medical Research Foundation, Brisbane, QLD, Australia; 3 NUTRIM<br />

School of Nutrition and Translational Research in Metabolism,<br />

Maastricht University, Maastricht, Netherlands; 4 QIMR Berghofer<br />

Medical Research Institute, Brisbane, QLD, Australia<br />

Introduction: Individuals with non-alcoholic steatohepatitis<br />

(NASH) frequently have increased hepatic iron stores, which<br />

appear to potentiate liver injury. The underlying mechanism of<br />

iron loading in these patients however has remained elusive.<br />

Our group has previously shown a greater severity of injury in<br />

Hfe -/- mice fed a high calorie diet (HCD) compared to wild type<br />

controls. Despite iron loading, mice fed a HCD had markedly<br />

reduced expression of hepcidin (Hamp1), the master regulator<br />

of systemic iron homeostasis. BMP (bone morphogenetic<br />

protein)-SMAD signalling is known to be important in hepcidin<br />

induction; we hypothesised that BMP signalling is altered<br />

in a fatty acid-rich environment. We developed an in vitro<br />

model of NAFLD to examine the consequences of accumulated<br />

fatty acids on the BMP-SMAD signalling pathway. Methods:<br />

The mouse hepatocyte cell line (AML12) was serum-starved<br />

for 4 hours before treatment with free fatty acids (FFA; 2mM<br />

of oleate and palmitate in a 2 to 1 ratio), ferric ammonium<br />

citrate (Fe, 100μM) and/or BMP6 (50ng/ml) for 12-hours.<br />

RNA and protein were extracted for further downstream analysis.<br />

Results: Serum-starved AML12 cells treated with BMP6<br />

had increased expression of Hamp1, Smad7, Id1 and Atoh8<br />

indicating an intact BMP6 signalling pathway. Cells cultured<br />

with FFA however had a blunted response to BMP6 treatment.<br />

This was evident by the significant decrease in expression of<br />

Hamp1, Smad7, Id1 and Atoh8 (p ≤ 0.05) in cells treated with<br />

FFA despite BMP6 treatment. Additionally, we investigated the<br />

levels of phospho-SMAD1/5/8, a BMP6 signalling intermediary,<br />

and found that while phosphorylation of SMAD1/5/8<br />

was stimulated with BMP6 treatment the induction was significantly<br />

blunted with FFA treatment (p ≤ 0.001). Discussion: Iron<br />

loading in NAFLD has the potential to increase oxidative stress<br />

and lipid peroxidation, thus contributing to progressive liver<br />

disease. The mechanisms underlying iron loading in NAFLD<br />

are largely unknown and reduced BMP-SMAD signalling in<br />

response to FFA treatment represents a plausible mechanistic<br />

link, by decreasing hepcidin expression, thus facilitating duodenal<br />

iron absorption via ferroportin.<br />

Disclosures:<br />

Darrell H. Crawford - Advisory Committees or Review Panels: Roche Products<br />

Pty Ltd, Bristol Myers Squibb, Gilead Sciences, Novartis, MSD, Abbvie, Jansen;<br />

Consulting: Roche Products Pty Ltd; Grant/Research Support: Roche Products<br />

Pty Ltd; Speaking and Teaching: Roche Products Pty Ltd, Bristol Myers Squibb,<br />

Gilead Sciences, MSD<br />

The following authors have nothing to disclose: Nishreen Santrampurwala, Kim<br />

Bridle, Janske Reiling, Laurence J. Britton, Lesley-Anne Jaskowski, Nathan Subramaniam<br />

976<br />

Increased sensitivity and compromised time-dependent<br />

adaptation to methoxamine in a rat model of severe<br />

steatosis<br />

Wilhelmus J. Kwanten 1 , Paul Fransen 2 , Joris G. De Man 1 , Benedicte<br />

Y. De Winter 1 , Peter P. Michielsen 1,3 , Sven M. Francque 1,3 ,<br />

Denise Van Der Graaff 1 , Michiel Landen 1 ; 1 Laboratory of Experimental<br />

Medicine and Pediatrics (LEMP) - Gastroenterology &<br />

Hepatology, University of Antwerp, Wilrijk, Belgium; 2 Laboratory<br />

of Physiopharmacology, University of Antwerp, Wilrijk, Belgium;<br />

3 Gastroenterology Hepatology, Antwerp University Hospital<br />

(UZA), Edegem, Belgium<br />

NAFLD causes important intrahepatic vascular alterations and<br />

an increased portal blood pressure, even before the development<br />

of inflammation or fibrosis. Intrahepatic endothelial<br />

dysfunction and a decreased sensitivity to vasoconstriction to<br />

the α1-adrenoceptor agonist methoxamine (Mx) were reported.<br />

AIM To study the underlying mechanisms of the reported hyporesponsiveness<br />

to Mx and potential time-dependent effects.<br />

METHODS Male Wistar rats were fed a methione-choline deficient<br />

diet (MCDD) to induce steatosis or control diet (CD) diet<br />

for 4 weeks (n=4-5/group). Intrahepatic vascular resistance<br />

was studied by in situ isolated liver perfusion (flow: 10-50<br />

ml/min). Dose-response curves to Mx were constructed evaluating<br />

intrahepatic vascular tone. Finally, intrahepatic vascular<br />

tone was assessed as a function of time at a fixed dose<br />

(at Emax, 10-4 M Mx) and flow (30ml/min). RESULTS Basal<br />

intrahepatic resistance didn’t differ between steatotic and<br />

control rats, nor was the response to increasing flow rates.<br />

Dose-response curves to Mx showed a significantly increased<br />

sensitivity to Mx in steatosis compared to controls (-LogEC50:<br />

5.55 vs 5.20; p60min: 19.24>11.25mmHg). Steatotic<br />

livers showed a similar initial contraction, while time-dependent<br />

decrease was delayed and attenuated (5>60min:<br />

19.03>16.88mmHg)(fig 1). Significance was achieved after<br />

20min Mx. When expressed as percentage decrease compared<br />

to the level after 5min Mx, significance was already<br />

achieved after 10min. CONCLUSION Our results show a significant<br />

and time-dependent adaptation of intrahepatic vascular<br />

tone to Mx in control rats. This phenomenon should be taken<br />

into account when interpreting experiments with Mx pre-constriction.<br />

Secondly, although we didn’t observe an increased<br />

intrahepatic vascular resistance, response to Mx was significantly<br />

altered in steatosis, with a significant hyperresponsiviness<br />

and impaired time-dependent adaptation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Wilhelmus J. Kwanten, Paul<br />

Fransen, Joris G. De Man, Benedicte Y. De Winter, Peter P. Michielsen, Sven M.<br />

Francque, Denise Van Der Graaff, Michiel Landen


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 689A<br />

977<br />

NorUDCA reduces liver injury and improves glucose<br />

sensitivity in a mouse model of obesity and steatosis.<br />

Daniel Steinacher 1 , Thierry Claudel 1 , Tatjana Stojakovic 2 , Michael<br />

Trauner 1 ; 1 Gastroenterology and Hepatology, Medical University<br />

of Vienna, Vienna, Austria; 2 Clinical Institute of Medical and<br />

Chemical Laboratory Diagnostics, Medical University of Graz,<br />

Graz, Austria<br />

Background: Non-alcoholic fatty liver disease ranges from simple<br />

benign steatosis to the more severe forms of non-alcoholic<br />

steatohepatitis (NASH), cirrhosis and ultimately hepatocellular<br />

carcinoma. Currently there is no satisfactory drug therapy<br />

against NASH available. The leptin deficient mouse (ob/ob)<br />

mimics metabolic syndrome and suffers from hyperinsulinemia,<br />

insulin resistance, hyperlipidemia, hepatic steatosis and inflammation.<br />

Bile acids (BA) are not only fat-absorption facilitators<br />

but can also regulate metabolic processes directly. Currently,<br />

ursodeoxycholic acid (UDCA) is the only BA used as a therapeutic<br />

in humans, but has limited efficacy in NASH. NorUDCA<br />

is a side-chained shortened derivative of UDCA improving liver<br />

injury in mouse models of cholestatic liver and bile duct injury.<br />

Therefore, we aimed to explore whether NorUDCA improves<br />

liver damage and has beneficial metabolic effects in ob/ob<br />

mice. Material & Methods: ob/ob mice received either a diet<br />

supplemented with 0.5% NorUDCA or chow diet for 6 weeks.<br />

Intraperitoneal glucose and insulin tolerance tests (IPGTT, IPITT)<br />

were performed at week 5 and 6. Serum and urine biochemistry,<br />

liver and fat histology were analyzed, gene and protein<br />

expressions of key markers of inflammation, lipid and glucose<br />

metabolism were assessed by RT-PCR and Western Blotting.<br />

Results: Mice treated with NorUDCA showed a significant<br />

reduction in AST, ALT and AP serum levels (p


690A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

mice compared with wild-type mice displayed elevated plasma<br />

ALT (150 +/-11.94 vs 108+/-4), and more fat accumulation<br />

in the liver. These data suggest that BHB production by the<br />

beta-oxidation pathway, and sensing by GPR019a, protects<br />

from alcohol induced acute liver injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yonglin Chen, Irma Garcia-Martinez,<br />

Xinshou Ouyang, Rafaz Hoque, Wajahat Z. Mehal<br />

980<br />

Interferon lambda 3 (IL28B) and interferon lambda 4<br />

genotypes and resistance-associated variants in HCV<br />

genotype 1 and 3 infected patients<br />

Kai-Henrik Peiffer, Lisa Sommer, Simone Susser, Julia Dietz, Dany<br />

Perner, Caterina Berkowski, Stefan Zeuzem, Christoph Sarrazin; J.<br />

W. Goethe University, Medizinische Klinik 1, Frankfurt am Main,<br />

Germany<br />

Background and aims: In the era of pegylated interferon<br />

(Peg-IFN) single nucleotide polymorphisms (SNPs) in the interferon<br />

lambda 3 (IFN-L 3, also known as IL28B) and interferon<br />

lambda 4 genes (IFN-L4) were strong predictors for achieving<br />

SVR in patients with hepatitis C virus (HCV) infection. For many<br />

direct acting antiviral (DAA) combination regimens only weak<br />

or no correlation with IFN-L SNPs was observed. However, little<br />

is known about a potential selection of resistance-associated<br />

variants (RAVs) by IFN-L genotypes, which may influence virologic<br />

treatment response. This study aims to analyze the prevalence<br />

of rare and common RAVs in NS3, NS5A and NS5B<br />

to currently approved DAAs in a large European population<br />

in correlation to IFN-L3 and IFN-L4 genotypes. Materials and<br />

methods: Samples of 633 patients chronically infected with<br />

HCV genotype 1a (n=259), 1b (n=323) and 3 (n=51) were<br />

genotyped for rs12979860 (IFN-L3, also known as IL28B) and<br />

ss469415590 (IFN-L4) by real-time PCR. RAVs in NS3, NS5A<br />

and NS5B were detected via population-based sequencing and<br />

correlated with the IFN-L SNPs. Results: No significant correlation<br />

was found between IFN-L genotypes and rare and common<br />

RAVs within NS3 and NS5B genes (i.e. Q80K, C316N and<br />

S556G/N/R). In contrast, the NS5A RAV Y93H was detected<br />

frequently in HCV genotype 1b (14%) and was significantly<br />

associated with the beneficial rs12979860 (IFN-L3) C/C- and<br />

the ss469415590 (IFN-L4) TT/TT-genotype (p=0.0005 and<br />

p=0.0016, respectively). Therefore, prevalence of the Y93H<br />

variant is much higher in IFN-L3 C/C (27%) and IFN-L4 TT/TT<br />

(28%) than in non-C/C (10%) and non-TT/TT (10%) genotyped<br />

patients respectively. In addition, Y93H was found to be significantly<br />

associated with higher viral loads independent from<br />

the IFN-L genotype (p=0.000936 in all HCV genotypes). Summary:<br />

The major NS5A RAV Y93H is significantly associated<br />

with the presence of beneficial IFN-L3 and IFN-L4 SNPs and a<br />

high baseline viral load in HCV genotype 1 infected patients,<br />

which may explain a lack or inverse correlation of treatment<br />

response with IFN-L genotype in NS5A inhibitor containing<br />

IFN-free regimens and might be of clinical interest for consideration<br />

of IFN-free treatment options for patients with known<br />

IFN-L3 or IFN-L4 status. In addition Y93H was found to be<br />

associated with a higher viral load independent from the IFN-L<br />

genotype, which may partially explain in HCV genotype 1b the<br />

known paradox phenomenon of higher viral loads in patients<br />

with the beneficial IFN-L3 genotype.<br />

Disclosures:<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

The following authors have nothing to disclose: Kai-Henrik Peiffer, Lisa Sommer,<br />

Simone Susser, Julia Dietz, Dany Perner, Caterina Berkowski<br />

981<br />

MicroRNA-130a inhibit HCV production through regulating<br />

host immune responses and lipid metabolism<br />

Xiaoqiong Duan 1 , Limin Chen 1,2 ; 1 Institute of Blood Transfusion,Chinese<br />

Academy of Medical Sciences (CAMS) /Peking<br />

Union Medical College (PUMC), Chengdu, China; 2 Toronto General<br />

Research Institute,University of Toronto, Toronto, ON, Canada<br />

Background & Aims: MicroRNAs (MiRNAs)a class of small<br />

non-coding RNAs that regulate messenger RNA (mRNA)<br />

expression, play pivotal role in the regulation of viral infections.<br />

In our previous study, we demonstrated that over-expression<br />

of miR-130a inhibited HCV replication by restoring<br />

the host innate immune response. But the detailed mechanism<br />

remains unclear. In this study, we aimed to identify and validate<br />

the potential target genes of miR-130a to explore the<br />

possible regulation mechanism between miR-130a and host<br />

immune responses in the context of HCV infection. Methods:<br />

Four algorithms, miRanda, TargetScan, PITA and RNAhybrid<br />

were used to predict the potential targets of miR-130a. To validate<br />

the target genes, MiR-130a mimic were transfected to<br />

Huh7.5.1 cells for 48h and total RNAs were extracted and<br />

reverse-transcribed to cDNA. Quantitative real-time PCR and<br />

western blot were used to measure target gene expressions<br />

with or without miR-130a overexpression. Results: Three potential<br />

target genes pyruvate kinase (PKLR), interleukin 18 binding<br />

protein (IL18BP) and low density lipoprotein receptor (LDLR)<br />

were identified by software algorithms and validated by qRT-<br />

PCR. miR-130a over-expression inhibited the expression levels<br />

of PKLR, IL18BP and LDLR significantly both mRNA and at protein<br />

levels. Functional analysis suggested that these genes are<br />

involved in lipid metabolism and host immune response. Quite<br />

interestingly, Cholesterol-25-hydroxylase (CH25H), an interferon-stimulated<br />

gene playing critical role in lipid metabolism<br />

and anti-HCV innate immunity, was stimulated by miR-130a. In<br />

addition, miR-130a overexpression increased the expression<br />

levels of interferon γ (IFNγ) and inhibited HCV entry, possibly<br />

through its targeting to IL18BP and LDLR, respectively. Conclusion:<br />

miR-130a inhibits HCV production through regulating<br />

host immune responses and lipid metabolism. Mir-130a may<br />

be developed as a potential drug candidate for the treatment<br />

of HCV infection.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 691A<br />

Disclosures:<br />

The following authors have nothing to disclose: Xiaoqiong Duan, Limin Chen<br />

982<br />

Individualized DAA treatment duration for cure in<br />

patients with cirrhosis via modeling of early HCV kinetics<br />

Martina Gambato 3 , Laetitia Canini 1,4 , Sabela Lens 3 , Frederik<br />

Graw 5 , María-Carlota Londoño 3 , Susan L. Uprichard 1 , Zoe<br />

Mariño 3 , Enric Reverter Segura 3 , Concepció Bartres 3 , Patricia<br />

Gonzalez 3 , Scott Cotler 1 , Xavier Forns 3 , Harel Dahari 1,2 ; 1 The<br />

Program for Experimental & Theoretical Modeling, Department<br />

of Medicine, Division of Hepatology, Loyola University Medical<br />

Center, Maywood, IL; 2 Theoretical Biology & Biophysics Group,<br />

Los Alamos National Laboratory, Los Alamos, NM; 3 Liver Unit,<br />

Hospital Clínic, IDIBAPS and CIBEREHD, Barcelona, Spain; 4 Centre<br />

for Immunity, Infection and Evolution, University of Edinburgh,<br />

Edinburgh, United Kingdom; 5 Center for Modeling and Simulation<br />

in the Biosciences, BioQuant Center,, Heidelberg University, Heidelberg,<br />

Germany<br />

Background. While recent DAA-based regimens achieve SVR<br />

rates of over 90%, there is still substantial interest in optimizing<br />

length (and cost) of therapy, particularly in difficult-to-cure populations,<br />

such as those with cirrhosis. We aimed to investigate<br />

whether modelling of early HCV-RNA kinetics after therapy<br />

initiation could predict the minimal treatment duration needed<br />

to achieve SVR (cure). Methods. Fifty eight HCV patients who<br />

received DAA-based treatment for 12, 16 or 24 weeks at a<br />

single centre were included. HCV genotype, Child-Turcotte-<br />

Pugh [CTP], MELD, hepatic venous pressure gradient [HVPG],<br />

liver stiffness [LS], platelets count [PLT], albumin level [ALB]),<br />

IL28B, and treatment regimen were recorded. Viral load was<br />

assessed at baseline, at 4, 8 and 24 hours, at days 2, 3, 4,<br />

7, and at weeks 2, 3 and 4 after treatment initiation (or until<br />

target-not-detected (TND); LLOQ: HCV-RNA


692A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Donatella Palazzo, Elisa Biliotti,<br />

Francesca Tinti, Alessandra Bachetoni, Stefania Grieco, Andrea Cappoli, Raffaella<br />

Labriola, Paola Perinelli, Ilaria Umbro, Paola Rucci, Anna Paola Mitterhofer<br />

984<br />

Hepatitis C virus facilitates its persistent infection<br />

through ΔNp63-miR-181a-SIRT1/ DUSP6 signaling<br />

pathways<br />

Yun Zhou 1,2 , Guangyu Li 2,3 , Junping Ren 2,3 , Ying Zhang 1 , Jianqi<br />

Lian 1 , Changxing Huang 1 , Shunbin Ning 2,3 , Jonathan P. Moorman<br />

2,3 , Zhansheng Jia 1 , Zhi Q. Yao 2,3 ; 1 Department of Infectious<br />

Diseases, Tangdu Hospital, Fourth Military Medical University,<br />

Xi’an, China; 2 Department of Internal Medicine, Division of Infectious<br />

Diseases, James H. Quillen College of Medicine, East Tennessee<br />

State University, Johnson City, TN; 3 Center for Inflammation,<br />

Infectious Disease and Immunity, James H. Quillen College of Medicine,<br />

East Tennessee State University, Johnson city, TN<br />

T cells play a crucial role for viral clearance or persistence;<br />

however, the precise mechanisms that control their responses<br />

during viral infection remain largely unknown. MicroRNAs<br />

(miR) have been implicated as key regulators controlling<br />

diverse biological processes through posttranscriptional repression.<br />

Here we demonstrate that hepatitis C virus (HCV)-mediated<br />

regulation of ΔNp63 and miR-181a expression impairs<br />

CD4+ T cell responses to facilitate viral persistence via the<br />

dual specific phosphatase 6 (DUSP6) and SIRT1 signaling pathways.<br />

Specifically, a significant up-regulation of ΔNp63 leads<br />

to a decline of miR-181a expression, concomitant with over-expression<br />

of DUSP6 and SIRT1 in CD4+ T cells from chronically<br />

HCV-infected individuals compared to healthy subjects.<br />

The levels of ΔNp63 expression were found to be negatively<br />

associated with the levels of miR-181a, which in turn correlate<br />

to the expressions of DUSP6 and SIRT1 in these cells. Importantly,<br />

silencing ΔNp63/SIRT1 or reconstitution of miR-181a<br />

expression in CD4+ T cells leads to T cell senescence including<br />

reduced telomerase activity, shortened telomere length,<br />

increased senescence-associated β-galactosidaese (SA-β-gal)<br />

expression, and decreased EdU incorporation. Paradoxically,<br />

an increased IL-2 expression is observed in CD4+ T cells by<br />

these treatments, primarily due to a reduced frequencies of<br />

Foxp3+ regulatory T cells (Tregs) and increased or unchanged<br />

numbers of Foxp3- effector T cells (Teffs) along with manipulating<br />

the ΔNp63-miR181a-Sirt1 pathway. These findings provide<br />

novel mechanistic insights into how HCV induces T cell exhaustion<br />

via ΔNp63-miR-181a-DUSP6 signaling and premature T<br />

cell aging via ΔNp63-miR-181a-SIRT1 pathway. These two<br />

major pathways reveal new targets for therapeutic rejuvenation<br />

of impaired T cell responses during chronic viral infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yun Zhou, Guangyu Li, Junping<br />

Ren, Ying Zhang, Jianqi Lian, Changxing Huang, Shunbin Ning, Jonathan P.<br />

Moorman, Zhansheng Jia, Zhi Q. Yao<br />

985<br />

Effects of Resistance Mutations of NS5A Inhibitor<br />

on Viral Production and Susceptibility to Anti-HCV<br />

Reagents in Recombinant Hepatitis C Viruses with NS5A<br />

of Genotype 1b<br />

Sayuri Nitta 1,2 , Yasuhiro Asahina 1 , Takaji Wakita 2 , Takanobu<br />

Kato 2 ; 1 Gastroenterology and Hepatology, Tokyo Medical and<br />

Dental University, Tokyo, Japan; 2 Virology II, National Institute of<br />

Infectious Diseases, Tokyo, Japan<br />

Backgrounds and Aims: The direct acting antiviral agents<br />

(DAAs) for hepatitis C virus (HCV) have strong inhibitory<br />

potency to HCV. However, using DAAs could cause the emergence<br />

of resistance mutations because these agents directly target<br />

HCV proteins. Amino acid substitutions at L31 and Y93 of<br />

non-structural protein 5A (NS5A) in HCV genotype 1b strains<br />

have been reported to confer resistance to NS5A inhibitor,<br />

daclatasvir (DCV). Furthermore, strains with these mutations<br />

are often detected in DCV treatment naïve patients. In this<br />

study, we assessed the effects of resistance mutations of DCV<br />

on HCV life cycle and susceptibility to DCV and other anti-HCV<br />

reagents by use of recombinant HCV harboring NS5A of genotype<br />

1b strain, Con1. Materials and Methods: We constructed<br />

recombinant HCV harboring NS5A of Con1 (JFH1-5ACon1),<br />

and introduced reported DCV resistance mutations, L31M,<br />

L31V, L31I and Y93H, in solely or in combination. In vitro transcribed<br />

full-length HCV RNA of these strains were transfected<br />

into Huh-7.5.1 cells and evaluated HCV replication and infections<br />

virus production by measuring the extra- and intra-cellular<br />

HCV core antigen (Ag) and infectivity titers. Susceptibilities of<br />

these strains to various anti-HCV reagents were also investigated.<br />

Results: All these strains are capable of replication in<br />

Huh-7.5.1 cells. At day 3 post-transfection, intra-cellular HCV<br />

core Ag levels were comparable in these strains transfected<br />

cells. However, extra-cellular HCV core Ag levels of strains with<br />

Y93H were about 2-fold higher than that of JFH-1/5ACon1.<br />

To assess the affecting step of this mutation in HCV life cycle,<br />

we determined the infectivity titers of these recombinant HCV.<br />

The strains with Y93H indicated higher infectivity titers than<br />

that of JFH-1/5ACon1. In the analysis of susceptibility to DCV,<br />

strains with mutation at L31 or Y93 showed the relatively mild<br />

or moderate resistance, while those with mutations at both L31<br />

and Y93 showed severe resistance, consistent with previous<br />

reports. Strains with DCV resistance mutations exhibited similar<br />

level susceptibility to IFNα, IFNλ1, IFNλ3 or RBV. Interestingly,<br />

the strains with Y93H mutation were more sensitive to protease<br />

inhibitor simeprevir (SMV) in comparison with JFH-1/5ACon1.<br />

Conclusions:In the in vitro study with recombinant HCV with<br />

NS5A of Con1, we indicated that the strains with Y93H<br />

enhanced infectious virus production. This observation suggests<br />

that once strains with Y93H emerged, it might be persistent<br />

after treatment. From our findings in susceptibility analysis to<br />

SMV, regimen including SMV could be considerable to treat<br />

the patients infected HCV with Y93H mutation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sayuri Nitta, Yasuhiro Asahina,<br />

Takaji Wakita, Takanobu Kato<br />

986<br />

The impact of HLA-DQs and IFNL4 variant on hepatitis C<br />

virus-induced liver inflammation<br />

Daiki Miki 1,2 , Hidenori Ochi 1,2 , C. Nelson Hayes 1,2 , Hiromi Abe 1,2 ,<br />

Sakura Akamatsu 1,2 , Atsushi Ono 1,2 , Takashi Nakahara 1,2 , Yizhou<br />

Zhang 1,2 , Keiichi Masaki 1,2 , Hatsue Fujino 1,2 , Eisuke Miyaki 1,2 ,<br />

Hiromi Kan 1,2 , Takuro Uchida 1,2 , Nobuhiko Hiraga 1,2 , Masataka<br />

Tsuge 1,2 , Tomokazu Kawaoka 1,2 , Michio Imamura 1,2 , Yoshiiku<br />

Kawakami 1,2 , Hiroshi Aikata 1,2 , Kazuaki Chayama 1,2 ; 1 Laboratory<br />

for Digestive Diseases, RIKEN Center for Integrative Medical<br />

Sciences, Hiroshima, Japan; 2 Department of Gastroenterology and<br />

Metabolism, Hiroshima University Hospital, Hiroshima, Japan<br />

Background & Aims: Several previous <strong>studies</strong> including genomewide<br />

association <strong>studies</strong> (GWAS) revealed that genetic variation<br />

in the IFNL3 (IL28B)-IFNL4 locus and the MHC region were<br />

associated not only with hepatitis C virus (HCV) persistence<br />

but also with HCV-induced cirrhosis and hepatocellular carcinoma.<br />

Because these variants appear to be associated with<br />

the progression of chronic HCV infection, they might play an


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 693A<br />

important role in chronic liver inflammation, although the role<br />

of these variants is still controversial. We performed large scale<br />

genotyping of IFNL4 and HLA-DQA1/B1, which have been<br />

reported as susceptibility loci for HCV persistence in European<br />

and Japanese populations, and analyzed the association<br />

between these genetic variants and the degree of histological<br />

inflammation in liver biopsies of patients with HCV. Methods:<br />

We directly sequenced exon 2 of HLA-DQA1 and HLA-DBB1<br />

using 775 Japanese patients with chronic HCV infection who<br />

underwent liver biopsies. In addition, we also investigated an<br />

IFNL4 variant (ss469415590) by multiplex-PCR-based Invader<br />

assay. We compared the frequencies of these genetic variants<br />

between patients with severe inflammation (necroinflammatory<br />

activity grade 2-3; n=514) and those with mild inflammation<br />

(activity grade 1; n=261). Results: The IFNL4 major homozygote<br />

TT/TT was more likely to be present in patients with<br />

severe activity grade (P = 0.034). Among common HLA-DQ<br />

alleles, HLA-DQA1*0101 and HLA-DQB1*0501 were<br />

observed more frequently (P < 0.05) in patients with mild activity<br />

(odds ratio [OR] = 0.58 and 0.57, respectively), whereas<br />

HLA-DQA1*0301 and HLA-DQB1*0401 were observed more<br />

frequently in those with severe activity (OR = 1.27 and 1.59,<br />

respectively). After adjustment for multiple testing, only HLA-<br />

DQA1*0101 was significantly associated with activity grade,<br />

and the effect size of this allele was not reduced by adjusting<br />

for IFNL4 variant and age at biopsy. Subsequent haplotype<br />

analysis revealed one significant protective haplotype, HLA-<br />

DQA1*0101-HLA-DQB1*0501 (P = 2.18 × 10 -4 , OR = 0.51),<br />

as well as one suggestive risk haplotype, HLA-DQA1*0301-<br />

HLA-DQB1*0401 (P = 5.12 × 10 -3 , OR = 1.48). Finally, results<br />

of multiple logistic regression analysis suggested that each<br />

haplotype and IFNL4 variant were independently associated<br />

with activity grade. Conclusions: We identified liver inflammation-associated<br />

HLA-DQ alleles and haplotypes that differed<br />

from the HCV persistence/clearance-associated allele, HLA-<br />

DQB1*03. Our findings suggest that HLA-DQ molecules might<br />

play different roles in liver inflammation and viral clearance,<br />

even as part of the same anti-viral immune response.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Daiki Miki, Hidenori Ochi, C.<br />

Nelson Hayes, Hiromi Abe, Sakura Akamatsu, Atsushi Ono, Takashi Nakahara,<br />

Yizhou Zhang, Keiichi Masaki, Hatsue Fujino, Eisuke Miyaki, Hiromi Kan, Takuro<br />

Uchida, Nobuhiko Hiraga, Masataka Tsuge, Tomokazu Kawaoka, Michio<br />

Imamura, Yoshiiku Kawakami, Hiroshi Aikata<br />

987<br />

The Hepatitis C Virus modulates its replication and<br />

the degradation of its RNA through the subversion of<br />

KHSRP function<br />

Patrice Bruscella 1 , Camille Baudesson 1 , Hassan Danso 1 , Michele<br />

Trabucchi 2 , Jean-Michel Pawlotsky 1 , Cyrille Feray 1 ; 1 Unit 955,<br />

INSERM, Creteil, France; 2 INSERM U 1065, C3M, Nice, France<br />

During infection, the hepatitis C virus (HCV) genome is exposed<br />

to degradation by the cellular RNA decay machinery. Indeed,<br />

HCV RNA is a target for cellular RNAses due to the lack of<br />

a 5’ methylated cap and of a 3’ poly-A tail and to the presence<br />

of a destabilizing RNA element (the 3’ poly-U region).<br />

MiR-122 has been shown to protect the HCV genome from<br />

5’-end degradation. The goal of this work was to unravel the<br />

role of KHSRP, an AU-Rich Element (ARE) mRNA-binding protein<br />

capable to recruit 5’- and 3’-RNAses such as XRN1 and<br />

DIS3, respectively, involved in miRNA maturation during HCV<br />

infection. Methods: Transient gene knockdown was performed<br />

by means of siRNA transfection. Viral RNA stability was evaluated<br />

by Northern-Blot analysis and qRT-PCR. Viral replication<br />

was measured using a modified JFH1 strain expressing<br />

Gaussia luciferase. Gene, pri-miRNA and miRNA expressions<br />

were quantified by qRT-PCR. The post-translational modifications<br />

and subcellular localization of endogenous KHSRP were<br />

studied by means of in situ Proximity Ligation Assay. Direct<br />

interaction of KHSRP with pri-miR-122, HCV RNA and ARE<br />

mRNA was studied by RNA Immunoprecipitation experiments<br />

using plasmids expressing KHSRP-HA, KHSRP-HA-S193D<br />

(phosphomimetic) or KHSRP-HA-S193A. Results: XRN1 and<br />

DIS3 silencing increased JFH1 replication in Huh7 hepatoma<br />

cells and the stability of nonreplicative viral RNA, whereas<br />

KHSRP silencing inhibited both viral replication and degradation<br />

of non-replicative HCV RNA. KHSRP silencing dramatically<br />

decreased miR-122 maturation. Using the JFH1 strain and a<br />

con1 subgenomic replicon, we showed that the expression of<br />

structural HCV proteins was associated with a decrease of the<br />

amount of KHSRP mRNA. During infection, the AKT-dependent<br />

phosphorylation of KHSRP at position Ser193 led to nuclear<br />

relocalization of phospho-KHSRP concomitantly to maturation<br />

of nuclear pri-miR-122. Proximity Ligation Assay showed co-localization<br />

of KHSRP, HCV NS5A protein and HCV RNA in<br />

the cytosol. Conclusions: HCV down-regulates KHSRP gene<br />

expression in hepatocytes, thus protecting HCV RNA from<br />

intracellular degradation by XRN1 and DIS3 RNAses. HCV<br />

infection also induces the phosphorylation and relocalization<br />

of phospho-KHSRP to the nucleus, where it induces miR-122<br />

maturation, thereby increasing HCV RNA stability and favoring<br />

replication. This so far unknown regulatory mechanism simultaneously<br />

promotes viral replication through miR-122 maturation<br />

and limits viral RNA degradation.<br />

Disclosures:<br />

Jean-Michel Pawlotsky - Consulting: Abbvie, Achillion, Bristol-Myers Squibb, Gilead,<br />

Janssen, Merck; Speaking and Teaching: Bristol-Myers Squibb, Gilead,<br />

Merck, Janssen<br />

The following authors have nothing to disclose: Patrice Bruscella, Camille Baudesson,<br />

Hassan Danso, Michele Trabucchi, Cyrille Feray<br />

988<br />

Liver pDCs from HCV-Infected Patients Are Primed for<br />

Interferon-alpha Production<br />

Erin H. Doyle 1 , Adeeb Rahman 2 , Arielle L. Klepper 1 , Sang Kim 6 ,<br />

Brandy M. Haydel 3 , Sander S. Florman 4 , M. Isabel Fiel 5 , Thomas<br />

D. Schiano 4 , Andrea D. Branch 1 ; 1 Department of Liver Diseases,<br />

Icahn School of Medicine at Mount Sinai, New York, NY;<br />

2 Human Immune Monitoring Core, Icahn School of Medicine at<br />

Mount Sinai, New York, NY; 3 Center for Translational Transplant<br />

Research, Icahn School of Medicine at Mount Sinai, New York,<br />

NY; 4 Recanati Miller Transplantation Institute, The Mount Sinai<br />

Hospital, New York, NY; 5 Department of Pathology, The Mount<br />

Sinai Hospital, New York, NY; 6 Department of Anesthesiology,<br />

The Mount Sinai Hospital, New York, NY<br />

Background: Hepatitis C virus (HCV)-infected cells must be<br />

cleared from the liver to achieve a cure—whether through<br />

direct-acting antiviral drugs, or immune mechanisms. Interferon<br />

(IFN)α is an important anti-viral cytokine that is produced by<br />

plasmacytoid dendritic cells (pDCs), but its role in HCV clearance<br />

has been questioned because intrahepatic levels of IFNα<br />

mRNA are often below the limit of detection. This study combined<br />

a powerful new technique for cell phenotyping—mass<br />

cytometry time-of-flight (CyTOF)—with other methods to determine<br />

whether intrahepatic pDCs can and do produce IFNα.


694A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Methods: Cells were isolated from explants of 20 HCV-infected<br />

patients undergoing liver transplantation. Blood was obtained<br />

prior to surgery. Liver and peripheral blood mononuclear cells<br />

(PBMCs) were analyzed by CyTOF, FACS, microarray, and<br />

multiplex cytokine assays with and without 4 hr stimulation with<br />

R848 (TLR7/8 agonist/single-stranded RNA mimic) or Poly<br />

I:C (TLR9 agonist/double-stranded RNA analogue). Results:<br />

CyTOF delineated monocytes/macrophages, DC, NK, T and<br />

B cell subsets and revealed that pDCs are the only liver or<br />

blood cell that produces IFNα. Cytokine assays showed that<br />

liver pDCs were primed for robust IFNα production: secreted<br />

levels increased from 0.8 to 344.7 pg/mL after 4 hrs of R848<br />

stimulation. This increase was much higher than that of blood<br />

pDCs, which increased production from 2.4 to 115 pg/mL,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 695A<br />

as shown by increased ISG induction. Interestingly, SeP was<br />

found to regulate IFN signal transduction independent of the<br />

JAK-STAT pathway. In vivo, the ISGs Mx-1 and Oas2 were<br />

significantly increased in SepKO mice compared with WT mice<br />

following PolyI:C injection. To explore the clinical significance,<br />

hepatic gene expression was analyzed in 66 CHC patients;<br />

SeP was upregulated in the early stage of liver fibrosis and was<br />

significantly correlated with a poor outcome of IFN therapy.<br />

Conclusion: In CHC, SeP impaired innate immunity. The present<br />

findings would be useful for the development of novel therapeutic<br />

strategies for patients with chronic liver disease with various<br />

etiologies including NASH and CHC.<br />

Disclosures:<br />

Stanley M. Lemon - Advisory Committees or Review Panels: Merck, Santaris,<br />

Abbott, Gilead; Consulting: Achillion, Idenix; Grant/Research Support: Merck,<br />

Tibotec, Scynexis; Speaking and Teaching: Hoffman LaRoche<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Kazuhisa Murai, Masao Honda,<br />

Tetsuro Shimakami, Takayoshi Shirasaki, Hirofumi Misu, Toshinari Takamura,<br />

Seishi Murakami<br />

991<br />

Long noncoding RNAs regulate HCV replication and<br />

ISGs in a Jak-STAT independent manner<br />

Xiao Liu, Jian Hong, Dachuan Cai, Sae Hwan Lee, Dahlene Fusco,<br />

Esperance A. Schaefer, Cynthia Brisac, Anna Lidofsky, Wenyu Lin,<br />

Raymond T. Chung; Gastrointestinal Unit, Massachusetts General<br />

Hospital, Harvard Medical School, Boston, MA<br />

Background/ Aims: Long non-coding RNAs (lncRNAs) involve<br />

in many important cellular process regulations. However, their<br />

roles in the innate immune response including type-1 interferon<br />

(IFN) response to HCV infection are not well understood.<br />

We previously identified SART1, a host protein involved in<br />

RNA splicing and pre-mRNA processing, as a regulator of<br />

IFN antiviral effects (Lin et al, J Hepatology 2015, 62:1024).<br />

We sought to explore lncRNA regulators of HCV replication,<br />

the JAK-STAT pathway, and interferon-stimulated genes (ISGs).<br />

Methods: We performed SART1 siRNA knockdown and RNA-<br />

Seq in Huh7.5.1 cells with or without IFN treatment. Bioinformatic<br />

Homo sapiens Genome Browser Gateway analysis was<br />

performed to predict the location of lncRNAs and adjacent<br />

coding genes. Selected lncRNAs, Jak-STAT pathway genes and<br />

ISGs and their proteins, together with HCV replication, were<br />

monitored by qRT-PCR and Western blot in the JFH1 HCV infectious<br />

model. Results: We obtained 6699 lncRNAs regulated<br />

by SART1 and IFN in RNA-Seq. Our bioinformatic analysis<br />

of RNA-Seq data identified a cluster of 60 lncRNAs involved<br />

in IFN’s antiviral regulation. qRT-PCR confirmed that expression<br />

of 14 of these lncRNAs was consistent with the RNA-Seq<br />

data. We then identified 4 anti-sense lncRNA related to ISGs<br />

which negatively regulated the adjacent coding gene expression<br />

based on previous reports. Of these, we selected lncRNA<br />

RP11-670E13.5 (lncRNA-TRIM25) for further characterization<br />

of its regulation of HCV. We found that knockdown of lncRNA<br />

RP11-670E13.5 significantly increased HCV replication compared<br />

to Neg siRNA. We also found that siRNA lncRNA RP11-<br />

670E13.5 significantly reduced TRIM25, MX1, ISG15 mRNA<br />

compared to Neg siRNA in JFH1-infected Huh7.5.1 cells.<br />

Interestingly, STAT1, JAK1, OAS3, and RIG-I mRNA expression<br />

were not affected by the knock-down of lncRNA RP11-<br />

670E13.5. Moreover, we found that siRNA to STAT1 had no<br />

effect on lncRNA RP11-670E13.5 expression. These data suggest<br />

that lncRNA RP11-670E13.5 regulation of HCV is independent<br />

of the Jak-STAT signaling pathway. Conclusion: We<br />

found that lncRNA RP11-670E13.5 has important regulatory<br />

functions of antiviral innate immunity in the JFH1 HCV model.<br />

Our data suggested that lncRNA RP11-670E13.5 regulates<br />

HCV replication independently of the JAK-STAT pathway. We<br />

speculate that RP11-670E13.5 exerts its regulatory function on<br />

the enhancer and promoter activation of nearby protein-coding<br />

(target) genes such as TRIM25, MX1, and ISG15. Further characterization<br />

of the mechanism by which selected lncRNAs that<br />

associate with these genes regulate HCV replication and ISGs<br />

are warranted.<br />

Disclosures:<br />

Dahlene Fusco - Grant/Research Support: Gilead<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Xiao Liu, Jian Hong, Dachuan<br />

Cai, Sae Hwan Lee, Esperance A. Schaefer, Cynthia Brisac, Anna Lidofsky,<br />

Wenyu Lin<br />

992<br />

Evaluation of protective and therapeutic strategies<br />

against hepatitis C virus in a syngenic transplantation<br />

mouse model<br />

Matti Sällberg 1 , Sepideh Levander 1 , Fredrik Holmstom 1 , Gustaf<br />

Ahlén 1 , Lars Frelin 1 , Gang Long 2 , Daniel Rupp 2 , Ralf Bartenschlager<br />

2 ; 1 Laboratory Medicine, Karolinska Institutet, Stockholm,<br />

Sweden; 2 Molecular Virology and Infectious Diseases, University<br />

of Heidelberg, Heidelberg, Germany<br />

Key features of the hepatitis C virus (HCV) are its high genetic<br />

plasticity and its ability to establish chronic infection in hepatocytes.<br />

Development of prophylactic vaccines is hampered by<br />

the lack of immune competent small animal models supporting<br />

robust HCV replication. Herein, we devised an immune competent<br />

mouse model that is based on the syngenic transplantation<br />

of H-2 b -restricted Hep56 cells containing a self-replicating<br />

subgenomic HCV replicon RNA of genotype (gt) 2a. To allow<br />

non-invasive in vivo monitoring of tumor growth, cells were<br />

also stably transfected with a luciferase (Luc) reporter gene.<br />

As controls, Hep56 cells stably expressing an enzymatically<br />

active HCV NS3/4A (gt2a) protease complex were used.<br />

Upon subcutaneous injection, all cell lines generated palpable<br />

tumors with sizes peaking around day 8-16 post inoculation.<br />

Tumor growth correlated with an accumulation of inflammatory<br />

cells and central necrotic areas, concomitant with a decrease<br />

of HCV RNA and Luc DNA copy numbers. DNA vaccination<br />

resulted in a robust NS3/4A(gt2a)-specific CD4+ and CD8+<br />

T cell response protecting wild type, but not HCV NS3/4A(gt1a)-transgenic<br />

(Tg) mice against HCV replicon cell tumors suggesting<br />

cross-genotypic tolerance in NS3/4A(gt1a)-Tg mice.<br />

Hep56 cells stably expressing gt2a NS3/4A primed and<br />

boosted a strong CTL-dominated response, whereas the HCV<br />

replicon cells did not, despite comparable levels of antigen<br />

expression. Thus, similar to human infection, hepatocytes with<br />

replicating HCV RNA appear to be poorly immunogenic arguing<br />

that the replicon-cell based syngenic transplantation model<br />

shares features of infected hepatocytes in patients. Unlike other<br />

models, syngenic transplantation of HCV replicon cells can<br />

be rapidly applied to genetically modified immune competent<br />

mice of the widely used C57BL/6J background for the study<br />

of HCV-specific cellular immune responses. This model will be<br />

highly useful in vaccine development for hepatitis C<br />

Disclosures:<br />

Matti Sällberg - Consulting: Chrontech Pharma AB, Abbvie<br />

The following authors have nothing to disclose: Sepideh Levander, Fredrik Holmstom,<br />

Gustaf Ahlén, Lars Frelin, Gang Long, Daniel Rupp, Ralf Bartenschlager


696A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

993<br />

Analysis of AL-516, a Novel Potent Nucleotide Analog,<br />

Combination Drug Interactions in the Hepatitis C Virus<br />

(HCV) Subgenomic Replicon System<br />

Hua Tan, Kenneth Shaw, vivek Rajwanshi, Gary Wang, Leo Beigelman,<br />

David B. Smith, Lawrence M. Blatt, Julian A. Symons;<br />

Alios, South San Francisco, CA<br />

Background: Nucleotide analogs have emerged as an important<br />

component of interferon (IFN)-free combination therapies<br />

for the treatment of chronic hepatitis C (CHC) based on their<br />

potent activity and high barrier to the generation of viral resistance.<br />

AL-516, a novel monophosphate prodrug of a guanosine-based<br />

nucleotide analog, has been identified as a potent<br />

and selective inhibitor of NS5B-directed HCV RNA replication<br />

in the cell based replicon system. In this study, inhibition of the<br />

HCV replicon by AL-516 was examined in pairwise combinations<br />

with multiple compounds either registered for the treatment<br />

of CHC or currently in clinical development. Methods: Studies<br />

were performed using a Huh-7 cell line expressing a Firefly<br />

luciferase-encoding HCV 1b subgenomic replicon. Compounds<br />

were added to cells in a checkerboard fashion and inhibition<br />

of HCV replication measured by luminescence. Data were analyzed<br />

using two drug interaction models; Isobologram analysis<br />

using the Loewe additivity model and the Bliss-Independence<br />

model using Pritchard’s MacSynergy II software. Results: In<br />

the HCV 1b replicon, AL-516 exhibits potent antiviral activity<br />

with an EC 50<br />

of 6.5 nM. When tested in pairwise combinations<br />

with representative members of the NS3/4A protease<br />

inhibitors, NS5A inhibitors, nucleoside/tide and non-nucleoside<br />

polymerase inhibitors, cyclophilin A inhibitors, type I and<br />

type III IFN’s or RBV, AL-516 demonstrated significant synergy<br />

(>100 mM2%), synergy (25-100 mM2%) or additivity (0-25<br />

mM2%), no antagonistic effects were observed. For approved<br />

compounds, combination of AL-516 with the NS3/4A protease<br />

inhibitor, simeprevir, exhibited significant synergy with<br />

a synergy volume of 114.1 mM2%. AL-516 also exhibited<br />

significant synergistic interactions with the HCV NS5A inhibitor,<br />

daclatasvir, (117.1 mM2%) and additive-to-synergistic<br />

interactions with the investigational uridine-based nucleoside<br />

polymerase inhibitor, AL-335, (33.5 mM2%). Conclusions:<br />

Future IFN-free therapy for CHC will require a combination<br />

of compounds with different mechanisms of action. AL-516<br />

demonstrates an in vitro antiviral profile that suggests it may<br />

become an important component of IFN-free combination therapy.<br />

To this end, AL-516 is currently advancing towards human<br />

clinical trials for CHC.<br />

Disclosures:<br />

Kenneth Shaw - Employment: Alios Biopharma<br />

David B. Smith - Employment: Alios BioPharma<br />

Lawrence M. Blatt - Management Position: Alios BioPharma<br />

Julian A. Symons - Employment: Alios BioPharma<br />

The following authors have nothing to disclose: Hua Tan, vivek Rajwanshi, Gary<br />

Wang, Leo Beigelman<br />

994<br />

Regulation of Hepatitis C Virus Infection by Long<br />

Non-Coding RNAs<br />

Tetsuro Shimakami, Masao Honda, Takayoshi Shirasaki, Fanwei<br />

Liu, Masaya Funaki, Kazuhisa Murai, Shuichi Kaneko; Kanazawa<br />

University, Kanazawa, Japan<br />

Background Recently, large numbers of non-coding RNAs<br />

have been identified that reportedly have a wide range of<br />

functions. Non-coding RNAs are categorized into two groups:<br />

the first is composed of short non-coding RNAs, such as siRNA<br />

and miRNA, and the second contains long non-coding RNAs,<br />

which are generally more than 200 ntds in length and possess<br />

a 5ʹ cap and 3ʹ poly-A tail. The role of short non-coding RNAs<br />

in hepatitis C virus (HCV) infection, such as miR-122, has been<br />

characterized considerably more than that of long non-coding<br />

RNAs. Here, we studied the role of long non-coding RNAs<br />

in HCV infection. Method To identify long non-coding RNAs<br />

whose expression is specifically altered by HCV infection,<br />

Huh-7.5 cells were infected with the HJ3-5 virus, which is a<br />

Ia/IIa chimera. Total cellular RNAs were extracted, followed<br />

by amplification of only poly-A tail RNAs, and subjected to<br />

sequence analysis using a next-generation sequencer. We<br />

selected 26 candidate long non-coding RNAs whose expression<br />

was specifically and significantly increased by HCV infection.<br />

Next, we designed 2–3 siRNAs for each candidate and<br />

tested the effect of knockdown of each long non-coding RNA in<br />

HCV cell culture. Result The knockdown of 4 long non-coding<br />

RNAs significantly suppressed HCV replication. From these,<br />

we focused on lncRNA-H, which has already been registered<br />

in the lncRNA database and whose expression in human livers<br />

has been confirmed by several reports; however, its role in<br />

HCV infection remains unknown. The expression of lncRNA-H<br />

was induced by HCV infection in a multiplicity of infection-dependent<br />

manner in HCV cell culture, and NS5A seemed to be<br />

the inducer of lncRNA-H expression. This induction was also<br />

observed in an HCV-infected chimpanzee as well as HCV-infected<br />

chimeric mice. Furthermore, when we compared the<br />

expressions of lncRNA-H in human liver between pre- and<br />

post-eradication of HCV, the expression of lncRNA-H was significantly<br />

decreased by the eradication of HCV. The knockdown<br />

of lncRNA-H by siRNAs suppressed HCV replication<br />

as well as infectious virus production, in accordance with the<br />

knockdown of lncRNA-H. This effect seemed to be through the<br />

suppression of HCV-IRES-dependent translation. Furthermore,<br />

we compared the amount of lncRNA-H in HCV-infected human<br />

livers and serums derived from IL28B genotype (rs8099917)<br />

major and minor allele patients. Interestingly, the amount of<br />

lncRNA-H from minor allele patients was significantly higher<br />

than in patients with the major both in livers and serums. Discussion<br />

These results suggest that lncRNA-H could be a new<br />

target for the treatment of HCV infection as well as a new<br />

biomarker for interferon-based therapies.<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Tetsuro Shimakami, Masao<br />

Honda, Takayoshi Shirasaki, Fanwei Liu, Masaya Funaki, Kazuhisa Murai<br />

995<br />

Iron chelation restores mitophagy suppressed by hepatitis<br />

C virus<br />

Yuichi Hara 1 , Sohji Nishina 1 , Tasuku Hirayama 2 , Hideko Nagasawa<br />

2 , Keisuke Hino 1 ; 1 Kawasaki Medical College, Kurashiki,<br />

Japan; 2 Gifu Pharmaceutical University, Gifu, Japan<br />

Background and aim: Oxidative stress is present in chronic<br />

hepatitis C to a greater degree than in other inflammatory<br />

liver disease and closely related to disease progression. We<br />

recently reported that hepatitis C virus (HCV) core protein<br />

suppresses mitophagy by interacting with Parkin (Am J Pathol<br />

2014), which may amplify and sustain HCV-induced oxidative<br />

stress. Therefore, the restoration of mitophagy is a critical therapeutic<br />

intervention for preventing disease progression including<br />

liver cancer development in chronic hepatitis C. On the other


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 697A<br />

hand, iron loss has been reported to trigger PINK1/Parkin-independent<br />

mitophagy. The aim of this study was to determine<br />

whether iron chelation restores mitophagy suppressed by HCV.<br />

Methods: HCV-JFH1 infected cells were treated for 24h with a<br />

final concentration of 1mM deferiprone (DFP), iron chelator.<br />

Transgenic mice expressing the HCV polyprotein were administered<br />

0.075g of DFP dissolved in water through gastric tube<br />

for 2 months. The effect of DFP on mitophagy and mitochondrial<br />

function was examined in vitro and in vivo. Results: As<br />

reported previously, suppressed mitophagy was confirmed in<br />

HCV-JFH1 infected cells. DFP treatment increased the expression<br />

of microtubule-associated protein light chain 3 (LC3)-II and<br />

decreased the expression of p62 in dose-dependent manner,<br />

but did not affect translocation of Parkin to mitochondria in<br />

HCV-JFH1 infected cells. Electron microscopy also revealed<br />

that DFP significantly increased the number of mitophagosomes<br />

in HCV-JFH1 infected cells and liver specimens from transgenic<br />

mice expressing the HCV polyprotein. These results showed<br />

that DFP restored mitophagy suppressed by HCV core protein.<br />

DFP also decreased mitochondrial membrane potential<br />

and reactive oxygen species (ROS) production, but not ATP<br />

production in HCV-JFH1 infected cells. Next, we measured<br />

mitochondrial ferrous iron content, using a highly specific<br />

turn-on fluorescent probe (Ac-MT-FluNox1) that is specific to<br />

labile ferrous iron in mitochondria. Mitochondrial ferrous iron<br />

content was significantly higher in cells with HCV-JFH1 infection<br />

than in those without HCV-JFH1 infection in the absence of<br />

DFP treatment. DFP decreased mitochondrial ferrous iron content<br />

in dose dependent manner regardless of HCV infection.<br />

Conclusion: These results indicated that iron chelation restores<br />

mitophagy suppressed by HCV. The mechanisms underlying<br />

mitophagy induced by iron chelation remains to be unknown,<br />

but decrease in mitochondrial ferrous iron and disrupted mitochondrial<br />

function may be critical for inducing mitophagy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yuichi Hara, Sohji Nishina,<br />

Tasuku Hirayama, Hideko Nagasawa, Keisuke Hino<br />

996<br />

Cell-specific peripheral innate immune responses to<br />

immunomodulatory stimulation in chronic HCV<br />

Jacinta A. Holmes 1,2 , Narelle A. Skinner 2 , Mario Congiu 2 , Rosemary<br />

M. Millen 2 , Lijia Yu 2 , William Sievert 3 , Paul V. Desmond 1 ,<br />

Kumar Visvanathan 2 , Alex J. Thompson 1,2 ; 1 Gastroenterology, St<br />

Vincent’s Hospital, Fitzroy, VIC, Australia; 2 Immunology Research<br />

Centre, St Vincent’s Hospital, Melbourne, VIC, Australia; 3 Gastroenterology,<br />

Monash Health, Monash University, Melbourne, VIC,<br />

Australia<br />

INTRODUCTION: Our preliminary data showed chronic HCV<br />

(CHC) patients (pts) who achieve SVR have potent and early<br />

immune responses, with chemokine, cytokine and interferon-stimulated<br />

gene (ISG) induction. The pattern is similar in CC<br />

IFNL3 genotype (gt) pts. There are no data regarding which<br />

cell types contribute most to immunomodulatory responses or<br />

which immune pathways are important in CHC. We hypothesised<br />

monocytes (MC) would display the greatest induction<br />

of ISGs post-stimulation, particularly in CC pts, and viral sensing<br />

TLR pathways would generate the greatest responses. We<br />

performed a pilot study to characterise cell-specific ISGs in<br />

CHC PBMCs pre/post immunomodulatory stimulation. METH-<br />

ODS: Treatment-naïve HCV1 F0-3 pts were enrolled. PBMCs<br />

were stimulated with TLR2, 3, 4, 7/8, 9 ligands and IFNa.<br />

PBMCs were sorted into NKs, NKTs and MC using a cell<br />

sorter. RNA was extracted, and ISG expression (MX1/ISG15)<br />

measured (rtPCR). Levels were compared between cell types,<br />

pre/post-stimulation (fold-change), and according to IFNL3 gt.<br />

RESULTS: 10 pts were included: 50% CC, 40% male, median<br />

age 53 years. At baseline non-CC pts demonstrated higher ISG<br />

mRNA, predominantly from NK/NKT cells (Fig.1a). Post-stimulation,<br />

ISG induction was significantly higher in MC (131 to<br />

168-fold increase) than NK/NKT cells (maximum 60-fold-increase)<br />

following TLR7/8 and IFNa stimulation (Fig.1b). MC<br />

from CC pts demonstrated a 2-3 fold-increase in MX1 than<br />

non-CC pts following stimulation (Fig.1c). CONCLUSIONS:<br />

The data identify MC as key immunomodulatory cells, and<br />

demonstrate the importance of the TLR7/8 pathway in immune<br />

responses in CHC, supporting clinical development of TLR7<br />

agonists. Furthermore, MC from CC IFNL3 gt pts demonstrated<br />

greater immune reactivity, and may contribute to the more<br />

potent immune responses observed during therapy, leading to<br />

greater viral clearance. Unraveling these differences in immune<br />

responses may lead to new therapeutic targets, particularly for<br />

patients who fail IFN-free therapies.<br />

Disclosures:<br />

Jacinta A. Holmes - Grant/Research Support: Roche, Merck, AASLD, St Vincent’s<br />

Hospital Research Endowment Fund<br />

William Sievert - Advisory Committees or Review Panels: Gilead Sciences, Bristol<br />

Myers Squibb, Merck; Speaking and Teaching: Merck, Bristol Myers Squibb<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

The following authors have nothing to disclose: Narelle A. Skinner, Mario Congiu,<br />

Rosemary M. Millen, Lijia Yu, Paul V. Desmond, Kumar Visvanathan<br />

997<br />

Role of very-low-density lipoprotein (VLDL) biogenesis<br />

in hepatitis C virus (HCV) production: a reassessment in<br />

primary human adult hepatocytes<br />

Veronique Pene 1 , Maxime Villaret 1 , Matthieu Lemasson 1 , Lynda<br />

Aoudjehane 2,3 , Jean-François Méritet 1,4 , Filomena Conti 3,5 , Yvon<br />

Calmus 3,5 , Arielle R. Rosenberg 1,4 ; 1 EA 4474 “Hepatitis C Virology”,<br />

Université Paris Descartes, Paris, France; 2 Human HepCell,<br />

Hôpital Saint-Antoine, Paris, France; 3 UMRS 938 “CDR Saint-Antoine”,<br />

Inserm, Paris, France; 4 Service de Virologie, AP-HP, Hôpital<br />

Cochin, Paris, France; 5 Unité de Transplantation Hépatique,<br />

AP-HP, Hôpital Pitié-Salpêtrière, Paris, France<br />

Background and Aim. In the blood of HCV-infected patients,<br />

infectivity is mainly supported by viral particles associated<br />

with triglyceride-rich lipoproteins containing apolipoprotein B<br />

(ApoB) and ApoE. These complexes are believed to assemble<br />

within the hepatocyte, which is both the primary replication site


698A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of HCV and the cell type specialized in the secretion of VLDL.<br />

The known requirements for VLDL biogenesis are (i) ApoB, a<br />

structural component essential for VLDL scaffolding, and (ii)<br />

microsomal triglyceride transfer protein (MTP), the rate-limiting<br />

enzyme that lipidates ApoB. However, <strong>studies</strong> with the classical<br />

HCV culture system in hepatocarcinoma-derived Huh-7 sublines<br />

suggested that MTP inhibitors might not efficiently block<br />

HCV production unless high, cytotoxic concentrations are used.<br />

Moreover, the role of ApoB remains a matter of controversy<br />

whereas ApoE appears to be a prerequisite for HCV production<br />

in cell line culture. Here we have reassessed the role of VLDL<br />

biogenesis in HCV morphogenesis and secretion using a more<br />

relevant HCV culture system in primary human adult hepatocytes<br />

(PHH), which, contrary to Huh-7 cells, secrete authentic<br />

VLDL and infectious particles. Methods. PHH were infected with<br />

the HCV strain JFH1, and either treated with increasing doses<br />

of various MTP inhibitors or depleted for ApoB or ApoE (RNA<br />

interference technologies). Cultures were evaluated for production<br />

of intracellular and extracellular infectious virus (focus-formation<br />

assay), viral load (RT-qPCR), secretion of ApoB and<br />

ApoE (ELISA), and cytotoxic effects. Results. The pharmacological<br />

MTP inhibitor CP-346086, which efficiently inhibited<br />

VLDL secretion in PHH, also reduced both HCV production<br />

and infectivity at moderate, non-cytotoxic doses. The grapefruit<br />

flavonoid naringenin, reported to inhibit MTP activity indirectly<br />

through a PPARa-mediated mechanism, induced a significant<br />

reduction of HCV production even at doses that did not reduce<br />

VLDL secretion. ApoB depletion did not seem to affect HCV production<br />

or infectivity significantly even though it inhibited VLDL<br />

biogenesis. ApoE depletion caused only a moderate reduction<br />

of HCV infectivity in PHH whereas it almost completely abolished<br />

HCV production in Huh-7.5.1 cells. Conclusion. These<br />

data in differentiated human hepatocytes show that neither<br />

ApoB nor ApoE is essential for HCV morphogenesis whereas<br />

MTP function is absolutely required for HCV production and<br />

infectivity, albeit not necessarily via its role in VLDL biogenesis.<br />

MTP inhibitors, either direct (CP-346086) or indirect (naringenin),<br />

appear as promising host-targeting drugs to combat<br />

HCV infection in complement with directly acting antivirals.<br />

Disclosures:<br />

The following authors have nothing to disclose: Veronique Pene, Maxime Villaret,<br />

Matthieu Lemasson, Lynda Aoudjehane, Jean-François Méritet, Filomena Conti,<br />

Yvon Calmus, Arielle R. Rosenberg<br />

998<br />

A miR-122/miR-224-based model to identify severe<br />

fibrosis and cirrhosis in patients with chronic hepatitis C<br />

Kevin Appourchaux 2 , Emilie Estrabaud 2 , Matthieu Resche-Rigon<br />

3,1 , Martine Lapalus 2 , Michelle Martinot-Peignoux 2 , Nathalie<br />

Boyer 4 , Michel Vidaud 5 , Pierre Bedossa 6 , Patrick Marcellin 2 , Tarik<br />

Asselah 2,4 ; 1 INSERM, Paris, France; 2 INSERM UMR1149, Paris,<br />

France; 3 INSERM UMR1153, Paris, France; 4 Hepatology, Beaujon<br />

Hospital, Clichy, France; 5 INSERM UMR745, Paris, France; 6 Service<br />

d’Anatomie Pathologique, Beaujon Hospital, Clichy, France<br />

Background and aims Staging fibrosis is crucial in patients<br />

with chronic hepatitis C (CHC) because it reflects the prognosis.<br />

MiRNAs regulate the expression of up to 60% of mRNAs.<br />

The liver-enriched miR-122 enhances HCV replication. MiRNAs<br />

are increasingly investigated as biomarkers because of their<br />

high stability compared to mRNAs and proteins. We aimed<br />

to identify miRNAs associated with fibrosis in patients with<br />

CHC. Patients and Methods A total of 202 patients with CHC<br />

were consecutively enrolled. The inclusion criteria were having<br />

at least one biopsy and no HCC. Serums and biopsies<br />

samples were available for respectively 106 and 86 patients<br />

(26 paired liver/serum). Among patients with available serum,<br />

63.2% were male, the mean age was 48.8 years and the<br />

mean HCV viral load was 5.81 logUI/mL. According to the<br />

Metavir scoring system, 24.5%, 16%, 34.9%, 24.6% of the<br />

patients were respectively F1, F2, F3 and F4. Among patients<br />

with available biopsies, 54.2% were male, the mean age was<br />

49.9 years and the mean HCV viral load was 5,6 logUI/mL.<br />

Regarding the stages of fibrosis, 27,1%, 31,2%, 22,9%, and<br />

18,8% of the patients were respectively F1, F2, F3 and F4.<br />

MiR-20a, -27b, -29a, -92a, -122, -146a, -222, -224 were<br />

selected for the study because of their role during liver fibrosis.<br />

The expression of miRNAs was assessed by RT-qPCR. Results<br />

An increased expression of hepatic miR-224 (p=0.003) was<br />

observed in patients with mild and moderate fibrosis (F1-F2)<br />

compared to those with severe fibrosis and cirrhosis (F3-F4).<br />

A reduction of hepatic miR-122 (p=0.008) was observed in<br />

patients with F3-F4 compared to those with F1-F2. For hepatic<br />

miR-20a, -27b, -29a, -92a, -122 and -146a, no difference was<br />

observed between patients with F3-F4 and those with F1-F2.<br />

In the serums, none of the miRNAs tested, was significantly<br />

deregulated between F3-F4 and F1-F2. The multivariate analysis<br />

of clinical data (albumin, platelets count, aspartate aminotransferase<br />

(AST), alkaline phosphatase) and hepatic miRNAs,<br />

allowed us to build a model combining hepatic miR-122 and<br />

miR-224, platelets count and AST. The AUC of the model was<br />

0.90 while in the same cohort, FIB-4 and APRI had, respectively,<br />

an AUC of 0.78 and 0.85. Conclusions Interestingly,<br />

hepatic miR-122 and miR-224 were respectively decreased<br />

and increased in patients with severe fibrosis and cirrhosis.<br />

Increased expression of miR-224 might suggest a premalignant<br />

condition to HCC. The model combining the assessment of<br />

hepatic miR-122 and miR-224, platelets count and AST was<br />

more accurate than FIB-4 and APRI to distinguish patients with<br />

F3-F4 from those with F1-F2.<br />

Disclosures:<br />

Nathalie Boyer - Board Membership: MSD, JANSSEN, Gilead, Abbvie; Speaking<br />

and Teaching: BMS<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

The following authors have nothing to disclose: Kevin Appourchaux, Emilie<br />

Estrabaud, Matthieu Resche-Rigon, Martine Lapalus, Michelle Martinot-Peignoux,<br />

Michel Vidaud, Pierre Bedossa<br />

999<br />

Different HCV genotypes have different genetic barriers<br />

to resistance to N5A inhibitors, explaining different<br />

treatment outcomes<br />

Slim FOURATI 1,2 , Stephane Chevaliez 1,2 , Alexandre Soulier 1,2 ,<br />

Marion Lavert 1,2 , Lila Poiteau 1,2 , Jean-Michel Pawlotsky 1,2 ; 1 Virology,<br />

Hôpital Henri Mondor, AP-HP,, National Reference Center for<br />

Viral Hepatitis B, C and Delta,, Créteil, France; 2 INSERM U955,<br />

Université Paris-Est, Créteil, France<br />

The genetic barrier to resistance of an antiviral drug depends<br />

on the number and type of nucleotide substitutions required to<br />

overcome the drug’s selective pressure. It is a key factor for<br />

the development of HCV drug resistance in the clinical setting.<br />

The goal of this study was to assess whether different<br />

HCV genotypes and subtypes have different genetic barriers<br />

to resistance to NS5A inhibitors that could explain the clinical<br />

results of NS5A inhibitor-containing regimens. Methods:<br />

744 unselected sequences extracted from the European HCV<br />

database (https://euhcvdb.ibcp.fr) were examined at 9 amino


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 699A<br />

acid positions located in domain I of NS5A (residues 28, 29,<br />

30, 31, 32, 58, 62, 92 and 93); amino acid changes at these<br />

positions are known to be responsible for reduced susceptibility<br />

to one or several NS5A inhibitors, including ledipasvir,<br />

daclatasvir, ombitasvir and/or elbasvir. The genetic barrier<br />

was calculated as the sum of transitions (scored as 1) and/<br />

or transversions (scored as 2.5) required for evolution to any<br />

drug resistance mutation in NS5A. Comparison between genotypes<br />

and subtypes (1a, 1b) was performed based on the most<br />

frequent codon in the genotype/subtype. Results: The genetic<br />

barrier was identical across all genotypes only at NS5A positions<br />

29 and 32 (P29S and P32L scored 1 for all 6 genotypes).<br />

In contrast, a higher barrier to resistance was found in subtype<br />

1b than in subtype 1a at positions M/L28T (score: 3.5<br />

vs 1, respectively), M/L28V (2.5 vs 1), M/L28A (3.5 vs 2),<br />

Q/R30E/H/K (3.5 vs 2.5) and H/P58D (7.5 vs 2.5). Conversely,<br />

subtype 1b displayed a lower genetic barrier than 1a<br />

to acquire H/P58S (1 vs 3.5, respectively). Genotypes 3 and<br />

4 displayed a lower genetic barrier to acquire A/R30S than<br />

genotype 1 (2.5 vs 3.5, respectively). Residue Y93 was highly<br />

conserved across genotypes 1, 2 and 3. Indeed, in contrast to<br />

genotype 6 which displayed a high genetic barrier to acquire<br />

93C (score 2.5) and 93H (score 5), these amino acid substitutions<br />

required a minimal score of 1 for genotypes 1, 2 and 3.<br />

Conclusions: HCV subtype 1a displays a lower genetic barrier<br />

to resistance than subtype 1b at several NS5A positions at the<br />

nucleotide level. This could account for the higher rate of failure<br />

of NS5A inhibitor-containing regimens in patients infected with<br />

HCV subtype 1a. In addition, genotype-specific amino acid<br />

changes (e.g. R30S in genotype 4, P58S in subtype 1b) could<br />

be explained by a differential genetic barrier between different<br />

genotypes and subtypes.<br />

Disclosures:<br />

Slim FOURATI - Speaking and Teaching: Gilead<br />

Stephane Chevaliez - Advisory Committees or Review Panels: Janssen; Speaking<br />

and Teaching: Gilead, BMS<br />

Jean-Michel Pawlotsky - Consulting: Abbvie, Achillion, Bristol-Myers Squibb, Gilead,<br />

Janssen, Merck; Speaking and Teaching: Bristol-Myers Squibb, Gilead,<br />

Merck, Janssen<br />

The following authors have nothing to disclose: Alexandre Soulier, Marion Lavert,<br />

Lila Poiteau<br />

1000<br />

Establishment and Characterization of a New Permissive<br />

Cell Line for HCV Infection<br />

Hitoshi Omura, Tetsuro Shimakami, Takayoshi Shirasaki, Kazuhisa<br />

Murai, Masaya Funaki, Fanwei Liu, Takehiro Hayashi, Taro<br />

Yamashita, Masao Honda, Shuichi Kaneko; Kanazawa university,<br />

Kanazawa city, Japan<br />

Background The in vivo HCV infection model has been needed<br />

to clarify the mechanism of persistent infection of HCV in human<br />

livers. Huh-7.5, which is a subline of human hepatoma cell line,<br />

Huh-7, has been used most frequently for propagation of HCV<br />

due to its vigorous ability to support for HCV replication. Huh-<br />

7.5 is quite useful for the development of many direct-acting<br />

antivirals (DAAs). Huh-7.5 is known to have the defect in its<br />

antiviral interferon signaling, therefore, a new cell line with<br />

intact interferon signaling is needed to mimic HCV infection<br />

in human livers. Here, we have established and characterized<br />

a new permissive cell line for HCV infection with intact innate<br />

antiviral interferon signaling. Method The human hepatoma<br />

tissue was obtained from an HCV-infected patient at a surgical<br />

resection, and then transplanted to mice. The adherent cells to<br />

mice were removed, followed by the MACS cell separation to<br />

remove mice fibroblasts, and used for developing a new cell<br />

line, designated as KH cell. We examined the permissiveness<br />

of KH cells for HCV and characterized by comparing with<br />

Huh-7.5 cells. Result The HCV RNA was no longer detected in<br />

KH cells by qRT-PCR from the original patient. The karyotype<br />

analysis for KH and Huh-7.5 cells revealed that the two cell<br />

lines were distinct from each other. KH cells could support HCV<br />

replication as efficiently as Huh-7.5 cells after transfection of in<br />

vitro transcribed HCV RNA. Additionally, the infection of cell<br />

culture-derived HCV to KH cells showed the entry of HCV into<br />

the cells, followed by efficient replication. KH cells replicating<br />

HCV could also produce infectious viruses, suggesting that KH<br />

cells had the ability to support a full life cycle of HCV as could<br />

Huh-7.5 cells. KH cells had same amount of miR-122 as did<br />

Huh-7.5. We measured EC50 for IFN-α in KH and Huh-7.5<br />

cells. We then compared EC50 of IFN-α between both and<br />

found a similarity. Interestingly, the sequence analysis of RIG-I<br />

at its 55 amino acid position for KH cells showed wild type,<br />

T55. RIG-I from Huh-7.5 cells was mutant, T55I, which is a<br />

representative mutation of Huh-7.5 cells. In accordance with<br />

this result about RIG-I, the transfection of HCV RNA into KH<br />

cells induced robust ISGs induction, such as MX1, IFN-β, and<br />

OAS2, on the other hand, no induction was observed in Huh-<br />

7.5 cells. Discussion We have established and characterized a<br />

new permissive cell line for HCV infection with intact antiviral<br />

interferon signaling. This cell line could be useful to understand<br />

the underlying mechanism of persistent HCV infection under<br />

pressure of host antiviral innate immunity.<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Hitoshi Omura, Tetsuro Shimakami,<br />

Takayoshi Shirasaki, Kazuhisa Murai, Masaya Funaki, Fanwei Liu, Takehiro<br />

Hayashi, Taro Yamashita, Masao Honda<br />

1001<br />

PD-1 blockade enhances cytopathic control of HCV by<br />

antiviral CD8 T cells in infectious HCV co-culture model<br />

Keisuke Ojiro 1 , Xiaowang Qu 1,2 , Jang-June Park 1 , Hyosun Cho 1,3 ,<br />

Kyong-Mi Chang 1 ; 1 Division of Gastroenterology, University of<br />

Pennsylvania, Philadelphia, PA; 2 Translational Medicine Institute,<br />

University of South China, Chenzhou, China; 3 Pharmacy, Duksung<br />

Women’s University, Seoul, Korea (the Republic of)<br />

Background: T-cell expression of inhibitory costimulatory<br />

receptor PD-1 and CTLA-4 in chronic viral infection and tumor<br />

environment is associated with functional tolerance that may<br />

be reversed by antibody-mediated blockade in vitro and in<br />

vivo. Thus, PD-1 blockade is in both pre-clinical and clinical<br />

development in various settings. In this study, we examined<br />

the effect of PD-1 blockade on T cell mediated virus control<br />

and hepatocyte survival using the infectious HCV co-culture<br />

system. Method: Target cells for HCV replication and antigen<br />

presentation were established by lentiviral transduction of HLA<br />

A2.1, PDL1 and/or GFP expression in Huh7.5 hepatoma<br />

cells. Highly infectious JFH derived strain Jc1/Gluc2A clone<br />

(genotype 2a) was modified by site-directed mutagenesis to<br />

encode well-defined genotype 1a-derived NS3 1073 or NS5B<br />

2594 epitopes. HCV-specific CD8 T cells were expanded from<br />

peripheral lymphocytes of HCV-resolvers following 1-2 weeks<br />

of HCV peptide stimulation or engineered by transducing CD8<br />

T cells from uninfected blood donors by lentivirus encoding<br />

HCV-TCR specific for HLA-A2 restricted NS3 1073- or NS5B<br />

2594-epitopes. HCV-specific or non-specific T cells were co-cultured<br />

with HCV-infected Huh7.5 cells at varying E/T ratio, culture<br />

duration and inhibitory receptor blockade, with the level


700A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of HCV-expression in Huh7.5 cells confirmed by anti-NS5A Ab<br />

and HCV-specific T cell activation examined by MHC peptide<br />

tetramer, intracellular cytokine staining and CD107a mobilization<br />

using FACS. Result: Jc1-1073-1a and Jc1-2594-1a with<br />

genotype 1a-derived HCV epitopes could infect over 90% of<br />

Huh7.5 cells and support endogenous processing/presentation<br />

of 1a-based CD8 T cell epitopes to activate HCV-specific CD8<br />

T cells in-vitro. Co-culture with HCV-specific CD8 T cells and<br />

HCV-replicating Huh7.5 cells resulted in reduced number of<br />

total and HCV-infected Huh7.5 cells. PD-1 blockade enhanced<br />

HCV-specific CD8 T cell activation with reduced HCV expression<br />

as well as viability of Huh7.5 cells. Conclusion: PD-1<br />

blockade may enhance virus control through cytopathic pathway<br />

in an infectious HCV co-culture system, highlighting the<br />

need for caution in clinical application of checkpoint inhibitors.<br />

*acknowledgment: HCV and T-cell reagents were generously<br />

provided by Drs. Takaji Wakita, Charles Rice, Brett Lindenbach,<br />

Stanley Lemon, Jim Riley, Annelise Vuidepot, Peter Molloy<br />

and Alan Bennett.<br />

Disclosures:<br />

Kyong-Mi Chang - Consulting: Genentech, Tekmira, Alnylam; Stock Shareholder:<br />

BMS (spouse employment)<br />

The following authors have nothing to disclose: Keisuke Ojiro, Xiaowang Qu,<br />

Jang-June Park, Hyosun Cho<br />

1002<br />

HCV RNA and HCV core antigen are frequently detectable<br />

in stool of men chronically infected with HCV: Is<br />

feces a potential source of infection?<br />

Benjamin Heidrich 1,3 , Eike Steinmann 4 , Iris Plumeier 2 , Janina<br />

Kirschner 1 , Lisa Sollik 1 , Szilvia Ziegert 1 , Patrick Lehmann 1 ,<br />

Michael Engelmann 4 , Tim Lankisch 1 , Michael P. Manns 1,3 , Dietmar<br />

H. Pieper 2,3 , Heiner Wedemeyer 1,3 ; 1 Dept. of Gastroenterology,<br />

Hepatology and Endocrinology, Hannover Medical School, Hannover,<br />

Germany; 2 Microbial Interactions and Processes Research<br />

Group, Centre for Infection Research, Braunschweig, Germany;<br />

3 Partner site Hannover-Braunschweig, German Center for Infection<br />

Research (DZIF), Hannover-Braunschweig, Germany; 4 Institute for<br />

Experimental Virology, TWINCORE Centre for Experimental and<br />

Clinical Infection Research, Hannover, Germany<br />

Background and Aims: HCV is one of the leading causes of<br />

liver cirrhosis and hepatocellular carcinoma worldwide. Transmission<br />

of HCV is thought to be mainly parenteral. However,<br />

unprotected anal intercourse seems to be a risk factor for acquisition<br />

of HCV in men having sex with men (MSM) as well as in<br />

heterosexual partnerships. HCV has been detected in blood,<br />

saliva, and bile. Surprisingly, there are no <strong>studies</strong> investigating<br />

the presence of HCV in stool. Methods: Stool samples of 98<br />

patients were prospectively collected, mixed with RNAlater<br />

and stored at -20°C. RNA was isolated using TRIzol reagent<br />

and transcribed in cDNA by using the SuperScript III kit with<br />

random hexamers. Specific HCV primers were used to identify<br />

samples positive for HCV RNA. PCR results were confirmed by<br />

Sanger sequencing. HCV RNA positive samples were tested for<br />

occult blood using the hemoCARE guajak test and were tested<br />

for HCVcoreAg using the Architect HCVAg from Abbott. Additionally,<br />

viral stability of recombinant HCV particles was investigated<br />

in vitro by incubation of the genotype 2a chimeric virus<br />

Jc1 with different bile and stool suspensions. Results: Overall,<br />

68/98 (69%) stool samples tested positive for HCV RNA. Men<br />

tested more often positive for HCV RNA in stool than women<br />

(83% vs. 52%; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 701A<br />

factor for HCV infection in Vero cells. We also supplemented<br />

Vero cells with the human Apolipoprotein E (ApoE), one of<br />

the host factors important for virus production, and found that<br />

ApoE supplementation enabled them to produce infectious<br />

viruses. Finally, we established Vero cells expressing all these<br />

factors, miR-122, human SR-B1 and ApoE. In this cell line, HCV<br />

replication and infectious virus production could be observed<br />

after HCVcc infection, indicating the reconstitution of the entire<br />

life cycle of HCV. In conclusion, we demonstrate that miR-122,<br />

SR-B1 and ApoE are necessary and sufficient for the reconstitution<br />

of the complete HCV life cycle in non-human non-hepatic<br />

Vero cells. These identified factors may be key determinants for<br />

species and organ tropism of HCV infection. The established<br />

novel HCV cell culture system with Vero cells will be useful to<br />

understand the species and organ specific restriction factors of<br />

HCV and to establish the mass culture system of viral particles<br />

in non-cancer cells for HCV vaccine.<br />

Disclosures:<br />

The following authors have nothing to disclose: Asako Murayama, Nao Sugiyama,<br />

Takaji Wakita, Takanobu Kato<br />

1004<br />

Tim-3 Inhibits Monocyte Functions via T-bet during Hepatitis<br />

C Virus Infection<br />

Ying Zhang 1 , Wenjing Yi 1 , Peixin Zhang 1 , Yan Liang 1 , Zhi Q.<br />

Yao 2 , Zhansheng Jia 1 ; 1 Department of Infectious Diseases, Tangdu<br />

Hospital, Fourth Military Medical University, Xi’an, China; 2 Department<br />

of Internal Medicine, Division of Infectious Diseases, James<br />

H. Quillen College of Medicine, East Tennessee State University,<br />

Johnson City, TN<br />

Hepatitis C virus (HCV) dysregulates innate immune responses<br />

and induces persistent viral infection. We have previously<br />

demonstrated that T-cell immunoglobulin and mucin domain<br />

protein-3 (Tim-3) plays a pivotal role in negative regulation<br />

of Toll-like receptor (TLR)-mediated innate immune responses.<br />

While it is clear that Tim-3 is up-regulated on monocyte/macrophages<br />

(M/M) during HCV infection, little is known about the<br />

transcription factor that controls its expression in these cells.<br />

Recent <strong>studies</strong> have revealed that Tim-3 expression is controlled<br />

by the transcription factor T-box expressed in T cells (T-bet).<br />

In this study, we further investigated the regulatory effects of<br />

T-bet on Tim-3 transcription/translation in M/M during HCV<br />

infection. Our results demonstrate that T-bet is constitutively<br />

expressed on resting peripheral blood CD14+ M/M. M/M<br />

from chronically HCV-infected subjects at baseline exhibit<br />

significantly increased T-bet expression that is positively correlated<br />

with the Tim-3 expression. Up-regulation of T-bet is also<br />

observed in CD14+ M/M incubated with HCV+ Huh7.5 cell in<br />

a time-dependent manner. In addition, HCV core protein can<br />

induce T-bet expression in primary M/M or monocytic THP-1<br />

cells, which is reversible by blocking the HCV core/gC1qR<br />

interactions. Moreover, the HCV core-induced up-regulation of<br />

T-bet and Tim-3 expression in CD14+ monocytes can be abrogated<br />

by incubating the cells with SP600125 - an inhibitor for<br />

JNK signaling pathway. Notably, silencing T-bet gene expression<br />

decreases Tim-3 expression and enhances IL-12 secretion<br />

as well as STAT-1 phosphorylation. These data suggest that<br />

T-bet, induced by HCV core/gC1qR through JNK pathway,<br />

enhances Tim-3 expression, leading to dampened M/M function<br />

during chronic HCV infection. These findings provide new<br />

information regarding Tim-3 transcriptional/translational regulation<br />

via T-bet during HCV infection and novel immunotherapy<br />

targets to combat this global epidemic viral infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ying Zhang, Wenjing Yi, Peixin<br />

Zhang, Yan Liang, Zhi Q. Yao, Zhansheng Jia<br />

1005<br />

Downregulation of let-7 family miRs is strongly associated<br />

with increasing stages of hepatic fibrosis in chronic<br />

hepatitis C (HCV) Genotype 3a<br />

Manish C. Choudhary 1 , Sadaf Dar 1 , Ekta Gupta 2 , Jaswinder S.<br />

Maras 1 , Syed N. Kazim 3 , Gayatri Ramakrishna 1 , Shiv K. Sarin 1 ;<br />

1 Research, Institute of Liver and Biliary Science, New Delhi, India;<br />

2 Virology, Institute of Liver and Biliary Sciences, New Delhi, India;<br />

3 Centre for Interdisciplinary Research in Basic Sciences, Jamia<br />

Milia Islamia, New Delhi, India<br />

Background: Chronic hepatitis C (CHC) infection is a major<br />

cause of liver fibrosis and end stage liver disease. Circulating<br />

miRNAs have evolved as a reliable biomarker for various<br />

pathological conditions. However, there is limited data on systematic<br />

miRNA based biomarker study on hepatic fibrogenesis<br />

in HCV genotype 3a (G3a). Aims: To profile circulating repertoire<br />

of miRNAs in the plasma of HCV G3a infected patients<br />

with different stages of hepatic fibrosis.Patients and Methods:<br />

47 subjects with CHC and histological assessment by two independent<br />

hepato-pathologists were categorized based on stage<br />

of hepatic fibrosis: F0-1 (n=32), F3-4 (n=15) were compared<br />

with healthy controls (n=28). Differentially expressed miRNAs<br />

in plasma of CHC F0,1 (n=4) and F3,4 (n=4) were studied<br />

and compared with controls using miRCURY LNA TM Serum/<br />

Plasma Focus panel. These miRNAs were analysed in context<br />

of fibrosis progression by making comparison between CHC<br />

F0,1 and F3,4 patients. Subsequent validation of array data<br />

was done by stem-loop RT-PCR. miRNA target gene prediction<br />

was done by multiMiR R analysis. Results: 185 miRNAs were<br />

screened and 31 miRNAs were commonly identified in all the<br />

groups. 4 miRNA were significantly down regulated (p


702A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1006<br />

Type III interferon responses contribute to hepatitis virus<br />

infection and depended on IFNL3-IFNL4 variant<br />

Tsunamasa Watanabe 2,1 , Susumu Tsutsumi 2 , Kentaro Matsuura 3 ,<br />

Etsuko Iio 3 , Noboru Shinkai 3 , Kayoko Matsunami 3 , Fumio Itoh 1 ,<br />

Yasuhito Tanaka 2 ; 1 Department of Internal Medicine, Division of<br />

Gastroenterology and Hepatology, St. Marianna University School<br />

of Medicine, Kawasaki, Japan; 2 Department of Virology and Liver<br />

Unit, Nagoya City University Graduate School of Medical Sciences,<br />

Nagoya, Japan; 3 Department of Gastroenterology and<br />

Metabolism, Nagoya City University Hospital, Nagoya, Japan<br />

Introduction: Type III interferon, which consist of interferon-l<br />

(IFNL)1, 2, 3, and 4, has received considerable attention in<br />

hepatotropic hepatitis C virus (HCV), as many independent<br />

genome-wide association <strong>studies</strong> have identified a strong association<br />

between IFNL3-IFNL4 variants and outcome of HCV.<br />

Materials and Methods: We employed an original primary<br />

human hepatocytes (PHH) systems, isolated from chimeric<br />

mice harboring human hepatocytes without cryopreservation,<br />

to investigate innate immune responses against hepatotropic<br />

viruses including mimic of HCV infections and hepatitis B virus<br />

(HBV) infection. The antiviral effects of IFNL were also investigated<br />

in vivo using chronically HCV-infected chimeric mice<br />

transplanted with human hepatocytes of IFNL3-IFNL4 favorable<br />

or unfavorable genotype. Results: In vivo study using pegylated<br />

IFNL1 revealed that the initial viral decline of HCV-RNA by<br />

IFNL1 was similar in both IFNL3-IFNL4 favorable and unfavorable<br />

genotype. Compared to the effect of pegylated IFN-a,<br />

however, the antiviral effects of IFNL was limited, because of<br />

the killer activity of NK cells that was predominantly induced<br />

by only IFN-a, not IFNL1. On the other hands, type III IFN<br />

responses using PHH culture in vitro against 5’-triphosphate<br />

single stranded RNA (PPP-RNA) transfection (mimic of HCV<br />

infection) were varied from IFNL3-IFNL4 variants, although the<br />

levels of IFN-λ genes were significantly higher in PHHs with<br />

favorable genotype than that with unfavorable genotype by<br />

PPP-RNA. The knockdown experiment by small interfering RNA<br />

indicated that cytosol RNA sensor was the key molecule of IFNL<br />

production. When PHHs were infected with HBV, IFN-λ genes<br />

were induced after HBV infection in PHHs with favorable genotype,<br />

but not unfavorable genotype. Conclusions: Our results<br />

suggest that a response of human hepatocytes against hepatotropic<br />

viruses could be contribute to type III IFN production and<br />

the responses might be depended to the IFNL3-IFNL4 variants.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Tsunamasa Watanabe, Susumu<br />

Tsutsumi, Kentaro Matsuura, Etsuko Iio, Noboru Shinkai, Kayoko Matsunami,<br />

Fumio Itoh<br />

1007<br />

Altered frequencies and function of MAIT cells during<br />

treatment of patients with chronic HCV, HIV and HCV/<br />

HIV co-infections<br />

Michelle Spaan 2 , Sebastiaan Hullegie 1 , Kim Kreefft 2 , Gertine van<br />

Oord 2 , Boris Beudeker 2 , Bart J. Rijnders 1 , Robert J. de Knegt 2 ,<br />

Mark Claassen 1 , Andre Boonstra 2 ; 1 Department of Internal Medicine,<br />

Erasmus MC, Rotterdam, Netherlands; 2 Department of<br />

Gastroenterology and Hepatology, Erasmus MC, Rotterdam, Netherlands<br />

Background Mucosal invariant T (MAIT) cells comprise a subpopulation<br />

of T cells that can be activated by a broad range of<br />

bacterial products and cytokines to produce IFN-γ and express<br />

cytolytic enzymes. Loss of MAIT cells is observed in HIV infection,<br />

which is thought to compromise anti-bacterial immunity<br />

and microbial translocation in the gut. Since little is known<br />

on the MAIT cells during HCV infection, we compared their<br />

phenotype and function in comparison to HIV and HCV/ HIV<br />

co-infection, and determined the effect of IFN-based and DAA<br />

antiviral therapy on MAIT cells for HCV. Methods Peripheral<br />

blood was collected before, during and 24 weeks after therapy<br />

from patients with chronic HCV (n=27), virologically supressed<br />

chronic HIV (n=10), HCV/HIV co-infection (n=9) and healthy<br />

individuals (n=12) before, during and after therapy. Patients<br />

with HCV were treated with pegIFN-α/ribavirin with and without<br />

telaprevir, or boceprevir or with IFN-free therapy. MAIT<br />

cells were identified on the basis of CD161 and Vα7.2, and<br />

were stimulated with E.coli, IL-12/IL-18 or IL-18/IFN-α for 18<br />

hours. NK cells and MAIT cells were analysed for the expression<br />

of CD38, CD69 and IFN-γ by flowcytometry. Results Compared<br />

to healthy individuals, the frequency of MAIT cells was<br />

significantly decreased in patients with chronic HCV, HIV and<br />

HCV/ HIV co-infection (5.6%, 3.2%, 2.0% and 1.0% within<br />

CD3+ T cells, respectively). Expression of CD38 on MAIT cells<br />

was comparable in chronic HCV, HIV and healthy individuals,<br />

but was significantly increased in patients with HIV/HCV<br />

co-infection. MAIT cells from healthy controls and the 3 patient<br />

populations were responsive to IFNα in vitro as evidenced by<br />

enhanced frequencies of CD69 expressing and IFN-γ producing<br />

cells. After successful IFN-based therapy the frequencies of<br />

MAIT cells did not normalize. However, we observed that the<br />

functionality of MAIT cells was differentially affected during the<br />

course IFN-α-based therapy and was highly dependent on the<br />

treatment regimen and patient group. Importantly, viral load<br />

decline by IFN-free DAA treatment led to strongly enhanced<br />

IFN-γ levels. Conclusion We show that the frequencies of MAIT<br />

cells are reduced in blood of patients with chronic HCV, HIV<br />

and in HCV/HIV co-infection compared to healthy individuals.<br />

The potent effects on MAIT cells of exposure to IFN-α, both in<br />

vitro and in vivo, and HCV and HIV RNA levels warrant more<br />

detailed <strong>studies</strong> on the interplay between infections to the activity<br />

of this specialized T cell subpopulation.<br />

Disclosures:<br />

Bart J. Rijnders - Advisory Committees or Review Panels: BMS, Abbvie, Gilead;<br />

Grant/Research Support: MSD, Gilead<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

Andre Boonstra - Grant/Research Support: BMS, Janssen Pharmaceutics, Merck,<br />

Roche, Gilead<br />

The following authors have nothing to disclose: Michelle Spaan, Sebastiaan Hullegie,<br />

Kim Kreefft, Gertine van Oord, Boris Beudeker, Mark Claassen<br />

1008<br />

Interferon-based treatment activates regulatory CD4 +<br />

T-cells to exert counteractive antiviral immunity<br />

Bettina Langhans, Hans Dieter Nischalke, Philipp Lutz, Benjamin<br />

Krämer, Christian P. Strassburg, Jacob Nattermann, Ulrich Spengler;<br />

Department of Internal Medicine I, University of Bonn, Bonn,<br />

Germany<br />

Background and aims: Regulatory CD4 + T cells (Tregs) accumulate<br />

during chronic hepatitis C virus (HCV) infection and<br />

inhibit anti-viral T cell responses. The availability of direct-acting<br />

antiviral agents (DAAs), which can be applied in combination<br />

with pegylated interferon (IFN) or as IFN-free regimens<br />

have revolutionized treatment options. However, their effect on<br />

altered immunoregulation in hepatitis C is unclear. Here, we<br />

studied Tregs before and after elimination of HCV with DAAs<br />

combined with IFN versus IFN-free therapy. Methods: Using


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 703A<br />

multi-color flowcytometry we analyzed frequency and function<br />

of Tregs and effector T cells before and at the end of treatment<br />

(treatment week (TW) 12) from HCV-patients after succesful<br />

HCV elimination with sofosbuvir (SOF) in combination with<br />

IFN plus ribavirin (n=14) versus patients treated with SOF plus<br />

daclatasvir (DCV) and simeprevir (SMV), respectively (n=14).<br />

In addition, we measured serum levels of immunomodulatory<br />

cytokines. Results: Unlike IFN-free therapy, treatment with SOF<br />

plus IFN significantly increased numbers of peripheral Tregs<br />

(% Foxp3 + CD25 + CD4 + T cells TW0 vs. TW12 (MW±SD):<br />

5.5±2.0 vs. 7.2±2.4; p90%) of sustained<br />

virological response at week 12 (SVR12) in patients with<br />

HCV infection genotype 4 (Abergel et al. ILC 2015, J Hepatol<br />

2015; 62 (Suppl 2):S219; Poordad et al. ILC 2015, J Hepatol<br />

2015; 62 (Suppl 2):S261). In the former study, two of the three<br />

patients with virological relapse had a genotype 4r subtype.<br />

In the later, NS5A variants were found at baseline in 4 out of<br />

13 patients with virologic failure and in 13 out of 13 patients<br />

at failure. In addition, the number of pre-treatment baseline<br />

RAPs (Resistance Associated Polymorphisms) was associated<br />

with virological failure in a study conducted by Sarrazin et<br />

al. (AASLD, Hepatology 2014; 60 (Suppl) 1128A). Aim: The<br />

objective of this study was to assess the NS5A natural polymorphisms<br />

of the genotype 4 subtypes in order to investigate<br />

the potential role of subtypes and RAPs in treatment outcome<br />

with regimens containing an NS5A inhibitor. Methods: We<br />

collected from the NCBI, European, and Japanese HCV databases,<br />

47 HCV NS5A Genotype 4 sequences and proceeded<br />

to multiple Sequence Alignment using Clustal W2. We analyzed<br />

16 subtype 4a, 7 subtype 4b, 3 subtype 4d, 2 subtype<br />

4m, 3 subtype 4o and 16 subtype 4r (no 4c or 4f were available).<br />

We assessed the prevalence by subtypes of the main<br />

RAPs: L28M, L30R, L31M, P58T and Y93H. Results: Seventy<br />

five percent of the 16 genotype 4r strains included 2 RAPs<br />

(L30R and L31M), in comparison with the 16 genotype 4a<br />

which had only one RAP in all of the strains included (L31M)<br />

(p


704A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

the innate immune response following Poly(I:C) and HCV-RNA<br />

transfection. LECT2 enhanced RIG-I-mediated IFNβ induction<br />

and increased ISG induction through the activation of IRF3,<br />

IRF7 and JAK/STAT pathway. To confirm the biologic significance<br />

of these results, we generated LECT2 transgenic (Tg)<br />

and knockout (KO) mice. LECT2-Tg mice showed increased ISG<br />

induction following Poly(I:C) treatment in the liver. Conversely,<br />

LECT2-KO mice showed impaired ISG induction following IFN<br />

or Poly(I:C) treatment to 30–50% of that of WT mice. Conclusion:<br />

Gene expression profiling of primary hepatocytes treated<br />

with IFNα, IL28B, or both in combination revealed that LECT2,<br />

which enhances the innate immune response and suppresses<br />

HCV replication, was specifically induced by IL28B. LECT2<br />

might participate in prolonged ISG induction by IL28B and in<br />

the unique innate antiviral immune system of IL28B.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Stanley M. Lemon - Advisory Committees or Review Panels: Merck, Santaris,<br />

Abbott, Gilead; Consulting: Achillion, Idenix; Grant/Research Support: Merck,<br />

Tibotec, Scynexis; Speaking and Teaching: Hoffman LaRoche<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Takayoshi Shirasaki, Masao<br />

Honda, Tetsuro Shimakami, Kazuhisa Murai, Hirofumi Misu, Toshinari Takamura,<br />

Seishi Murakami<br />

1011<br />

MicroRNA 17 Host Gene Protein and its targets PTEN<br />

and NF-κB in Chronic Hepatitis C Virus Infection: Relation<br />

to Hepatic inflammation, fibrosis and steatosis.<br />

Hoda El Aggan 1 , Sahah A. Mahmoud 2 , Nevine M. El Deeb 3 ,<br />

Ehab M. Hassona 1 , Sally El Demiry 1 ; 1 Department of Medicine<br />

(Hepatobiliary Unit), Faculty of Medicine, University of Alexandria,<br />

Alexandria, Egypt; 2 Department of Medical Biochemistry, Faculty<br />

of Medicine, University of Alexandria, Alexandria, Egypt; 3 Department<br />

of Pathology, Faculty of Medicine, University of Alexandria,<br />

Alexandria, Egypt<br />

Background/Aim: MicroRNAs (miRs) regulate post-transcriptional<br />

gene expression in various biological processes. The<br />

polycistronic miR-17~92 cluster is comprised of six miRs and<br />

its primary transcript also encodes for a polypeptide of 70<br />

amino acids designated as the miR-17 host gene (MIR17HG)<br />

protein. The present study was designed to evaluate the plasma<br />

levels of MIR17HG protein, an index of miR-17~92 cluster<br />

activity, and hepatic expression of its targets phosphatase and<br />

tensin homolog (PTEN) and nuclear factor kappa-B (NF-κB)<br />

in patients with chronic hepatitis C virus (HCV) infection in<br />

relation to hepatic steatosis, inflammation and fibrosis. Methods:<br />

Thirty treatment-naïve patients with chronic HCV infection<br />

[18 patients with chronic hepatitis C (CHC) and 12 patients<br />

with cirrhosis] and 15 healthy subjects were included in the<br />

study. Quantitative determination of plasma levels of MIR17HG<br />

protein was performed using quantitative sandwich enzyme<br />

immunoassay. Core liver biopsies obtained from patients with<br />

chronic HCV infection were assessed for METAVIR histological<br />

activity grade and fibrosis stage and steatosis grade. Immunohistochemical<br />

staining of liver specimens was done using<br />

monoclonal antibody against PTEN and NF-κB. Results: Plasma<br />

MIR17HG protein levels showed a significant increase in<br />

patients with chronic HCV infection compared with healthy subjects<br />

(P = 0.012) and in patients with cirrhosis compared with<br />

patients with CHC (P < 0.001). By plotting a receiver-operating<br />

characteristic curve, the sensitivity and specificity of plasma<br />

MIR17HG protein levels in discriminating patients with cirrhosis<br />

from patients with CHC were 100% and 88.9% respectively<br />

at a cut-off level of 45.5 pg/ml (AUC = 0.995). The plasma<br />

MIR17HG protein levels were inversely correlated with hepatic<br />

PTEN expression and positively correlated with serum levels of<br />

aminotransferases, METAVIR histological activity grade and<br />

fibrosis stage, steatosis grade and hepatic NF-κB expression<br />

(P < 0.05). The hepatic PTEN expression showed positive correlations<br />

with serum gamma glutamyl transpeptidase levels,<br />

METAVIR fibrosis stage and steatosis grade (P < 0.05) and an<br />

inverse correlation with hepatic NF-κB expression (P = 0.006).<br />

Conclusion: Activation of the miR-17~92 cluster may play an<br />

important role in the pathogenesis of HCV-related liver injury<br />

via PTEN inhibition and NF-κB activation and could be a possible<br />

therapeutic option in chronic HCV infection. Circulating<br />

MIR17HG protein could be a useful biomarker for disease<br />

progression.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hoda El Aggan, Sahah A. Mahmoud,<br />

Nevine M. El Deeb, Ehab M. Hassona, Sally El Demiry<br />

1012<br />

The FGL2:FcγRIIB Immunosuppressive Pathway Inhibits<br />

HCV Specific Antiviral T Cell Responses in Chronic Infection<br />

Ramzi Khattar, Kaveh Farrokhi, Hassan Sadozai, Vanessa Rojas<br />

Luengas, Gary Levy, Nazia Selzner; University Health Network,<br />

Toronto, ON, Canada<br />

Background and Aims: Hepatitis C virus (HCV) evades immune<br />

detection by limiting the magnitude and responsiveness of antiviral<br />

T cells. FGL2 is a potent immunosuppressive effector molecule<br />

of CD4 + CD25 + TIGIT + FOXP3 + Treg that exerts its activity<br />

by binding to the inhibitory FCγRIIB receptor expressed on antigen<br />

presenting cells. Binding of FGL2 to the FCγRIIB receptor<br />

inhibits the maturation of dendritic cells, limiting responsiveness<br />

of antiviral T cells. We examined the relationship between<br />

FGL2 and HCV specific T cell responses in a cohort of HCV<br />

patients. The study also examined the efficacy of a novel neutralizing<br />

antibody towards FGL2 to restore HCV specific T cell<br />

proliferation in vitro from HCV patients. Methods: 61 patients<br />

with chronic HCV infection, 15 SVR and 8 healthy controls<br />

were recruited to the study. TIGIT expression on CD4+CD25+-<br />

FOXP3+ Treg and FGL2 production by T cells, dendritic cells<br />

and macrophages was measured in patient PBMC by flow<br />

cytometry. FGL2 levels were measured in patient samples by<br />

ELISA and T cell reactivity to HCV by ex vivo peptide re-stimulation<br />

using genotype specific HCV peptide arrays. Results: In<br />

patients chronically infected with HCV, there was a significant<br />

increase in frequency of CD4 + CD25 + FOXP3 + TIGIT + Treg cells<br />

(42.5±4.1) compared to SVR patients (30.1±2.8) or healthy<br />

controls (14.5±1.5) (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 705A<br />

Disclosures:<br />

The following authors have nothing to disclose: Ramzi Khattar, Kaveh Farrokhi,<br />

Hassan Sadozai, Vanessa Rojas Luengas, Gary Levy, Nazia Selzner<br />

1013<br />

Novel effect of NS5A inhibition on Augmentation of<br />

HCV specific immunity and SVR<br />

Shikha Shrivastava 1 , Anu Osinusi 2 , Shyam Kottilil 1 ; 1 Institute of<br />

Human Virology, University of Maryland, Baltimore, Baltimore,<br />

MD; 2 Gilead Sciences Inc,, Foster City,, CA<br />

Background: We previously demonstrated that HCV clearance<br />

is associated with increased HCV specific immunity in<br />

chronic hepatitis C virus (HCV) genotype 1 (GT-1) infected<br />

patients during treatment with the Interferon (IFN) free regimen<br />

of sofosbuvir and ribavirin (SPARE). In this study, we<br />

evaluated the HCV specific immune responses in patients who<br />

relapsed previous treatment with sofosbuvir and ribavirin but<br />

were retreated with sofosbuvir and ledipasvir. Aim: To analyze<br />

the changes in HCV specific immunologic responses associated<br />

with viral clearance with combination DAA therapy of<br />

Sofosbuvir and Ledipasvir in chronic HCV GT1 patients who<br />

relapsed and lacked enhancement of HCV specific immune<br />

response when treated with sofosbuvir and ribavirin (SPARE).<br />

Methods: HCV GT-1 patients (N=14) that relapsed after treatment<br />

with sofosbuvir and ribavirin for 24 weeks were retreated<br />

with Sofosbuvir and Ledipasvir for 12 weeks. Phenotypic and<br />

functional changes within the adaptive and innate immune<br />

compartments of peripheral blood mononuclear cell (PBMC)<br />

at baseline and end of treatment (EOT) were analyzed. HCV<br />

specific immune responses were quantified by ELISA, ELISpot<br />

and intracellular flow cytometry (IFN-gamma, IL-2, TNF-a) after<br />

stimulation with overlapping peptides spanning the entire HCV<br />

genome. Results: In the SPARE study, SVR was associated with<br />

an enhancement of HCV-specific immune response (IFN-G positive<br />

peripheral CD8+ T cells). During the retreatment of these<br />

relapsers from SPARE with Sofosbuvir and Ledipasvir for 12<br />

weeks, all patients have attained SVR 12. Suppression of HCV<br />

was associated with decline in T cell exhaustion markers at EOT<br />

such as CD8+CD57+ (9.3 + 0.2 % vs. 6.2+ 0.6% p=0.02),<br />

CD8+Tim3+ (67+10.2 % VS. 51+9.3 % p=0.03), CD8+PD1+<br />

(17+1.8 % VS. 12+1.5% p=0.03). All patients were able to<br />

augment peripheral HCV-specific T cell IFN-gamma responses<br />

at the end of treatment compared to baseline when treated with<br />

sofosbuvir and ledipasvir (p=0.003). SVR was associated with<br />

augmentation of HCV-specific T cells response. Conclusions:<br />

Addition of NS5A inhibitor to sofosbuvir is associated with<br />

augmentation of HCV specific immunity and SVR in patients<br />

who previously failed sofosbuvir and ribavirin therapy. These<br />

findings demonstrate a novel effect of NS5A inhibitors in inducing<br />

host response aiding HCV clearance and achieving SVR.<br />

Disclosures:<br />

Anu Osinusi - Employment: gilead sciences<br />

The following authors have nothing to disclose: Shikha Shrivastava, Shyam Kottilil<br />

1014<br />

Differential microRNA expression in the liver during the<br />

advanced stage of chronic hepatitis C<br />

Rika Horii, Masao Honda, Takayoshi Shirasaki, Hikari Okada,<br />

Mikiko Nakamura, Shuichi Kaneko; Kanazawa University Graduate<br />

School of Medicine, Kanazawa, Japan<br />

BACKGROUND & AIMS: MicroRNAs (miRNAs) play important<br />

roles in development, metabolism, infection, and cancer.<br />

We previously analyzed the changes of miRNA expression<br />

associated with the progression of chronic hepatitis C (CHC)<br />

and showed that the expression of 7 miRNAs (miR-10a, -19a,<br />

-27a, -195, -199a, -214, and -218) were most significantly<br />

different between patients with early stage fibrosis (F1–2) and<br />

advanced stage fibrosis (F3–4). In this study, we evaluated<br />

the functional relevance of these miRNAs on Hepatitis C virus<br />

(HCV) replication and pathogenesis of liver disease. METH-<br />

ODS: MiRNAs were obtained from biopsy specimens of CHC<br />

patients infected with genotype 1b HCV. The functional relevance<br />

of miRNAs on HCV replication was evaluated in Huh-7<br />

cells using the infectious genotype 1a clone pH77S.3/Gluc2A<br />

with a Gaussia reporter gene. HCV translation (HCV-IRES) was<br />

monitored in the stably transformed IRES reporter cell line RCF-<br />

26. MiRNA-transfected Huh-7 cells were infected with HCV<br />

particles in oleic acid- or chenodeoxycholic acid (CDCA)–containing<br />

medium and HCV replication and cellular metabolism<br />

were evaluated. RESULTS: Among 7 miRNAs differentially<br />

expressed between early and advanced stage fibrosis (miR-<br />

10a, -19a, -27a, -195, -199a, -214, and -218), most miRNAs<br />

other than miR-19a were significantly up-regulated in Huh-7<br />

cells following HCV infection. Conversely, overexpression of<br />

miR-10a, -27a, -195, -199, and -214 significantly repressed<br />

HCV replication in Huh-7 cells, while miR-19a and -218 had<br />

no effect on HCV replication. We focused on miR-10a, as<br />

it was not previously characterized in HCV replication and<br />

liver pathogenesis. MiR-10a had no effect on HCV translation;<br />

however, its overexpression decreased hepatic triglyceride synthesis<br />

by suppressing lipid synthetic genes such as SREBP1 and<br />

FASN. Moreover, CDCA treatment in HCV replicating Huh-7<br />

cells increased hepatic concentrations of cholesterol but miR-<br />

10a decreased hepatic cholesterol content and suppressed<br />

HCV replication. We found that miR-10a targeted liver receptor<br />

homolog-1, a key regulatory enzyme of bile acid biosynthesis<br />

from cholesterol 7α-hydroxylase/CYP7A1. Thus, miR-10a<br />

suppresses the storage of triglyceride and cholesterol and the<br />

synthesis of bile acid, suppressing HCV replication. This would<br />

protect liver cell damage in the advanced fibrosis stage of liver<br />

disease. Conclusion: MiR-10a possibly plays an important role<br />

in cellular lipid and bile acid metabolism. The findings of this<br />

report might serve to establish a new therapeutic strategy for<br />

HCV-related advanced liver disease.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Rika Horii, Masao Honda, Takayoshi<br />

Shirasaki, Mikiko Nakamura<br />

1015<br />

The tumor suppressor IQGAP2 plays an essential role<br />

in antiviral activity of Interferon alpha against HCV<br />

through the NF-κB pathway.<br />

Cynthia Brisac, Shadi Salloum, Victor Yang, Stephane Chevaliez,<br />

Esperance A. Schaefer, Jian Hong, Charles Carlton-Smith, Nadia<br />

Alatrakchi, Anna Lidofsky, Dahlene Fusco, Xiao Liu, Dachuan Cai,<br />

Lee F. Peng, Wenyu Lin, Raymond T. Chung; MGH, Boston, MA<br />

Type I Interferon (IFN) has been the backbone of treatment for<br />

hepatitis C virus (HCV) infection over the past two decades. For<br />

some countries, it is likely to remain the only treatment since<br />

they will not be able to afford new high cost IFN-free regimen.<br />

Recently, using unbiased genome-wide siRNA screens, we<br />

identified hundreds of novel genes required for IFN response<br />

against HCV (Zhao H, J. Hepatology, 2013 and Fusco DN,


706A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Gastroenterology, 2014) shedding light on new pathways of<br />

IFN. It is now clear that JAK/STAT and the hundreds of Interferon<br />

Stimulated Genes (ISGs) it stimulates is not the only pathway<br />

mediating IFN’s antiviral effects in the liver. One of the<br />

genes we identified is IQ-motif containing GTPase activating<br />

protein 2 (IQGAP2). IQGAP2 is predominantly expressed in<br />

the liver and has been described as a tumor suppressor. Here<br />

we show that IQGAP2 also mediates IFN anti-HCV response<br />

in infected hepatoma cells. Indeed, IQGAP2 siRNA-mediated<br />

knock-down rescues HCV infection from IFN inhibition. We<br />

show that IQGAP2 is not itself an ISG and its knockdown<br />

does not affect JAK/STAT activation, suggesting that IQGAP2<br />

acts independently of the JAK/STAT pathway. We found that<br />

IQGAP2 interacts with RelA/p65, a subunit of NF-κB transcription<br />

factor. Interestingly, our data suggest that IFN alone<br />

is sufficient to activate RelA and stimulate NF-κB-dependent<br />

transcription in hepatoma cells; this process requires IQGAP2<br />

expression. Furthermore, we found that RelA silencing mimics<br />

IQGAP2 knock-down effects in rescuing HCV from IFN inhibition<br />

in IFN-deficient cells. This suggests that IQGAP2 and RelA<br />

mediate IFN anti-HCV response downstream of IFN binding to<br />

its receptor, a mechanism different from the well characterized<br />

role of NF-κB in IFN production. Finally, we investigated the<br />

effect of IQGAP2 or RelA knock-down on the induction by<br />

IFN of 23 ISGs with previously described anti-HCV properties.<br />

We found that IQGAP2 and RelA are both required for the<br />

full induction of two thirds of the tested genes suggesting that<br />

IQGAP2 and RelA mediate IFN anti-HCV response by controlling<br />

ISG induction in parallel of the JAK/STAT pathway.<br />

Altogether, our data demonstrate a previously unrecognized<br />

function for IQGAP2 in regulating innate antiviral response<br />

against HCV, and probably other hepatotropic pathogens.<br />

Since IQGAP2 is predominantly expressed in the liver, understanding<br />

the mechanism of RelA activation by IQGAP2 may<br />

open novel opportunities to prevent NF-kB associated liver disease.<br />

Disclosures:<br />

Stephane Chevaliez - Advisory Committees or Review Panels: Janssen; Speaking<br />

and Teaching: Gilead, BMS<br />

Dahlene Fusco - Grant/Research Support: Gilead<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Cynthia Brisac, Shadi Salloum,<br />

Victor Yang, Esperance A. Schaefer, Jian Hong, Charles Carlton-Smith, Nadia<br />

Alatrakchi, Anna Lidofsky, Xiao Liu, Dachuan Cai, Lee F. Peng, Wenyu Lin<br />

1016<br />

Single-strand RNA-induced monocyte differentiation<br />

generates pro-fibrogenic M2 macrophages in chronic<br />

hepatitis C Virus infection<br />

Banishree Saha, Karen Kodys, Gyongyi Szabo; Medicine, UMASS<br />

Med School, Worcester, MA<br />

Purpose: Innate immune responses, including monocyte and<br />

macrophage activation, contribute to the pathogenesis of<br />

chronic Hepatitis C Virus (HCV) infection and liver fibrosis.<br />

However the mechanisms that lead to liver fibrosis during<br />

HCV infection are not well understood. In response to the<br />

tissue microenvironment, monocytes undergo differentiation<br />

into polarized states including M1 (pro-inflammatory) or M2<br />

(anti-inflammatory/pro-fibrotic) macrophages. We hypothesized<br />

that during chronic HCV infection monocytes undergo<br />

differentiation that affect liver disease progression. Methods:<br />

Healthy human monocytes (n=10-15) were cultured with cellfree<br />

HCV or exosomes isolated from HCV-infected Huh7.5<br />

hepatoma cells for 5-7 days. Circulating monocytes and liver<br />

macrophages from HCV-infected patients and controls were<br />

analyzed by flow cytometry and western blotting. Results:<br />

Co-culture of healthy monocytes with cell-free HCV or exosomes<br />

derived from HCV-infected Huh7.5 cells resulted in differentiation<br />

of monocytes into macrophages (MΦ) with high CD14<br />

and CD68 expression. Exosomes from HCV-infected Huh7.5<br />

cells contained HCV single stranded (ss) RNA. These MΦ displayed<br />

M2 markers (CD206, CD163 and DC-SIGN) and produced<br />

pro- and anti-inflammatory cytokines. The HCV-induced<br />

M2-MΦ led to hepatic stellate cell (LX2) activation indicated<br />

by increased expression of collagen, TIMP-1 and α-SMA and<br />

this was prevented by anti-TGFβ blocking antibody administration.<br />

The HCV-induced monocyte activation and differentiation<br />

into M2-MΦ was prevented by TLR8 but not TLR3 or<br />

TLR7 knock-down. TLR8 ligands, independent of HCV, caused<br />

monocyte differentiation and M2-MΦ polarization suggesting<br />

a role for TLR8 in this process. Furthermore, HCV ssRNA alone<br />

induced monocyte to M2-MΦ differentiation and polarization,<br />

demonstrating that HCV ssRNA signals via TLR8 induce monocyte<br />

differentiation and polarization. In vivo in patients with<br />

chronic HCV infection, we observed a significant increase in<br />

the expression of M2 markers (CD206, CD163) on circulating<br />

monocytes and in the liver. Furthermore, we found a significant<br />

increase in collagen + CD206 + CD14 + circulating monocytes<br />

and the increased frequency of collagen + CD206 + CD14 + circulating<br />

monocytes correlated with liver fibrosis in HCV-infected<br />

patients. Conclusion: Our data identify a new mechanism by<br />

which cell-free and exosome-packaged HCV ssRNA interacts<br />

with monocytes and induces TLR8-mediated differentiation to a<br />

pro-fibrogenic, M2-MΦ. These M2-MΦ promote liver fibrosis.<br />

We also identified M2-marker plus collagen-expressing circulating<br />

monocytes as potential biomarkers of fibrosis in chronic<br />

HCV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Banishree Saha, Karen Kodys,<br />

Gyongyi Szabo<br />

1017<br />

Elucidating Mechanisms Underlying Development of<br />

Liver Disease in HIV/HCV Coinfection<br />

Jinhee Hyun 3 , Masato Yoneda 1 , Eugene R. Schiff 1 , Emmanuel<br />

Thomas 1,2 ; 1 Schiff Center for Liver Diseases, University of Miami<br />

Miller School of Medicine, Miami, FL; 2 Sylvester Comprehensive<br />

Cancer Center, Miami, FL; 3 Cell Biology, University of Miami<br />

School of Medicine, Miami, FL<br />

BACKGROUND: Patients with HIV that are coinfected with<br />

HCV are at increased risk for rapidly progressive liver disease<br />

and subsequently the development of Hepatocellular Carcinoma<br />

(HCC). Specifically, HCC develops earlier in coinfected<br />

patients and these patients are more symptomatic than those<br />

with only HCV infection at diagnosis suggesting that both viruses<br />

increase the propensity for malignant transformation. However,<br />

the genetic and cellular based mechanisms underpinning how<br />

HCV initiates and subsequently induces liver pathology and<br />

why coinfection with HIV results in significantly worse hepatic<br />

disease remains to be clarified. In addition, the specific cell<br />

types that contribute to these clinical outcomes are unknown.<br />

In this study, we specifically are testing the hypothesis that the<br />

robust viral RNA dependent innate immune response that we<br />

have characterized, that drives inflammation in HCV infected<br />

livers, is augmented with HIV coinfection through activation<br />

of viral DNA and RNA sensing pathways. METHODS: To this<br />

end, we have developed novel in-vitro models that utilize HCV<br />

infected primary human hepatocytes (PHHs) and HIV infected<br />

primary kupffer cells (PKCs). Specifically, the JFH-1 strain was<br />

used to study HCV while the BAL (R5) and IIIB (X4) strains were


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 707A<br />

used for HIV infection. A coculture model of PHHs an PKCs was<br />

also utilized. Specific RNA and DNA dependent innate and<br />

inflammatory pathways, such as RIG-I/TLR3 and IFI16/cGAS<br />

signaling, were studied as well as additional pathways involving<br />

the production of interferon and chemokines such as IP10/<br />

CXCL10, IL-1b and CCL5/Rantes that may drive subsequent<br />

liver disease. RESULTS: We found that similar to distinct stimuli<br />

including IFN-a and polyI:C, HCV and HIV stimulation of these<br />

primary liver cells resulted in rapid and robust upregulation of<br />

IP10 and other inflammatory cytokines at the mRNA and protein<br />

level. Surprisingly, microarray analyses of HIV treated liver<br />

cells demonstrated several innate immune pathways were activated<br />

by this virus in the liver. siRNA experiments demonstrated<br />

the induction of inflammatory cytokines was mediated by IRF3<br />

and NFkB-dependent pathways. CONCLUSIONS: Overall,<br />

our data demonstrate that the robust antiviral response that is<br />

observed in HCV infected livers and subsequently drives liver<br />

pathology is augmented with HIV/HCV coinfection. It is hoped<br />

that understanding the mechanism by which coinfected patients<br />

have more severe liver disease will enable the development of<br />

targeted therapeutics to abrogate these poor clinical outcomes.<br />

Disclosures:<br />

Eugene R. Schiff - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead, Merck, Janssen, Salix Pharmaceutical, Pfizer, Arrowhead; Consulting:<br />

Acorda; Grant/Research Support: Bristol Myers Squibb, Abbott / AbbVee, Gilead,<br />

Merck, Conatus, Medmira, Roche, Janssen, Orasure Technologies, Discovery<br />

Life Sciences, Siemens, Beckman Coulter, Siemens<br />

The following authors have nothing to disclose: Jinhee Hyun, Masato Yoneda,<br />

Emmanuel Thomas<br />

1018<br />

Expression of IFNλ4 in liver is closely associated with<br />

non-response to antiviral therapy through the regulation<br />

of basal expression of ISGs in chronic hepatitis C<br />

patients but not in hepatitis B patients.<br />

Miyako Murakawa 1,3 , Yasuhiro Asahina 1 , Fukiko Kawai-Kitahata<br />

1 , Hiroko Nagata 1 , Syun Kaneko 1 , Sayuri Nitta 1 , Takako<br />

Watanabe 1 , Yasuhiro Itsui 1,3 , Mina Nakagawa 1 , Sei Kakinuma 1 ,<br />

Sayuki Iijima 2 , Yasuhito Tanaka 2 , Mamoru Watanabe 1 , Yujiro<br />

Tanaka 3,1 ; 1 Gastroenterology and Hepatology, Tokyo Medical<br />

and Dental University, Tokyo, Japan; 2 Nagoya City University,<br />

Nagoya, Japan; 3 Medical Education Research and Development,<br />

Tokyo Medical and Dental University, Tokyo, Japan<br />

Backgrounds & Aims: The innate immune system has an essential<br />

role in host antiviral defense. IL28B SNPs are associated<br />

with the response to anti-HCV therapy both in IFN-based and<br />

IFN-free regimens. Recently, IFNλ4 gene was considered as<br />

one of the mechanisms responsible for poor response influenced<br />

by IL28B unfavorable SNPs. However, little is known<br />

about IFNλ4 expressions and its roles in antiviral innate immunity<br />

including intrahepatic gene expressions of IFN-stimulated<br />

genes (ISGs). The aim of this study is to evaluate the relationship<br />

between IFNλ4 expression and viral liver diseases, and to<br />

investigate the effect of IFNλ4 on intrahepatic gene expressions<br />

responsible for poor response to IFN therapy. Methods: 1)<br />

IFNλ4 expressions were analyzed in HCVcc-infected Huh7.5.1<br />

cells, poly (I:C)-treated Huh7, HepG2, HeLa, HEK293T cells<br />

and B lymphocytes, and HBV-expressing HepG2.215 cells.<br />

2) The expressions of IFNλ4 were analyzed in liver samples<br />

obtained from 49 chronic hepatitis C (CHC), 13 chronic hepatitis<br />

B (CHB) and 3 autoimmune hepatitis (AIH) patients. All<br />

CHC patients were treated with PEG-IFN/ribavirin (PR) therapy<br />

and the relationship between IFNλ4 expressions and<br />

treatment responses. 3) Antiviral ISGs (ISG15, Mx1, RIG-I,<br />

TLR3, SOCS2) and IFN-suppressive genes (RNF125, SOCS1,<br />

SOCS3, A20, RNF11) were analyzed in all liver samples,<br />

and the relationship with IFNλ4 expressions were accessed.<br />

Results: 1) IFNλ4 expression was induced by poly (I:C) treatment<br />

as well as HCVcc infection in IL28B-minor cell lines, but<br />

not induced by HBV. 2) IFNλ4 mRNA was detected in 11 of<br />

22 liver samples of CHC patients with IL28B-unfavorable SNP,<br />

but not in CHC patients with favorable genotype, CHB and AIH<br />

patients. IFNλ4 expression was associated with non-response<br />

to PR therapy (p < 0.001). 3) Intrahepatic expressions of antiviral<br />

ISGs (ISG15 and Mx1) were significantly up-regulated in<br />

nonvirological responders (NR) compared with those in virological<br />

responders (VR) while expressions of suppressive ISGs<br />

(RNF125, SOCS1, SOCS3 and RNF11) were down-regulated<br />

in NR. Expressions of antiviral ISGs (Mx1 and RIG-I) were significantly<br />

higher in IL28B-unfavorable patients in whom IFNλ4<br />

mRNA was detected compared with those in IFNλ4-undetected<br />

patients. However, expressions of suppressive ISGs (RNF125,<br />

SOCS1, SOCS3 and RNF11) were significantly lower in<br />

IFNλ4-detected patients. Conclusions: Our data suggest IFNλ4<br />

have a specific role only in CHC. Intrahepatic expression of<br />

IFNλ4 is associated with higher expression of antiviral ISGs<br />

and lower expressions of suppressive ISGs at baseline, resulting<br />

in the poor responsiveness for IFNα-based therapy for HCV<br />

infection.<br />

Disclosures:<br />

Sei Kakinuma - Grant/Research Support: The Japanese Society of Gastroenterology,<br />

MSD Co., Ltd.<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

Mamoru Watanabe - Grant/Research Support: MSD Co., Ltd.<br />

The following authors have nothing to disclose: Miyako Murakawa, Yasuhiro<br />

Asahina, Fukiko Kawai-Kitahata, Hiroko Nagata, Syun Kaneko, Sayuri Nitta,<br />

Takako Watanabe, Yasuhiro Itsui, Mina Nakagawa, Sayuki Iijima, Yujiro Tanaka<br />

1019<br />

Lipid-Free apolipoprotein E signaling induces an ABCG1<br />

driven cholesterol efflux that markedly decreases HCV<br />

replication<br />

Emilie Crouchet 1,2 , Mathieu Lefèvre 1,3 , Eloi R. Verrier 1,2 , Thomas<br />

F. Baumert 1,2 , Catherine Schuster 1,2 ; 1 U1110, Inserm, Strasbourg,<br />

France; 2 Strasbourg University Hospital, Université de Strasbourg,<br />

Strasbourg, France; 3 CNRS UPR 9002, Strasbourg, France<br />

There is an intimate link between the hepatitis C virus (HCV)<br />

life cycle and its host cell’s lipid metabolism. HCV is associated<br />

with lipoproteins in blood, forming lipo-viro-particles, which<br />

are the infectious form of the virus. Apolipoprotein E (apoE), a<br />

key lipo-viro-particle component, plays an essential role in HCV<br />

entry, assembly, and egress. Although both the lipoprotein-associated<br />

and lipid-free forms of apoE can be found in blood,<br />

the role of lipid-free apoE in the HCV life cycle is unknown. In<br />

this study, we investigated the effect of lipid-free apoE on the<br />

HCV life cycle using the HCVcc model system and primary<br />

human hepatocytes (PHH). A dose-dependent decrease in HCV<br />

replication was observed when Huh 7.5.1 cells or PHHs were<br />

treated with increasing amounts of lipid-free apoE. We show<br />

that lipid-free apoE acts on HCV replication independently of<br />

previously described apoE receptors, and independently of<br />

the lipid-free apoE recycling process. By investigating the role<br />

of lipid-free apoE on actors of hepatic lipid metabolism we<br />

observed that extracellular lipid-free apoE markedly increases<br />

cholesterol efflux via the ATP binding cassette sub-family G<br />

member 1 protein (ABCG1). Our findings highlight a new<br />

mechanism in lipid metabolism regulation and may provide<br />

new therapeutic strategies to fight chronic HCV infection.


708A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Emilie Crouchet, Mathieu Lefèvre,<br />

Eloi R. Verrier, Thomas F. Baumert, Catherine Schuster<br />

1020<br />

Substitution in amino acid 70 of hepatitis C virus core<br />

protein changes the adipokine profile via toll-like receptor<br />

2/4 signaling<br />

Masahiko Tameda, Kazushi Sugimoto, Yoshiyuki Takei; Mie University,<br />

Tsu, Japan<br />

Background & Aims: It has been suggested that amino acid<br />

(aa) substitution at position 70 from arginine (70R) to glutamine<br />

(70Q) in the genotype 1b hepatitis C virus (HCV) core protein<br />

is associated with insulin resistance and worse prognosis.<br />

However, the precise mechanism is still unclear. The aim of<br />

this study was to investigate the impact of the substitution at<br />

position 70 in HCV core protein on adipokine production by<br />

murine and human adipocytes. Methods: The influence of treatment<br />

with HCV core protein (70R or 70Q) on adipokine production<br />

by both 3T3-L1 and human adipocytes were examined<br />

with real-time PCR and enzyme-linked immunosorbent assay<br />

(ELISA), and triglyceride content was also analyzed. The effects<br />

of toll-like receptor (TLR)2/4 inhibition on IL-6 production by<br />

3T3-L1 induced by HCV core protein were examined. Results:<br />

IL-6 production was significantly increased and adiponectin<br />

production was reduced without a change in triglyceride content<br />

by treatment with 70Q compared to 70R core protein in<br />

both murine and human adipocytes. IL-6 induction of 3T3-L1<br />

cells treated by 70Q HCV core protein was significantly inhibited<br />

with anti-TLR2 antibody by 42%, and by TLR4 inhibitor<br />

by 40%. Conclusions: Our study suggests that extracellular<br />

HCV core protein with substitution at position 70 enhanced IL-6<br />

production and reduced adiponectin production from visceral<br />

adipose tissue, which can cause insulin resistance, hepatic<br />

steatosis, and ultimately development of HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Masahiko Tameda, Kazushi<br />

Sugimoto, Yoshiyuki Takei<br />

1021<br />

Variability of the hepatitis delta virus quasispecies in<br />

chronic infection is high and similar to that of the hepatitis<br />

C virus quasispecies<br />

Maria Homs 1,2 , Josep Gregori 3 , Hadi Karimzadeh 4 , Damir G.<br />

Cehic 3 , Maria Blasi 1,2 , Rosario Casillas 3 , David Tabernero 1,2 ,<br />

Josep Quer 1,3 , Michael Roggendorf 4 , Mar Riveiro-Barciela 5 , Rafael<br />

Esteban 1,5 , Francisco Rodriguez-Frias 1,2 , Maria Buti 1,5 ; 1 Centro de<br />

Investigacion Biomedica en Red de Enfermedades Hepaticas y<br />

Digestivas, CIBERehd, Barcelona, Spain; 2 Microbiology, Hospital<br />

Vall d’Hebron, Barcelona, Spain; 3 Liver Diseases, Research Institute<br />

Vall d’Hebron, Barcelona, Spain; 4 Institut für Virologie, Technische<br />

Universität München, München, Germany; 5 Hepatology,<br />

Hospital Vall d’Hebron, Barcelona, Spain<br />

Background Hepatitis D virus (HDV) is a satellite RNA virus with<br />

a small 1.7-kb genome that uses the hepatitis B virus envelope<br />

to infect hepatocytes. The complexity of the hepatitis D viral<br />

population has not been described. Hepatitis C virus (HCV)<br />

is another RNA virus that circulates in infected individuals as<br />

a quasispecies; the NS5A region is one of the most complex<br />

of the non-structural encoding HCV genes. Aim To study the<br />

complexity of the HDV quasispecies in patients with chronic<br />

hepatitis D and compare it with that of the HCV quasispecies.<br />

Samples and methods Fifteen serum samples from 10<br />

chronic HCV- and 5 chronic HDV-infected treatment-naïve<br />

patients were analyzed. Samples with similar viremia levels<br />

were selected (mean HCV-RNA 6.51 log copies/mL, SD 0.57<br />

and HDV-RNA 6.31 log copies/mL, SD 0.61 p=0.84). HCV-<br />

RNA was quantified by COBAS-AmpliPrep (Roche) (detection<br />

limits 1.63-7.83 log IU/mL) and HDV-RNA, by an in-house<br />

method (detection limits 3.5-8.5 log copies/mL). HCV-RNA<br />

values were converted to log copies/mL. One essential region<br />

in each virus was deep-sequenced: the NS5A of HCV (positions<br />

6299-6735, 436 bp) and the C-terminal region of HDV<br />

antigen (positions 912-1298, 386 bp). Abundance filters for<br />

deep-sequencing were set at 0.25%. Quasispecies complexity<br />

was evaluated by determining the mutation frequency (Mf)<br />

and nucleotide diversity (Pi). Both parameters take into account<br />

amplicon length, and represent the number of differences/site<br />

with respect to the most represented haplotye (Mf) or between<br />

each pair of genomes (Pi). Results are presented as mean and<br />

standard deviation (SD). Results A total of 187,606 sequences<br />

were analyzed. Mf and Pi showed a normal distribution and<br />

significantly correlated (R=0.985, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 709A<br />

gene. Cellular and viral protein expression was evaluated by<br />

western blot using antibodies vs. HCV-NS5A, SOD1, SOD2,<br />

catalase, thioredoxin-1, MAT1, MAT2, PKR, STAT1 and actin.<br />

Results: SAM treatment decreased HCV-RNA levels 50-70%<br />

compared to untreated control (24-72h). Total glutathione levels<br />

increased in both cell lines about 50-60% compared to<br />

control without SAM (6h post-treatment in replicon cells and 2h<br />

post-treatment in parental cells). Transcriptional expression of<br />

SOD1, SOD2 and thioredoxin-1 was increased (0.5, 2.5 and<br />

2 fold-times respectively compared to control) at different time<br />

points (24-72h). This effect was not observed for catalase. Interestingly,<br />

there was no significant change in ROS levels in both<br />

cell types upon SAM treatment at all times assessed, contrary<br />

to the observed with PDTC exposition where an average of<br />

30% reduction was detected (0.5-24h). In other hand, MAT1<br />

expression was increased (2.5 fold-times at 48h) and MAT2<br />

was decreased upon SAM exposition (24h) compared with<br />

untreated cells at both transcriptional and translational level<br />

in both cell lines. SAM treatment does not modify STAT1 and<br />

PKR expression compared to untreated cells (24-72h), however<br />

both protein levels were increased upon IFN-a treatment used<br />

as a control. Conclusions: A possible mechanism by which<br />

SAM decreases HCV expression could involve modulating<br />

antioxidant enzymes systems, biosynthesis of glutathione and<br />

switching MAT2/MAT1 turnover in Hepatitis C virus expressing<br />

cells. Proteins PKR and STAT1 were stimulated in the presence<br />

of IFN-a but not with SAM, suggesting that HCV down-regulation<br />

mediated by SAM is independent of IFN-a pathway. This<br />

may have clinical application in terms of understanding the<br />

pathophysiology of the disease. This work was supported by<br />

CONACYT-BASICA CB-2010-01155082.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sonia A. Lozano-Sepulveda, Eduardo<br />

Bautista-Osorio, Paula Cordero, Linda E. Muñoz, Ana Maria G. Rivas-Estilla<br />

1023<br />

Pre-treatment levels of miR-29b are associated with<br />

response to treatment and stage of fibrosis in patients<br />

with chronic hepatitis C [CHC]<br />

Hossein Sendi 1 , Marjan Mehrab-Mohseni 2 , Mark W. Russo 2 ,<br />

Philippe J. Zamor 2 , Paul A. Schmeltzer 2 , Nury Steuerwald 2 , Judith<br />

Parsons 2 , Mark G. Clemens 4 , Keith Kaplan 3 , W. Carl Jacobs 3 ,<br />

William A. Ahrens 3 , Herbert L. Bonkovsky 1 ; 1 Gastroenterology,<br />

Wake Forest Baptist Medical Center, Winston Salem, NC; 2 Liver,<br />

Carolinas Medical Center, Charlotte, NC; 3 Pathology, Carolinas<br />

Medical Center, Charlotte, NC; 4 Biology, UNC Charlotte, Charlotte,<br />

NC<br />

Both responses to treatment and long-term outcomes of hepatitis<br />

C virus (HCV) infection are critically affected by host genetic<br />

factors. We aimed to determine whether pre-treatment levels<br />

of miR122 or miR-29b play any role in response to treatment<br />

or stage of fibrosis in patients with CHC. A total of 25 CHC<br />

patients (HCV genotype 1) were included in this study, among<br />

whom 11 were treated with IFN plus Ribavirin (with or without<br />

Telaprevir) and 14 remained treatment naïve. The patients<br />

were classified according to their virological responses: early<br />

viral response (EVR): patients whose HCV RNA levels at 12<br />

weeks were reduced by 2-log 10<br />

or more compared to baseline.<br />

Non-EVR (n-EVR): patients whose HCV RNA levels were not<br />

reduced or reduced less than 2- log 10<br />

. Using quantitative real<br />

time PCR, we compared pre-treatment levels of hepatic miR-<br />

122, and miR-29b (normalized to SNORD44) in the patients<br />

based on their response to treatment. We found that pre-treatment<br />

levels of miR-122 were slightly lower in patients with EVR<br />

than those with n-EVR (49 vs 59, p>0.05). However, pre-treatment<br />

levels of miR-29b were significantly lower in patients<br />

with EVR than those with n-EVR (0.015 vs 0.130, p


710A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

duced HCV (JFH-1; genotype 2a). IRF9, STAT1, and STAT2<br />

were overexpressed by lentiviral transduction, or their expression<br />

was silenced with siRNAs. The expression of the negative<br />

regulators of type I IFN signaling, USP18 and ISG15, was<br />

studied in HCV-infected liver and IFN-λ 3<br />

treated hepatoma<br />

cells, and their regulation by U-ISGF3 was analyzed. Results<br />

The level of U-ISGF3, but not tyrosine phosphorylated STAT1,<br />

is significantly elevated in response to IFN-λ and IFN-β during<br />

chronic HCV infection. U-ISGF3 prolongs the expression of a<br />

subset of ISGs and restricts HCV chronic replication. However,<br />

paradoxically, high levels of U-ISGF3 also conferred unresponsiveness<br />

to IFN-α therapy. As a mechanism of U-ISGF3-induced<br />

resistance to IFN-α, we found that ISG15, a U-ISGF3-induced<br />

protein, sustains the abundance of USP18. Silencing ISG15<br />

decreased the level of USP18 in IFN-λ 3<br />

treated hepatoma cells<br />

and restored USP18-mediated attenuation of STATs phosphorylation<br />

after treatment of IFN-α. Conclusions Our data demonstrate<br />

that U-ISGF3 induced by IFN-λs and -β drives prolonged<br />

expression of a set of ISGs, leading to chronic activation of<br />

innate responses and conferring a lack of response to IFN-α in<br />

HCV-infected liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pil Soo Sung, HyeonJoo Cheon,<br />

Do Youn Park, Hyung-Il Seo, Su-Hyung Park, Seung Kew Yoon, George R Stark,<br />

Eui-Cheol Shin<br />

1026<br />

The expression of Immune-miRs could contribute to the<br />

immunopathogenesis of HCV and be modified by 1(OH)<br />

vitamin D3 supplementation<br />

Yasuteru Kondo 1 , Tatsuki Morosawa 1 , Masashi Ninomiya 1 ,<br />

Yasuyuki Fujisaka 1 , Yasuhito Tanaka 2 , Takayuki Kogure 1 , Jun<br />

Inoue 1 , Teruyuki Umetsu 1 , Tooru Shimosegawa 1 ; 1 Gastroenterology,<br />

Tohoku University Hospital, Sendai, Japan; 2 Nagoya City<br />

University, Nagoya, Japan<br />

[Background] We reported that the expression profiles of serum<br />

miRNAs could be important biomarkers for the diagnosis of<br />

various liver diseases (PLoS One 2013). It has been reported<br />

that various kinds of miRNAs known as immune-miRs could contribute<br />

to the immune regulation. [Aim] The aim of this study is<br />

to analyze the biological significance of immune-miRs expression<br />

on specific immune cells in chronic hepatitis C (CH-C)<br />

patients. [Material and Methods] Patients: Permission for the<br />

study was obtained from the ethics committee at our institute<br />

(2013-1-268). Twenty CH-C patients, 20 chronic hepatitis B<br />

patients (CH-B), and 10 healthy subjects were examined by<br />

deep sequencing analysis (Illumina G2X). Moreover, 50 CH-C<br />

patients (genotype 1b and high viral load) and control patients<br />

(CH-B, NASH, and healthy subjects) were examined by validation<br />

analysis. Deep sequencing analysis: Illumina deep<br />

sequencing for the initial screening to determine the read numbers<br />

of miRNAs expression in serum and PBMCs was carried<br />

out. Immune cells isolation: CD3 + T cells, CD14 + monocytes,<br />

CD19 + B cells and CD56 + NK cells were isolated using the<br />

magnetic beads method. Transfection of HCV individual protein<br />

expressing plasmids: The HCV individual expression plasmids<br />

(HCV-E1, E2, Core, NS3, NS4 and NS5A) were transfected<br />

into immune cells to identify the HCV proteins responsible for the<br />

suppression of miR146b-5p. Silencing and forced expression of<br />

miR146b-5p: Transfections of miRNA mimic (#MC10105) and<br />

miRNA inhibitor into THP-1, Jurkat, Molt-4, Raji and human<br />

primary lymphoid cells using 4D-nucleofector TM were carried<br />

out. [Results] The expressions of five miRNAs (miR146b-5p,<br />

miR-181a-2-3p, miR-204-5p, miR374a-3p and miR374a-5p)<br />

in the serum were significantly lower than those of the control<br />

groups. Among them, miR146b-5p was abundantly expressed<br />

in PBMCs. The expression of miR146b-5p in CD3 + T cells and<br />

CD14 + monocytes in CH-C patients was significantly lower<br />

than those in control groups. The silencing of miR146b-5p in<br />

the monocytes could significantly enhance the expression of<br />

CXCL10, TGF-β and IL10 under the stimulation (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 711A<br />

a treatment history with PEG-IFN/RBV (10 were prior null<br />

responders and 11 were prior relapsers). The SVR rate was<br />

86 % (6/7 patients) in treatment naïve patients, 82 % (9/11<br />

patients) in prior relapsers, and 40 % (4/10 patients) in prior<br />

null responders. the NS3 and the NS5A RAVs at baseline were<br />

detected in 71% (20/28 patients) and 81% (17/21 patients)<br />

by ultra-deep sequencing. No difference was found in the<br />

number of the patients with the NS3 RAVs before treatment<br />

between SVRs and non-SVRs (85% [11/19] vs. 71% [5/9],<br />

p=0.91). However, at the early phase of the treatment, increasing<br />

prevalence of the RAVs was more frequently found in non-<br />

SVRs more frequently than in SVRs (78 % [7/9], 32 % [6/19],<br />

p2 and p2 and an unadjusted p-value of


712A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 713A<br />

suggesting that HCV-ICs persist in an even higher proportion<br />

of patients. Studies are ongoing to assess the clinical implications<br />

of persistent HCV-specific ICs. CONCLUSIONS. We<br />

demonstrated for the first time that HCV-ICs persist in a majority<br />

of patients who had spontaneously cleared HCV or achieved<br />

SVR months to years earlier. Evidence of ongoing presence of<br />

HCV proteins after HCV clearance extends our understanding<br />

of pathogenesis. When using HCV-Ag assays, serum sample<br />

denaturation may result in failure to differentiate active from<br />

resolved HCV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ke-Qin Hu, Wei Cui<br />

1032<br />

Very low prevalence of Hepatitis C in HIV negative<br />

men who have sex with men in Tel Aviv: A prospective<br />

cohort analysis<br />

Ella Veitsman 1,7 , Adir Yanko 2 , Yuval Livnat 3,4 , Margalit Lorber 5,6 ,<br />

Ziv Ben Ari 1,7 ; 1 center for liver diseases, sheba medical center,<br />

Ramat Gan, Israel; 2 Israel LGBT organization, Levinsky clinic, Tel<br />

Aviv, Israel; 3 Israel AIDS Task force, Tel Aviv, Israel; 4 Tel Aviv University,<br />

Tel Aviv, Israel; 5 Immunology Department, Rambam Health<br />

Care Campus, Haifa, Israel; 6 The Ruth and Bruce Rappaport<br />

School of medicine, Israel institute of Technology, Haifa, Israel;<br />

7 Sackler School of medicine, Tel Aviv University, Tel Aviv, Israel<br />

Background: In light of highly effective existing and upcoming<br />

antiviral treatments, it is necessary to define additional risk<br />

groups of Hepatitis C virus (HCV) transmission. The occurrence<br />

of sexual transmission of HCV in human immunodeficiency<br />

virus (HIV) positive men who have sex with men (MSM) is well<br />

described and on the rise all over the world. Outbreaks of hepatitis<br />

C in this population have been linked to several risk factors<br />

including unprotected anal sex, use of recreational drugs,<br />

and concomitant sexually transmitted infections. This has led<br />

some experts to suggest that HIV negative MSM also have a<br />

high rate of hepatitis C but are simply not getting tested for it.<br />

However epidemiological <strong>studies</strong> on possible sexual transmission<br />

of hepatitis C in HIV negative MSM have shown significantly<br />

variable results in different countries. Aim: To determine<br />

the prevalence of positivity for antibodies to hepatitis C virus<br />

(anti-HCV) and the potential for sexual transmission of the<br />

virus in HIV negative MSM. Patients and methods: Participants<br />

were prospectively recruited from a day-MSM clinic in Tel Aviv<br />

during the period from July 2014 to April 2015. Participants<br />

self-completed a short questionnaire providing demographic<br />

and behavioral data. They also donated a sample of blood<br />

that was subsequently tested for antibodies to HCV and other<br />

selected pathogens (HIV, syphilis, gonorrhea and chlamydia).<br />

Results: In total 730 MSM HIV negative participants, median<br />

age 28.0 (range 16-64) completed the questionnaire and<br />

underwent HCV testing. The majority of the tested population<br />

was born in Israel. 85 participants (11.87%) were immigrants<br />

from other countries. Overall three of 730 tested were found<br />

sero-positive for antibodies to HCV. All three were born in<br />

Israel. The overall prevalence was 0.41% (95% CI: 0.21-<br />

0.8%). Further HCV RNA testing was negative in all three anti<br />

HCV positive men. Therefore no one was found harboring<br />

a replicating virus. Also no HCV sero-conversion was found<br />

during the study period. None of the three anti HCV positive<br />

participants reported known risk factors for HCV contraction<br />

except for unprotected sex in one of them. Conclusion: Our<br />

findings suggest that in Israel Hepatitis C among MSM without<br />

diagnosed HIV is much less prevalent than in the general<br />

population. There is no evidence of elevated rates of sexual<br />

transmission of HCV among MSM without HIV-infection.<br />

Disclosures:<br />

Yuval Livnat - Grant/Research Support: Abbvie, Neopharm Israel, Janssen<br />

The following authors have nothing to disclose: Ella Veitsman, Adir Yanko, Margalit<br />

Lorber, Ziv Ben Ari<br />

1033<br />

The Antibody response to hepatitis B virus vaccination<br />

is not influenced by the hepatitis C virus viral load in<br />

patients with chronic hepatitis C<br />

Roseane P. Medeiros, Marta Lopes, Daniel F. Mazo, Claudia P.<br />

Oliveira, Patricia M. Zitteli, João Renato R. Pinho, Flair J. Carrilho,<br />

Mario G. Pessoa; Gastroenterology, University of Sao Paulo<br />

School of Medicine, Sao Paulo, Brazil<br />

Background: Some immunogenicity <strong>studies</strong> of anti-HBV vaccine<br />

in patients with chronic HCV infection have demonstrated a<br />

diminished response ranging from 63.6% to 72.9% on seroconversion<br />

rate, compared to 90.9% to 93.9% in healthy controls.<br />

One possible explanation is the high HCV viral load in<br />

some patients. Aims: To evaluate the impact of HCV viral load<br />

on anti-HBV vaccination response in treatment naive chronic<br />

HCV patients without cirrhosis. Methods: 110 chronic HCV<br />

adult patients without cirrhosis were randomized to receive<br />

anti-HBV vaccination regimen at standard 3 doses (0, 1 and<br />

6 months) of 20ug, or higher dose of 40ug. Response to vaccination<br />

was measured by titers of anti-HBs 1 and 6 month<br />

after the last dose of anti-HBV vaccine. Healthy controls were<br />

negative to anti-HCV, anti-HBc, HBsAg and anti-HBs antibodies,<br />

and received standard 3 doses of 20ug at intervals of 0,<br />

1 and 6 months, and were also evaluated for anti-HBs titers.<br />

Results: Of the 110 HCV vaccinated patients, the seroconversion<br />

rate (anti-HBs ≥ 10IU/mL) was 74.5% (82/110). Out of<br />

the 45 healthy controls vaccinated with standard dose, we<br />

observed a seroconversion rate of 93.3% (42/45). Variables<br />

included in the logistic regression model were: Age, Gender,<br />

HCVRNA and Liver Fibrosis by Metavir (F0/F1 versus F2/F3).<br />

Age was the only variable that negatively influenced anti-HBs<br />

titers (p=0.003) Conclusions: Our study had demonstrated that<br />

chronic HCV patients without cirrhosis presented impaired anti-<br />

HBV vaccine response compared to healthy controls, similar to<br />

data previously demonstrated in the literature. This impairment<br />

is apparently not influenced by HCV viral load.<br />

Disclosures:<br />

João Renato R. Pinho - Employment: Hospital Israelita Albert Einstein<br />

The following authors have nothing to disclose: Roseane P. Medeiros, Marta<br />

Lopes, Daniel F. Mazo, Claudia P. Oliveira, Patricia M. Zitteli, Flair J. Carrilho,<br />

Mario G. Pessoa


714A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1034<br />

Approved All-oral Sofosbuvir Regimens are Safe and<br />

Highly Effective in Patients with Hereditary Bleeding<br />

Disorders<br />

Christopher Walsh 1 , Kimberly Workowski 2 , Norah Terrault 3 , Paul<br />

Sax 4 , Alice Cohen 5 , Christopher L. Bowlus 6 , Arthur Y. Kim 7 , Robert<br />

H. Hyland 8 , Jing Wang 8 , Luisa M. Stamm 8 , Diana M. Brainard 8 ,<br />

John G. McHutchison 8 , Annette Von Drygalski 9 , Frank S. Rhame 10 ,<br />

Michael W. Fried 11 , Peter Kouides 12 , Gayle Balba 13 , K. Rajender<br />

Reddy 14 ; 1 Mount Sinai Hospital, New York, NY; 2 Emory University,<br />

Atlanta, GA; 3 University of California, San Francisco, San<br />

Francisco, CA; 4 Brigham and Women’s Hospital, Boston, MA;<br />

5 Newark Beth Israel Center - Barnabas Health, Newark, NJ; 6 University<br />

of California, Davis, Sacremento, CA; 7 Massachusetts General<br />

Hospital, Boston, MA; 8 Gilead Sciences, Inc, Foster City, CA;<br />

9 University of California, San Diego, San Diego, CA; 10 Abbott<br />

NW Infectious Disease Clinic, Minneapolis, MN; 11 University of<br />

North Carolina at Chapel Hill, Chapel Hill, NC; 12 The Mary M<br />

Gooley Hemophilia Center, Rochester, NY; 13 Georgetown University<br />

Hospital, Washington, DC; 14 University of Pennsylvania,<br />

Philadelphia, PA<br />

Background: Patients with bleeding disorders have high rates<br />

of Hepatitis C Virus (HCV) infection due to exposure to contaminated<br />

blood products prior to 1992. They have historically<br />

been excluded from HCV clinical trials due to concern for<br />

unique adverse events, comorbidities, and/or their underlying<br />

diagnosis. This study evaluated the safety and efficacy of ledipasvir/sofosbuvir<br />

(LDV/SOF, genotype [GT] 1 and 4 infection)<br />

and SOF+RBV (GT 2 and 3 infection) in patients with bleeding<br />

disorders with or without HIV coinfection. Methods: GT1 or<br />

4 HCV-infected patients received LDV/SOF (90mg/400mg)<br />

fixed dose combination daily for 12 weeks; treatment was<br />

extended to 24 weeks for those who were treatment experienced<br />

with cirrhosis following FDA approval of LDV/SOF.<br />

GT 2 or 3 HCV-infected patients received SOF 400 mg daily<br />

+RBV (1000 or 1200 mg/day divided BID) for 12 weeks or<br />

24 weeks, respectively. Results: 120 patients were enrolled at<br />

13 sites in the USA. The majority were male (94%), Caucasian<br />

(76%), and IL28B non-CC genotype (69%). 103 patients were<br />

infected with GT1 HCV infection, 10 with GT2, 6 with GT3,<br />

and 1 with GT4. The most common bleeding disorders were<br />

Hemophilia A (65%) and Hemophilia B (26%). Patients with<br />

Type 1 or 3 Von Willebrand’s Disease (3% each), Type 2 Von<br />

Willebrand’s Disease (2%), and Factor XI deficiency (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 715A<br />

similar between groups. During treatment, patients receiving<br />

FDC showed improvement of PRO scores (+6.0% in activity/<br />

energy (AE) of CLDQ-HCV, +5.0% in fatigue score (FS) of FAC-<br />

IT-F, +6.8% in physical component of SF-36; all p


716A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

nosed between 1998-2014) with SCD in remission (with sickle<br />

cell history > 30 years) All patients were placed LDV 90 mg<br />

+ SOF 400 mg a day; with food for 12 weeks Patient characteristics:<br />

the table will be included in the oral presentation<br />

(slide) or poster Medications allowed: Hydroxyurea, Tylenol,<br />

MS Contin, Methadone, Antibiotics (Levofloxacin, Doxycycline,<br />

Colchicine, Metronidazole), epoetin alpha Exclusion: Decompensated<br />

cirrhotics, HIV, HBV, Sepsis, brittle SCD with frequent<br />

flares, uncontrolled DM, cardiomyopathy, CHF NYHA class<br />

III-IV, CKD, chronic osteomyelitis, active IVDU, Daily alcohol ><br />

30 grams, Deferoxamine for 12 weeks. Side events: Nausea-<br />

9/24, constipation 6/24, diarrhea 2/24, headache 6/24,<br />

abdominal pain 3/24, renal colic 1/24, insomnia 8/24, anemia<br />

2/24, hyperbilirubinemia 4/24, UTI 3/24 Conclusion:<br />

This study demonstrates that LDV and SOF combination in SCD<br />

patients with CHC is safe and well tolerated; with an SVR12 of<br />

91.67% (22/24) with 8.3% (2/24) viral failure (in concomitant<br />

genotypes; 1a/4c and 1a/3c).<br />

Results<br />

(range 2-8 weeks) and remained negative thereafter. One<br />

patient who was RBV-ineligible and received SOF/LDV without<br />

RBV for only 12 weeks due to insurance issues developed virological<br />

relapse between 4 and 8 weeks after treatment discontinuation.<br />

Conclusions: Treatment with SOF/LDV with or without<br />

RBV for 12-24 weeks was very well tolerated and resulted in<br />

an excellent on-treatment virological response in HCV GT1<br />

relapsers to SMV+SOF treatment. SVR12 data will be reported<br />

when available.<br />

On-Treatment Virological Response Rates<br />

Disclosures:<br />

Surakit Pungpapong - Grant/Research Support: BMS, Gilead<br />

The following authors have nothing to disclose: Michael D. Leise, Kymberly D.<br />

Watt, Hugo E. Vargas, Andrew P. Keaveny, Bashar A. Aqel<br />

Disclosures:<br />

The following authors have nothing to disclose: Patrick Basu, Niraj J. Shah, Nimy<br />

John, M. Aloysius, Robert Brown<br />

1038<br />

Multicenter Experience using Sofosbuvir/Ledipasvir with<br />

or without Ribavirin to Treat Hepatitis C Genotype 1<br />

Relapsers after Simeprevir and Sofosbuvir Treatment<br />

Surakit Pungpapong 1,2 , Michael D. Leise 3 , Kymberly D. Watt 3 ,<br />

Hugo E. Vargas 4 , Andrew P. Keaveny 1,2 , Bashar A. Aqel 4 ;<br />

1 Transplant, Mayo Clinic, Jacksonville, FL; 2 Gastroenterology &<br />

Hepatology, Mayo Clinic, Jacksonville, FL; 3 Gastroenterology &<br />

Hepatology, Mayo Clinic, Rochester, MN; 4 Gastroenterology &<br />

Hepatology, Mayo Clinic, Scottsdale, AZ<br />

Background: Approximately 10-15% of patients with hepatitis<br />

C genotype 1 (HCV GT1) experience virological relapse after<br />

all-oral antiviral regimen using simeprevir (SMV) and sofosbuvir<br />

(SOF). The efficacy and safety of treating such relapsers using<br />

sofosbuvir/ledipasvir fixed-dose combination (SOF/LDV) with/<br />

without ribavirin (RBV) has not been described to date. Objective:<br />

To report the virological response and safety of SOF/<br />

LDV with or without RBV for 12-24 weeks in treating HCV<br />

GT1 relapsers after SMV+SOF treatment. Results: To date, 42<br />

patients (26% post-LT, 79% male, 17% non-white, 80% subtype<br />

1a, 86% IL28B CT/TT, 76% F3-4) started SOF/LDV with<br />

or without RBV at a median of 20 weeks (range 7-55 weeks)<br />

after the last dose of SMV+SOF treatment. Five and 25 patients<br />

started SOF/LDV with RBV (dose adjusted for renal function) for<br />

12 and 24 weeks, respectively; while 12 patients who were<br />

RBV-ineligible received SOF/LDV without RBV for 24 weeks.<br />

Minimal adverse events were reported in those without RBV;<br />

48% patients who received RBV developed significant anemia<br />

requiring RBV dose reduction and/or discontinuation. In<br />

LT recipients, minimal immunosuppression dose adjustments<br />

were required and no biopsy-proven acute rejection occurred.<br />

On-treatment virological response rates are shown in the Table.<br />

Of 39 patients who received treatment at least 4 weeks, all<br />

achieved undetectable HCV RNA at a median of 4 weeks<br />

1039<br />

RUBY-I: Ombitasvir/Paritaprevir/Ritonavir + Dasabuvir<br />

+/- Ribavirin in Non-cirrhotic HCV Genotype 1-infected<br />

Patients With Severe Renal Impairment or End-Stage<br />

Renal Disease<br />

Paul J. Pockros 1 , K. Rajender Reddy 2 , Parvez S. Mantry 3 , Eric<br />

Cohen 4 , Michael Bennett 5 , Mark S. Sulkowski 6 , David Bernstein 7 ,<br />

Thomas Podsadecki 4 , Daniel E. Cohen 4 , Nancy Shulman 4 , Deli<br />

Wang 4 , Amit Khatri 4 , Manal Abunimeh 4 , Eric Lawitz 8 ; 1 Scripps<br />

Clinic, La Jolla, CA; 2 University of Pennsylvania, Philadelphia,<br />

PA; 3 The Liver Institute at Methodist Dallas, Dallas, TX; 4 AbbVie<br />

Inc., North Chicago, IL; 5 Medical Associates Research Group,<br />

San Diego, CA; 6 Johns Hopkins University, Baltimore, MD; 7 North<br />

Shore University Hospital, Manhasset, NY; 8 The Texas Liver Institute,<br />

University of Texas Health Science Center, San Antonio, TX<br />

Background: HCV is common among patients (pts) with endstage<br />

renal disease. Co-formulated ombitasvir/paritaprevir/r<br />

(paritaprevir identified by AbbVie and Enanta, co-dosed with<br />

ritonavir [r]) and dasabuvir (3D) do not require dose adjustment<br />

in pts with renal insufficiency. In Phase 3 trials, 3D+/- RBV<br />

showed high SVR rates and low rates of discontinuation due to<br />

adverse events in HCV genotype 1 (GT1)-infected pts. RUBY-I is<br />

an ongoing open-label study evaluating 3D+/-RBV in pts with<br />

stage 4 or 5 chronic kidney disease (CKD) and GT1 infection.<br />

Methods: Cohort 1 of RUBY-I enrolled treatment-naïve, non-cirrhotic<br />

adults with GT1 infection and CKD stage 4 (estimated<br />

GFR 15-30 mL/min/1.73m 2 ) or 5 (eGFR


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 717A<br />

pts with stage 4 or 5 CKD in this ongoing trial, 3D+/-RBV has<br />

been well tolerated, with no premature treatment discontinuations.<br />

Hemoglobin decreases were managed with RBV interruption,<br />

which does not appear to affect efficacy. There have been<br />

no virologic failures to date.<br />

Cohort 1: Baseline characteristics.<br />

Disclosures:<br />

Paul J. Pockros - Advisory Committees or Review Panels: Janssen, Merck, BMS,<br />

Gilead, AbbVie; Consulting: Lumena, Beckman Coulter; Grant/Research Support:<br />

Intercept, Janssen, BMS, Gilead, Lumena, Beckman Coulter, AbbVie, RMS,<br />

Merck; Speaking and Teaching: AbbVie, Janssen, Gilead<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

Parvez S. Mantry - Consulting: Salix, Gilead, Janssen, Abbvie, BMS; Grant/<br />

Research Support: Salix, Merck, Gilead, Boehringer-Ingelheim, Mass Biologics,<br />

Vital Therapies, Santaris, mass biologics, Bristol-Myers Squibb, Abbive, Bayer-Onyx,<br />

Shinogi, Tacere, Intercept; Speaking and Teaching: Gilead, Janssen,<br />

Salix<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

David Bernstein - Advisory Committees or Review Panels: Gilead; Consulting:<br />

Abbvie, BMS, Merck, Janssen; Grant/Research Support: Gilead, Abbvie, BMS,<br />

Merck, Janssen, Genentech; Speaking and Teaching: Abbvie, BMS, Merck,<br />

Gilead<br />

Thomas Podsadecki - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Daniel E. Cohen - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Nancy Shulman - Employment: Abbvie<br />

Deli Wang - Employment: AbbVie Inc<br />

Amit Khatri - Employment: AbbVie, Inc; Patent Held/Filed: AbbVie, Inc; Stock<br />

Shareholder: AbbVie, Inc<br />

Manal Abunimeh - Employment: AbbVie<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

The following authors have nothing to disclose: Eric Cohen, Michael Bennett<br />

1040<br />

Impact of baseline albumin levels in a Phase 3 study<br />

(OPTIMIST-2) of 12 weeks of simeprevir (SMV) plus<br />

sofosbuvir (SOF) in treatment-naïve or -experienced<br />

patients with chronic HCV genotype 1 infection and cirrhosis<br />

Eric Lawitz 1 , Eric M. Yoshida 2 , Reem H. Ghalib 3 , Marcelo Kugelmas<br />

4 , Ronald Kalmeijer 5 , Katrien Janssen 6 , Oliver Lenz 6 , Bart<br />

Fevery 6 , Guy De La Rosa 7 , Rekha Sinha 5 , James Witek 5 ; 1 Texas<br />

Liver Institute, University of Texas Health Science Center, San Antonio,<br />

TX; 2 University of British Columbia, Vancouver, BC, Canada;<br />

3 Texas Clinical Research Institute, Arlington, TX; 4 South Denver<br />

Gastroenterology, PC, Denver, CO; 5 Janssen Research & Development,<br />

LLC, Titusville, NJ; 6 Janssen Infectious Diseases BVBA,<br />

Beerse, Belgium; 7 Janssen Global Services, LLC, Titusville, NJ<br />

Purpose: In the Phase 3 OPTIMIST-2 (NCT02114151) study<br />

in HCV GT1-infected pts with cirrhosis, SMV+SOF for 12wks<br />

achieved superiority in sustained virologic response 12wks<br />

after end of treatment (SVR12) vs a historical control. This analysis<br />

evaluated SVR12 by baseline (BL) albumin. Methods: Pts<br />

with documented cirrhosis received 12wks of SMV 150mg<br />

QD+SOF 400mg QD. Primary endpoint: SVR12. Univariate<br />

logistic regression established the relationship between BL albumin<br />

and SVR12. SVR12 rates were evaluated post hoc by<br />

BL albumin and presence of other factors: BL Q80K polymorphism;<br />

IL28B genotype; BMI; platelet count; prior treatment;<br />

sex. Results are presented for the non-VR excluded population.<br />

Results: 100pts were treated (male 81%; median age 58yrs;<br />

Black/African American 18%; IL28B non-CC 72%; GT1a 69%<br />

[of these: +/- Q80K 46/54%]; treatment-naïve 49%; BL albumin,<br />

median 39 [range 29–47] g/L). Overall SVR12 was 85%.<br />

A strong univariate relationship was observed between BL albumin<br />

and SVR12 (AUC ROC 0.79). Analysis of the correlation<br />

of BL albumin and SVR12 rates identified a level of 40g/L as<br />

an optimal cut-off. 51% and 49% pts had BL albumin


718A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Eric M. Yoshida - Advisory Committees or Review Panels: Hoffman LaRoche, Gilead<br />

Sciences Inc, Abbvie; Grant/Research Support: Abbvie, Hoffman LaRoche,<br />

Merck Inc, Vertex Inc, Jannsen Inc, Gilead Sciences Inc, Boeringher Ingleheim<br />

Inc, Astellas; Speaking and Teaching: Gilead Sciences Inc, Merck Inc<br />

Reem H. Ghalib - Grant/Research Support: Bristol Myers Squibb Pharmaceuticals,<br />

Vertex Pharmaceuticals, Janssen, Merck, Genentech, Idenix, Zymogenetics,<br />

Pharmasset, Anadys, Duke Clinical Research Institute, Achillion, Boehringer Ingelheim,<br />

Gilead Pharmaceuticals, Virochem Pharmaceuticals, Abbott, Medtronic<br />

Inc, Novartitis, Roche, Schering Plough, Salix, tibotec, Inhibitex, Takeda, Abbvie<br />

Marcelo Kugelmas - Advisory Committees or Review Panels: Merck, Abbvie,<br />

Janssen, Salix; Consulting: Abbvie, Merck, Gilead, Janssen; Grant/Research<br />

Support: Abbvie, Intercept, Hologic, Gilead, Janssen, Roche, Anadys, Salix;<br />

Speaking and Teaching: Abbvie, Gilead, Merck, Janssen, Salix<br />

Ronald Kalmeijer - Employment: Johnson and Johnson<br />

Katrien Janssen - Employment: Janssen Pharmaceutica NV<br />

Oliver Lenz - Employment: Janssen; Stock Shareholder: Janssen (J&J)<br />

Bart Fevery - Employment: Janssen Infectious Diseases BVBA; Stock Shareholder:<br />

Janssen Infectious Diseases BVBA<br />

Guy De La Rosa - Employment: Johnson & Johnson<br />

Rekha Sinha - Employment: Janssen Pharmaceutica<br />

James Witek - Employment: Johnson & Johnson; Stock Shareholder: Johnson &<br />

Johnson<br />

1041<br />

Twelve Weeks of Sofosbuvir plus Ribavirin is Effective<br />

for Treatment of Genotype 2 HCV in Difficult to Treat<br />

U.S. Veterans with Cirrhosis – Results of the VAL-<br />

OR-HCV Study<br />

Samuel B. Ho 1 , Alexander Monto 2 , Adam Peyton 3 , David E.<br />

Kaplan 4 , Sean Byrne 5 , Scott Moon 5 , Yanni Zhu 5 , Star Seyedkazemi<br />

5 , Lorenzo Rossaro 5 , Diana M. Brainard 5 , William Guyer 5 ,<br />

Obaid S. Shaikh 6 , Michael Fuchs 7 , Timothy R. Morgan 8 ; 1 Gastroenterology,<br />

VA San Diego Healthcare System, San Diego, CA;<br />

2 Gastroenterology, UCSF/SFVAMC, San Francisco, CA; 3 Gastroenterology,<br />

Miami VA Healthcare System, Miami, FL; 4 Gastroenterology,<br />

Philadelphia VA Medical Center, Philadelphia, PA; 5 Gilead<br />

Sciences, Foster CIty, CA; 6 VA Pittsburgh Healthcare System, Pittsburgh,<br />

PA; 7 Richmond VA Medical Center, Richmond, VA; 8 Gastroenterology,<br />

VA Long Beach Healthcare System, Long Beach, CA<br />

Background: The prevalence of chronic HCV in U.S. veterans<br />

is approximately 4%. Veterans often have medical or psychiatric<br />

comorbidities that limit their enrollment in registration trials.<br />

Registries and surveillance reports show lower efficacy<br />

and higher discontinuation rates among veterans than in other<br />

patients. Data are limited on use of sofosbuvir (SOF) and ribavirin<br />

(RBV) for treatment of patients with cirrhosis from genotype<br />

2 (GT2) HCV infection, due to relatively small number<br />

of such subjects enrolled in clinical trials. Therefore, we conducted<br />

an open-label study to evaluate the safety and efficacy<br />

of 12 weeks of SOF+RBV in GT2 HCV infected veterans with<br />

cirrhosis. Methods: VALOR-HCV was conducted exclusively at<br />

15 U.S. VA sites in both treatment-naïve (TN) and treatment-experienced<br />

(TE) GT2 patients with compensated cirrhosis. The<br />

study had expanded inclusion criteria and did not exclude<br />

patients with active alcohol and polysubstance abuse. Subjects<br />

received 12 weeks of oral SOF 400 mg daily + weight-based<br />

RBV 1000-1200 mg/day in divided doses. The primary endpoint<br />

was sustained virologic response at 12 weeks after end<br />

of treatment (SVR12). Results: 66 cirrhotic GT2 HCV infected<br />

subjects, 47 TN and 19 TE, were enrolled. All subjects were<br />

male, mean age was 63 years (range: 49 - 85 years) and<br />

mean BMI was 30.2 kg/m 2 (range: 17.7-51.0 kg/m 2 ). Fifteen<br />

percent of subjects were African American, 9% Hispanic,<br />

and 33% classified as intolerant to or ineligible for treatment<br />

with interferon. Among subject comorbidities were vascular<br />

(71.2%), metabolic (62.1%), and psychiatric (56.1%) disorders.<br />

Cirrhosis was diagnosed by FibroTest (58%), FibroScan<br />

(26%), biopsy (12%), and imaging (5%). Ten TN (21%) and 2<br />

TE (11%) subjects had SOF adherence rates of


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 719A<br />

respondents ascertained prior awareness of HCV infection status.<br />

Population percentages and p-values were weighted to<br />

adjust for sampling design. Results: 162 of 458 NHANES<br />

respondents who tested positive for HCV antibodies or RNA<br />

during 2003-2012 participated in the follow-up survey. Of<br />

these, 91 (59%) reported that they were already aware of<br />

their HCV status when informed of their positive HCV test in<br />

NHANES. Rates of prior HCV status awareness varied considerably;<br />

71% of adults born in 1945-1965 reported prior<br />

awareness, for example, compared to 28% of adults not in<br />

major risk groups (p=0.096). People living above 2x Federal<br />

poverty level (FPL) were more likely (72%) than those below 2x<br />

FPL (48%) to have been aware of their HCV status (p=0.019).<br />

Conclusions: HCV screening practices before 2010 appear to<br />

have identified just over half of total HCV cases in the non-institutionalized<br />

US population. Furthermore, large disparities exist<br />

in awareness of HCV status, particularly among marginalized<br />

and seemingly low-risk groups. With new therapies promising<br />

higher certainty of cure for people with HCV, development of<br />

effective screening policies is of great importance.<br />

Disclosures:<br />

Mark T. Linthicum - Employment: Precision Health Economics<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Gigi Moreno - Consulting: AbbVie<br />

Yuri Sanchez - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Darius Lakdawalla - Consulting: AbbVie Pharmaceuticals; Stock Shareholder:<br />

Precision Health Economics<br />

Steven Marx - Employment: AbbVie; Stock Shareholder: AbbVie<br />

The following authors have nothing to disclose: Brian R. Edlin<br />

1043<br />

Sustained Virologic Response to Direct Acting Antiviral<br />

Therapy for Chronic Hepatitis C Infection is Associated<br />

with Improvement of Components of the Metabolic Syndrome<br />

Mark Pedersen 1 , David W. Backstedt 2 , Bobby Kakati 2 , Myunghan<br />

Choi 3 , Anil B. Seetharam 4,5 ; 1 Department of Internal Medicine,<br />

Banner University Medical Center, Tempe, AZ; 2 Department of<br />

Gastroenterology, Banner University Medical Center, Phoenix, AZ;<br />

3 Arizona State University College of Nursing and Health Care<br />

Innovation, Tempe, AZ; 4 Banner Transplant and Advanced Liver<br />

Disease, Phoenix, AZ; 5 University of Arizona College of Medicine<br />

Phoenix, Phoenix, AZ<br />

Rationale: Chronic hepatitis C virus (HCV) infection has been<br />

linked to the metabolic syndrome. While direct acting antiviral<br />

(DAA) therapy for HCV is often successful in achieving<br />

sustained virologic response (SVR), pleiotropic effects on<br />

components of the metabolic syndrome have not been fully<br />

characterized. Aim: To evaluate the effect of DAA HCV therapy<br />

on components of the metabolic syndrome. Methods: A<br />

prospective, cohort study of consecutively enrolled, mono-infected<br />

chronic HCV patients initiated on treatment with a DAA<br />

regimen. Consecutively enrolled patients received one of the<br />

following: 1) pegylated interferon alfa 2a (IFN) + sofosbuvir<br />

(SOF) + ribavirin (RBV); 2) SOF + RBV; 3) SOF + simeprevir<br />

(SMV) or 4) ledipasvir (LDV) + SOF in accordance with AASLD<br />

guidelines. Metabolic parameters including: body mass index<br />

(BMI), systolic blood pressure (BP), diastolic BP, fasting glucose,<br />

total cholesterol, high density lipoprotein (HDL), low density<br />

lipoprotein (LDL), triglycerides, and fasting glucose were measured<br />

prior to and at the end of treatment. Subgroup analysis<br />

evaluated metabolic profile changes between groups achieving<br />

SVR at 12 weeks. Changes in metabolic parameters were<br />

analyzed with student’s t test, p < 0.05 significant. Results:<br />

Patients (n=218) completed treatment with a DAA based regimen:<br />

IFN + SOF + RBV (17%); SOF + RBV (54.6%); SMV +<br />

SOF (23.4%) and SOF + LDV (6%). Of 218 enrolled, 61.5%<br />

genotype 1 and 56.4% cirrhotic. Previous IFN exposure: 64%<br />

naïve followed by non-response (26%) and relapse (10%). The<br />

highest prevalence of concomitant metabolic syndrome parameter<br />

prior to treatment was hypertension (38.5%) followed by<br />

diabetes (16.5%) and hyperlipidemia (15.3%). For the entire<br />

cohort, end of treatment (ETR) was 98%, with 79% going on<br />

to achieve SVR 12 (19% relapsers). At ETR, significant reductions<br />

in BMI, systolic and diastolic BP, total cholesterol, and<br />

LDL were noted in the entire cohort (all p


720A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

analyzed with MEGA 5.0 software to identify RAVs to nine<br />

DAAs (Boceprevir, Telaprevir, Simeprevir, Paritaprevir, Daclatasvir,<br />

Ledipasvir, Omitasvir, Sofosbuvir and Dasabuvir) in the<br />

HCV genome, which were summarized from the up-to-date<br />

literature available. Results: This analysis collected 1459 fulllength<br />

HCV sequences from GenBank, 70.7% of which carried<br />

at least one dominant resistance variant. Geographically, the<br />

highest RAV frequency occurred in Asia (87.7%), followed<br />

by Africa (73.8%), Europe (69.0%) and America (63.3%).<br />

The highest RAV frequency was observed in genotype (GT) 6<br />

sequences (83.8%), followed by GT2 (83.2%), GT4 (77.6%),<br />

GT1 (67.6%) and GT3 (50.0%). Furthermore, 46.0%, 30.6%<br />

and 10.8% of sequences were observed RAVs to NS5A inhibitors,<br />

NS3 protease inhibitors and their combinations respectively.<br />

However, RAVs to NS5B nucleos(t)ide inhibitor (NI)<br />

and NI-based combinations were uncommon observed (< 3%<br />

of sequences). As expected, RAVs were rare (0.2% to 2.1%)<br />

detected to the IFN-free regimens recommended by the currently<br />

guideline. Conclusion: The global overall prevalence of<br />

DAA-induced RAVs was high without regard for geography or<br />

genotype. Similarly, RAVs to NS5A and NS3 inhibitors were<br />

mainly observed. In contrast, NS5B NI-based multi-DAA regimens<br />

had a low prevalence of RAVs, suggesting that NS5B<br />

NI-based DAA regimens are the most promising strategies for<br />

a long-term HCV infection cure. This result was consistent with<br />

the currently recommended IFN-free regimens.<br />

Disclosures:<br />

The following authors have nothing to disclose: Peng Hu, Zhi-wei Chen, Hu Li,<br />

Hong Ren<br />

SOLAR1 (NCT01938430) and SOLAR 2 (NCT02010255)<br />

<strong>studies</strong>. On treatment and post-treatment HCV RNA measurements<br />

from LT patients who received CyA or TAC were<br />

compared. Patients who took both or neither of the immunosuppresants<br />

were excluded. Final data for post-treatment Week<br />

12 for SOLAR 2 are pending and will be presented; however,<br />

relapse rates based on post-treatment Week 4 data were calculated.<br />

HCV RNA was measured using the Roche COBAS®<br />

Ampliprep®/Cobas TaqMan HCV Test, Version 2.0 (CAP/<br />

CTM HCV v2.0) with a lower limit of quantification (LLOQ) of<br />

15 IU/mL. Results: A total of 405 post-transplantation patients<br />

were included in the analysis: 309 (76%) patients had received<br />

TACand 96 (24%) had received CyA. The number of patients<br />

with decompensated disease receiving TAC and CyA was 73<br />

(24%) and 22 (23%) respectively. Median, Q1, and Q3 baseline<br />

HCV RNA values were similar among patients receiving<br />

TAC (6.5, 6.0, 6.8) and CyA (6.6, 6.1, 6.9). The percentage<br />

of patients with an HCV RNA < 800,000 IU/mL was 21%<br />

versus 18%, for TAC and CyA. The percent of patients who<br />

were


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 721A<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

Michael R. Charlton - Grant/Research Support: GIlead Sciences, Merck, Janssen,<br />

AbbVie, Novartis<br />

Xavier Forns - Consulting: Jansen, Abbvie; Grant/Research Support: Jansen,<br />

Gilead<br />

Gregory T. Everson - Advisory Committees or Review Panels: Roche/Genentech,<br />

Abbvie, Galectin, Boehringer-Ingelheim, Eisai, Bristol-Myers Squibb, HepC<br />

Connection, BioTest, Gilead, Merck; Board Membership: HepQuant LLC, PSC<br />

Partners, HepQuant LLC; Consulting: Abbvie, BMS, Gilead, Bristol-Myers Squibb;<br />

Grant/Research Support: Roche/Genentech, Pharmassett, Vertex, Abbvie, Bristol-Myers<br />

Squibb, Merck, Eisai, Conatus, PSC Partners, Vertex, Tibotec, GlobeImmune,<br />

Pfizer, Gilead; Management Position: HepQuant LLC, HepQuant LLC;<br />

Patent Held/Filed: Univ of Colorado; Speaking and Teaching: Abbvie, Gilead<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Michael P. Curry - Advisory Committees or Review Panels: Bristol Meyers Squib,<br />

Abbvie; Grant/Research Support: Gilead Sciences, Mass Biologics, Merck,<br />

Salix, Conatus; Stock Shareholder: Achilion<br />

1046<br />

Effectiveness of 8 or 12 week LDV/SOF in treatment-naïve<br />

patients with non-cirrhotic, genotype 1 Hepatitis<br />

C: Real-world experience from the TRIO Network<br />

Michael P. Curry 1 , Bruce Bacon 2 , Steven L. Flamm 3 , Naoky CS C.<br />

Tsai 4 , Douglas Dieterich 5 , Scott Milligan 6 , Kris V. Kowdley 7 ; 1 Beth<br />

Israel Deaconess Medical Center, Boston, MA; 2 Saint Louis University<br />

School of Medicine, St. Louis, MO; 3 Northwestern University<br />

Feinberg School of Medicine, Chicago, IL; 4 Queens Medical Center,<br />

University of Hawaii, Honolulu, HI; 5 Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 6 Trio Health Analytics, Newton,<br />

MA; 7 Liver Care Network, Swedish Medical Center, Seattle, WA<br />

BACKGROUND: The current recommendation for the treatment<br />

of genotype 1 non-cirrhotic patients who are naïve to treatment<br />

is 12 weeks ledipasvir/sofosbuvir (LDV/SOF) though patients<br />

with baseline HCV RNA 2. SUMMARY: An examination of a<br />

real-world heterogeneous Hepatitis C population revealed a<br />

preference for 12 weeks even when baseline HCV RNA is<br />


722A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

combination therapy of Asunaprevir and Daclatasvir in chronic<br />

hepatitis C patients improved NK cell activity by increasing<br />

the frequency of CD56 bright NK cell in peripheral blood. It was<br />

revealed that reduction of HCV load by DAAs treatment in<br />

chronic hepatitis C patients could restore NK cell activity that<br />

was reduced by HCV.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ikuo Nakamura, Yoshihiro Furuichi,<br />

Katsutoshi Sugimoto, Yoshiyuki Kobayashi, Fuminori Moriyasu<br />

1048<br />

Demographic Predictors of Direct Acting Antiviral<br />

Receipt in Individuals with Hepatitis C Virus Infection<br />

Fasiha Kanwal 1 , Jennifer R. Kramer 1 , Yumei Cao 1 , Thom Taylor 2 ,<br />

Donna Smith 1 , Steven Asch 2 , Hashem B. El-Serag 1 ; 1 Michael E<br />

DeBakey VA and Baylor College of Medicine, Houston, TX; 2 VA<br />

Palo Alto Health Care System, Menlo Park, CA<br />

Background: Direct acting antiviral agents (DAAs) are highly<br />

effective yet very expensive. The extent to which the receipt of<br />

DAAs varies among socio-demographic groups is unknown.<br />

We sought to examine how race, gender, and age predicted<br />

receipt of DAAs in a U.S. cohort of HCV infected veterans.<br />

Methods: We used the Veterans Administration (VA) Corporate<br />

Data Warehouse to examine a cohort of patients with a<br />

positive HCV RNA test that had an outpatient visit to one of<br />

the VA facilities nationwide in 2014. We excluded patients<br />

if they had achieved sustained virological response prior to<br />

12/2013. DAA receipt included any prescription for sofosbuvir,<br />

simeprevir, Harvoni or Viekira from 12/1/2013 to<br />

2/10/2015. We examined demographic predictors of DAA<br />

receipt including age, race/ethnicity (Black, Hispanic, non-Hispanic<br />

white, others), and gender while adjusting for HCV genotype,<br />

HIV co-infection, diabetes, depression, alcohol or drug<br />

abuse, cirrhosis, and comorbidity using multivariable hierarchical<br />

logistic regression models. We also examined interactions<br />

between race/ethnicity and age or gender. Results: We had<br />

145,596 patients in our cohort. Their mean age was 60.1<br />

years (sd=7.3), 37.0% were Black, 47.8% non-Hispanic white,<br />

and 4.2% were Hispanic. Majority (96.7%) were males and<br />

62.0% had genotype 1, 7% genotype 2, 4.6% genotype 3,<br />

0.7% genotype 4; 25.7% did not have an HCV genotype test.<br />

In the cohort, 13,238 (9.1%) received DAAs. Black patients<br />

were significantly less likely to receive DAAs than non-Hispanic<br />

whites (OR=0.79; 95%CI: 0.74-0.84). There were no<br />

significant differences in DAAs receipt between non-Hispanic<br />

whites and patients of other races/ethnicities. Women were<br />

more likely to receive DAAs than men (OR=1.12; 1.02-1.24).<br />

Younger patients (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 723A<br />

MELD Change from Baseline to Follow-up Week 12<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

Gregory T. Everson - Advisory Committees or Review Panels: Roche/Genentech,<br />

Abbvie, Galectin, Boehringer-Ingelheim, Eisai, Bristol-Myers Squibb, HepC<br />

Connection, BioTest, Gilead, Merck; Board Membership: HepQuant LLC, PSC<br />

Partners, HepQuant LLC; Consulting: Abbvie, BMS, Gilead, Bristol-Myers Squibb;<br />

Grant/Research Support: Roche/Genentech, Pharmassett, Vertex, Abbvie, Bristol-Myers<br />

Squibb, Merck, Eisai, Conatus, PSC Partners, Vertex, Tibotec, GlobeImmune,<br />

Pfizer, Gilead; Management Position: HepQuant LLC, HepQuant LLC;<br />

Patent Held/Filed: Univ of Colorado; Speaking and Teaching: Abbvie, Gilead<br />

Xavier Forns - Consulting: Jansen, Abbvie; Grant/Research Support: Jansen,<br />

Gilead<br />

The following authors have nothing to disclose: Hans Van Vlierberghe, Robert<br />

Brown<br />

Disclosures:<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Geoff McCaughan - Advisory Committees or Review Panels: Gilead<br />

Michael P. Curry - Advisory Committees or Review Panels: Bristol Meyers Squib,<br />

Abbvie; Grant/Research Support: Gilead Sciences, Mass Biologics, Merck,<br />

Salix, Conatus; Stock Shareholder: Achilion<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Jill M. Denning - Employment: Gilead Sciences, Inc.<br />

Sarah Arterburn - Employment: Gilead Sciences Inc.; Stock Shareholder: Gilead<br />

Sciences Inc.<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Princy N. Kumar - Consulting: Janssen, ViiV Healthcare; Grant/Research Support:<br />

Janssen, Merck, GSK, Gilead; Stock Shareholder: Merck, Phizer, Johnson&-<br />

Johnson, GSK, Gilead<br />

Eric M. Yoshida - Advisory Committees or Review Panels: Hoffman LaRoche, Gilead<br />

Sciences Inc, Abbvie; Grant/Research Support: Abbvie, Hoffman LaRoche,<br />

Merck Inc, Vertex Inc, Jannsen Inc, Gilead Sciences Inc, Boeringher Ingleheim<br />

Inc, Astellas; Speaking and Teaching: Gilead Sciences Inc, Merck Inc<br />

Massimo Colombo - Advisory Committees or Review Panels: BRISTOL-MEY-<br />

ERS-SQUIBB, SCHERING-PLOUGH, ROCHE, GILEAD, BRISTOL-MEYERS-SQUIBB,<br />

SCHERING-PLOUGH, ROCHE, GILEAD, Janssen Cilag, Achillion; Grant/<br />

Research Support: BRISTOL-MEYERS-SQUIBB, ROCHE, GILEAD, BRISTOL-MEY-<br />

ERS-SQUIBB, ROCHE, GILEAD; Speaking and Teaching: Glaxo Smith-Kline,<br />

BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH, ROCHE, NOVARTIS, GILEAD,<br />

VERTEX, Glaxo Smith-Kline, BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH,<br />

ROCHE, NOVARTIS, GILEAD, VERTEX, Sanofi<br />

Bart van Hoek - Advisory Committees or Review Panels: Janssen-Cilag, Bristol<br />

Meyers Squib, Gilead, Merck, Abbvie<br />

Jean-Francois Dufour - Advisory Committees or Review Panels: Bayer, BMS, Gilead,<br />

AbbVie, Novartis, Sillagen, Genfit<br />

Charles S. Landis - Grant/Research Support: Gilead, Abbvie, BMS<br />

David J. Mutimer - Advisory Committees or Review Panels: Gilead Sciences,<br />

AbbVie, Janssen, MSD, BMS; Speaking and Teaching: Gilead Sciences, AbbVie,<br />

Janssen, BMS<br />

Steven L. Flamm - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, AbbVie, Janssen, Salix; Consulting: Merck, Janssen, Bristol Myers<br />

Squibb, AbbVie, Salix, Gilead; Grant/Research Support: Janssen, Bristol Myers<br />

Squibb, Gilead, AbbVie; Speaking and Teaching: Salix<br />

Michael R. Charlton - Grant/Research Support: GIlead Sciences, Merck, Janssen,<br />

AbbVie, Novartis<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

1050<br />

Adherence and Discontinuation Rates of Sofosbuvir-Based<br />

Regimens: Modeling Real World Experience in<br />

a Large Managed Care Organization<br />

Pravin S. Kamble 2 , David R. Walker 1 , Steven Marx 1 , Ray Harvey 2 ,<br />

Claudia L. Uribe 2 , Suvapun Bunniran 2 , Jenna Collins 2 ; 1 HEOR,<br />

Abbvie, North Chicago, IL; 2 Outcomes, Comprehensive Health<br />

Insights, Louisville, KY<br />

BACKGROUND Discontinuation (DC) rates of sofosbuvir (SOF)<br />

based medications are generally very low in clinical trials (<<br />

1%). Adherence (ADH) rates on the other hand are rarely<br />

reported in clinical trials. Only a handful of <strong>studies</strong> have examined<br />

this issue outside of clinical trials, however, those <strong>studies</strong><br />

did not comprehensively attempt to adjust for factors that may<br />

impact DC and ADH rates. OBJECTIVES The objective was to<br />

assess medication ADH and DC rates in SOF-based regimens<br />

and identify factors potentially associated with poor ADH and<br />

DC rates in patients with the hepatitis C virus (HCV). METH-<br />

ODS A retrospective cohort study using administrative claims<br />

data from a large managed care organization from May 2013<br />

to Sept 2014. Plan types included Medicare Advantage Prescription<br />

Drug (MAPD) and commercial (COM) which provide<br />

medical and pharmacy benefits. The study cohort included<br />

patients initiating 12 or 24-week treatment on SOF. The index<br />

date was defined as the first prescription fill date for SOF from<br />

Nov 2013 to March or May 2014 with a follow up period of<br />

12 or 24 weeks. Continuous enrollment for 6 months pre-index<br />

and 4 or 6 months post index date was required. ADH was calculated<br />

using proportion of days covered (PDC). Patients with<br />

at least 85% PDC were categorized as adherent. Patients were<br />

deemed discontinued if a gap of >14 days exists between<br />

fills. Regression analyses were conducted to identify baseline<br />

covariates associated with ADH or DC among plan members.<br />

Covariates included age, gender, risk score, plan type, geographic<br />

region, treatment co-pay, prior treatment experience,<br />

use of interferon, treatment duration and baseline healthcare<br />

use and costs. RESULTS In total, 514 MAPD and 63 COM<br />

members initiated HCV treatment with DAA. Patients were 63%<br />

male, mean age of 60 years, 84% on 12-week treatment,<br />

and 36% of SOF members were also on peginteferon. The<br />

discontinuation rate was 18% and the percent of members<br />

considered non-adherent was 14%. Based on the regression<br />

analysis, only older age, higher comorbidity risk score and<br />

use of peginterferon were significantly positively associated<br />

with non-adherence after controlling for other covariates. For<br />

12-week treatment patients, being in a non-HMO plan or in<br />

Medicare were predictors of being at lower risk of medication<br />

discontinuation. CONCLUSIONS In a real world setting, 14%<br />

of members on a new DAA are not adherent and 18% had<br />

fill gaps greater than 14 days. Sicker, older members are less<br />

likely to be adherent, whereas non-HMO and MAPD members


724A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

have lower risk of discontinuation. Additional patient support<br />

may be needed to optimize ADH in these patients.<br />

Disclosures:<br />

Pravin S. Kamble - Employment: Comprehensive Health Insights Inc.<br />

David R. Walker - Employment: Abbvie; Stock Shareholder: Abbvie, Baxter<br />

Healthcare<br />

Steven Marx - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Ray Harvey - Consulting: Comprehensive Health Insights<br />

Claudia L. Uribe - Employment: Humana<br />

Jenna Collins - Consulting: Comprehensive Health Insights, Humana<br />

The following authors have nothing to disclose: Suvapun Bunniran<br />

1051<br />

Turquoise-III: 12-Week Ribavirin-Free Regimen Of<br />

Ombitasvir/Paritaprevir/R And Dasabuvir For Patients<br />

With HCV Genotype 1B And Cirrhosis<br />

Fred Poordad 1 , Jordan J. Feld 2 , Roger Trinh 3 , Yves J. Horsmans 4 ,<br />

Magdy Elkhashab 5 , Stefan Bourgeois 6 , Samuel S. Lee 7 , Christophe<br />

Moreno 8 , David Bernstein 9 , Ziad Younes 10 , Akshanth R. Polepally 3 ,<br />

Kevin Howieson 3 , Bo Fu 3 , Greg Ball 3 , Nancy Shulman 3 , Edward<br />

Tam 11 ; 1 The Texas Liver Institute/University of Texas Health Science<br />

Center, San Antonio, TX; 2 Toronto Centre for Liver Disease,<br />

University of Toronto, Toronto, ON, Canada; 3 AbbVie, Inc., North<br />

Chicago, IL; 4 Université Catholique de Louvain, Brussels, Belgium;<br />

5 Toronto Liver Centre, Toronto, ON, Canada; 6 ZNA Stuivenberg,<br />

Antwerpen, Belgium; 7 University of Calgary, Calgary, AB, Canada;<br />

8 CUB Hôpital Erasme, Université Libre de Bruxelles, Brussels,<br />

Belgium; 9 North Shore University Hospital, Manhasset, NY; 10 GastroOne,<br />

Germantown, TN; 11 LAIR Centre, Vancouver, BC, Canada<br />

Introduction: Hepatitis C virus (HCV) infected patients have<br />

historically been more difficult to cure when they have cirrhosis.<br />

Treatment with the 3 direct-acting antiviral (3D) regimen<br />

of ombitasvir, paritaprevir (identified by AbbVie and Enanta,<br />

boosted with ritonavir), and dasabuvir without ribavirin (RBV)<br />

for 12 weeks has demonstrated 12-week sustained virologic<br />

response (SVR12) rates of 100% in HCV genotype (GT) 1b<br />

patients without cirrhosis, and 99% in GT1b patients with<br />

compensated cirrhosis when co-administered with RBV for 12<br />

weeks. We report the safety and efficacy of the 3D regimen<br />

without RBV in patients with HCV GT1b infection and compensated<br />

cirrhosis. Methods: Patients enrolled in this phase<br />

3b, multicenter, open-label study received 12 weeks of 3D<br />

without RBV. Both treatment-naïve and peginterferon/RBV treatment-experienced<br />

patients with compensated cirrhosis with no<br />

history of decompensation were enrolled with the following<br />

criteria: hemoglobin ≥10 g/dL, albumin ≥2.8 g/dL, platelet<br />

count ≥25 x 10 9 /L, and creatinine clearance ≥30 ml/min.<br />

Efficacy was assessed by the percentage of patients achieving<br />

SVR (HCV RNA below the level of quantitation [LLOQ;<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 725A<br />

1052<br />

High prevalence of co-morbidities and complex polypharmacy<br />

with drug-drug interaction (DDI) potential in<br />

patients with chronic Hepatitis C (CHC). Consistent findings<br />

from large primary care databases in the United<br />

Kingdom, Germany, and France<br />

Fiona Marra 1,10 , Werner Leber 3 , Stephen T. Barclay 2 , Stefan<br />

Christensen 5 , Denis Ouzan 7 , Valerie Oules 8 , Peter S. McMahon 4 ,<br />

Karel Kostev 6 , Xavier Ansolabehere 9 ; 1 NHS Greater Glasgow and<br />

Clyde, Glasgow, United Kingdom; 2 Walton Liver Clinic, Glasgow,<br />

United Kingdom; 3 Centre for Primary Care and Public Health, London,<br />

United Kingdom; 4 IMS Health UK, London, United Kingdom;<br />

5 Center for Interdisciplinary Medicine (CIM) Infectious Diseases,<br />

Münster, Germany; 6 IMS Health Germany, Frankfurt, Germany;<br />

7 Institut Arnault Tzanck, Saint-Laurent-du-Var, France; 8 Hôpital<br />

Saint Joseph, Marseille, France; 9 IMS Health France, Paris, France;<br />

10 University of Liverpool, Liverpool, United Kingdom<br />

Background Comorbidities and co-medications with DDI in<br />

patients with CHC may pose a challenge to upscaling treatment<br />

with new direct-acting antivirals (DAAs) and should be<br />

considered when selecting treatment options since DDI potential<br />

varies among regimens. We describe prevalence of co-morbidities<br />

and co-medications in large samples of CHC patients in<br />

primary care in the United Kingdom (UK), Germany (GER) and<br />

France (FRA). Methods Patients diagnosed with CHC between<br />

2011 and 04/2015 were identified using Read codes in the<br />

Clinical Practice Research Data link (UK) and ICD10 codes<br />

on the IMS® Disease Analyzer (GER/FRA). Inclusion criteria:<br />

18+ years, no evidence of sustained virological response, minimum<br />

one-year follow-up. Prescriptions within 12 months of<br />

last active date were categorized for DDI potential as either<br />

RED (contraindicated) or AMBER (additional monitoring/dose<br />

reduction required) with at least one of currently licensed DAA<br />

(OBV/PTV/r+DSV, SMV, LDV/SOF, DCV) using www.hep-druginteractions.org.<br />

Read/ICD10 coded co-morbidities were<br />

analysed. Comorbidities of interest included those that may<br />

make CHC treatment more difficult to manage. Results 7,949<br />

patients (UK: 4,644/GER: 2,737/FRA: 568) were analysed.<br />

Variation in mean age (45.8/51.8/57.1 years) suggests differences<br />

in CHC populations; 65%/58%/59% were male.<br />

Between 12%-19% of patients were on RED drugs, however<br />

only 0.9%-2.9% were on drugs contraindicated for all DAA<br />

regimen. 28.9%/35.1%/39.1% were on two or more RED<br />

or AMBER drugs, polypharmacy increased with age {18-29<br />

years (15.1%), 70+years (56.1%)}. Most common therapeutic<br />

classes for DDI are presented in table. Most common co-morbidities<br />

in the UK were: depression (23.3%), cardio-vascular<br />

disease [CVD] (21.0%) and COPD/Asthma (14.4%); compared<br />

to chronic pain syndromes (22.7%/16.6%) and hypertension<br />

(22.1%/15.1%) in GER/FRA, depression (14.7%) in<br />

GER and diabetes (8.7%) in FRA. Conclusion We observed<br />

significant co-morbidity and co-prescribing with DDI potential<br />

in CHC patients in three countries. The large difference in<br />

proportion of co-medications contraindicated to all DAA vs.<br />

only to some suggests careful selection of the DAA regimen is<br />

required. Treating patients at a younger age likely reduces the<br />

risk of DDI.<br />

RED or AMBER co-medication with at least one DAA in CHC<br />

patients in primary care (n=7,949)<br />

Disclosures:<br />

Fiona Marra - Advisory Committees or Review Panels: Gilead, AbbVie, Merck;<br />

Consulting: Gilead; Speaking and Teaching: Gilead, Abbvie, Merck, Janssen<br />

Stephen T. Barclay - Advisory Committees or Review Panels: Gilead, Janssen,<br />

MSD, Abbvie, BMS; Speaking and Teaching: Gilead, Janssen<br />

Stefan Christensen - Advisory Committees or Review Panels: BMS, Abbvie, Janssen,<br />

ViiV, Gilead, MSD; Speaking and Teaching: Gilead, MSD, Abbvie, BMS,<br />

ViiV, Reckitt Benckiser, Janssen<br />

Valerie Oules - Board Membership: Abbvie; Speaking and Teaching: gilead,<br />

BMS, Abbvie, Jansen<br />

Peter S. McMahon - Consulting: Gilead Sciences; Employment: IMS Health<br />

Karel Kostev - Consulting: IMS Health<br />

The following authors have nothing to disclose: Werner Leber, Denis Ouzan,<br />

Xavier Ansolabehere<br />

1053<br />

Real-Word Effectiveness of Ledipasvir (LDV)/Sofosbuvir<br />

(SOF)-Based Therapy for Hepatitis C Virus (HCV):<br />

Preliminary Analysis in a Large Integrated Health Care<br />

System<br />

Jennifer B. Lai 2,1 , David J. Witt 2 ; 1 Gastroenterology, Kaiser Permanente,<br />

San Rafael, CA; 2 Infectious Diseases, Kaiser Permanente<br />

Medical Center, San Rafael, CA<br />

BACKGROUND AND AIMS: Second generation direct-acting<br />

antiviral agents have become an integral component of treatment<br />

for HCV infection but little is known about their real-world<br />

effectiveness, particularly in community practice. We report<br />

virologic responses of patients with HCV genotype (GT) 1 infection<br />

receiving LDV/SOF with or without ribavirin (RBV) therapy.<br />

METHODS: Approval was obtained from the Kaiser Permanente<br />

Northern California Institutional Review Board. Treatment naïve<br />

and treatment experienced (TE) HCV GT 1 patients who started<br />

LDV/SOF-based therapy were identified by electronic prescription<br />

and medical records and analyzed on an intent-to-treat<br />

basis. End-of-Treatment (EOT) and Sustained Viral Response 4<br />

and 12 (SVR4, SVR12) were defined as HCV RNA less than<br />

the lower limit of quantification (LLOQ) upon completion of<br />

therapy or at 4 and 12 weeks after completing therapy, respectively.<br />

Liver fibrosis was assessed; patients were deemed to be<br />

(a) cirrhotic if either (i) any history of biopsy-proven cirrhosis,<br />

(ii) clinical manifestation(s) of cirrhosis or (iii) AST to Platelet<br />

Ratio Index (APRI) > 2.0 in the absence of biopsy ≤ 3 years<br />

prior to day 1 (D1) of treatment, (b) non-cirrhotic if either (i)<br />

no evidence of cirrhosis per biopsy ≤ 3 years prior to D1 or<br />

(ii) APRI < 0.5 in the absence of biopsy ≤ 3 years prior to<br />

D1 or otherwise (c) undetermined fibrosis. RESULTS: 656 GT1<br />

patients started LDV/SOF-based therapy at the time of data<br />

analysis of which 267 patients started LDV/SOF/RBV therapy<br />

for 12 weeks and 389 patients started LDV/SOF therapy for up<br />

to 24 weeks based on clinician discretion. Respective baseline<br />

characteristics were median age (61, 60 years), male (69,<br />

64%), TE (78, 16%) and cirrhotic (45, 23%), non-cirrhotic (19,<br />

37%) and undetermined fibrosis (36, 40%). We report preliminary<br />

EOT and SVR4 rates of 100% (173/173) and 98%<br />

(57/58) for patients receiving LDV/SOF/RBV therapy versus<br />

99% (223/226) and 94% (58/62) for patients receiving LDV/


726A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

SOF therapy, respectively. Additional virologic response data,<br />

including SVR12 responses, will be reported. CONCLUSIONS:<br />

We found high preliminary EOT and SVR4 rates for patients<br />

with HCV GT 1 treated with LDV/SOF-based therapy of 94 to<br />

100% which are very comparable to the 94 to 100% virologic<br />

responses reported in the published ION study series.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jennifer B. Lai, David J. Witt<br />

1054<br />

Hepatitis C Cure Could Avoid Liver Transplant In Some<br />

Cirrhotic Patients On Dialysis Listed for Simultaneous<br />

Liver Kidney Transplantation<br />

Frank Czul, David Roth, Rodrigo M. Vianna, Cynthia Levy, Paul<br />

Martin, Kalyan R. Bhamidimarri; Hepatology, University of Miami-<br />

Miller School of Medicine, Miami, FL<br />

Background: Simultaneous liver kidney transplant (SLKT) is recommended<br />

in Hepatitis C (HCV) infected dialysis patients with<br />

cirrhosis but isolated kidney transplant (KT) can be considered<br />

in a subset that have early cirrhosis without portal hypertension.<br />

HCV cure could potentially avoid the need for liver transplantation<br />

in some patients listed for SLKT. Aim: To report our new<br />

strategy in SLKT wait-listed patients, whereby liver transplant<br />

(LT) was avoided in those who achieved HCV cure and did not<br />

demonstrate portal hypertension. Methods We describe our<br />

experience of 9 HCV infected patients with cirrhosis and end<br />

stage renal disease (ESRD) on hemodialysis who were initially<br />

listed for SLKT at the Miami Transplant Institute from 2011 to<br />

2014. Of the 9 patients, 6 (67%) were male with a mean age<br />

of 59.8 (+/-5.9) infected with HCV genotype 1 (78% GT1a).<br />

All of them underwent liver biopsy which confirmed the presence<br />

of cirrhosis, 8 (89%) Child A and 1 (11%) Child B, 1<br />

decompensated by small esophageal varices (EV) and 1 by EV<br />

+ ascites. The patients underwent HCV treatment, 2 (22%) with<br />

Pegylated interferon, Ribavirin and Boceprevir and 7 (78%)<br />

with Simeprevir and half dose Sofosbuvir. Results Among them,<br />

8 patients (89%) underwent trans-jugular liver biopsy with portal<br />

pressures measurements and 1 underwent percutaneous<br />

biopsy (see table). Seven (78%) patients achieved SVR12 and<br />

given the absence of portal hypertension, they were delisted<br />

from LT and were considered only for KT. The 2 patients who<br />

did not achieve SVR continued to be listed for SLKT. At the<br />

end of treatment there was a significant decrease on mean<br />

AST and ALT from 32.1 to 18.1 (p=0.047) and 31.6 to 15<br />

(p=0.0099). No other statistically significant differences were<br />

appreciated in the laboratory data. Conclusion HCV treatment<br />

could be offered to dialysis patients after they undergo KT.<br />

However, SLKT wait-listed patients could derive a huge benefit<br />

by pre-transplant HCV cure as LT could be avoided in those<br />

with compensated cirrhosis and devoid of portal hypertension.<br />

Characteristics of SLKT wait-listed patients<br />

Paul Martin - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Merck, Gilead, Janssen, Abbvie<br />

Kalyan R. Bhamidimarri - Advisory Committees or Review Panels: Gilead, Abb-<br />

Vie, Janssen; Grant/Research Support: Bristol Squibb Myers, Biotest, Synageva,<br />

Salix, Vital Therapies, Ocera Inc<br />

The following authors have nothing to disclose: Frank Czul, Rodrigo M. Vianna,<br />

Cynthia Levy<br />

1055<br />

Association of polymorphisms in the hepatitis C virus<br />

NS5B polymerase with reduced virologic response to<br />

sofosbuvir-based treatment<br />

Johannes Vermehren 1 , Simone Susser 1 , Christoph P. Berg 2 , Martin<br />

W. Welker 1 , Bernd Kronenberger 1 , Caterina Berkowski 1 , Stefan<br />

Zeuzem 1 , Christoph Sarrazin 1 ; 1 Medizinische Klinik 1, Klinikum<br />

der J. W. Goethe-Universität, Frankfurt am Main, Germany; 2 Universitätsklinikum<br />

Tübingen, Tübingen, Germany<br />

Background Sofosbuvir (SOF) is a nucleotide inhibitor of the<br />

hepatitis C virus (HCV) RNA polymerase NS5B that has been<br />

approved for the treatment of chronic HCV infection. SOF<br />

binds to the highly conserved active site of NS5B. Therefore,<br />

SOF has shown pangenotypic activity and a high barrier to<br />

resistance in vitro and in vivo. However, the importance of a<br />

number of minor variants within the HCV NS5B polymerase for<br />

treatment outcome of SOF-based therapy without other DAAs<br />

is unknown. Methods All consecutive patients with HCV GT1<br />

infection who received a SOF-based therapy with SOF being<br />

the only DAA were included in the study. Patients received<br />

either SOF in combination with PEG-IFN and ribavirin (RBV)<br />

or SOF in combination with RBV alone. Baseline samples were<br />

tested for the presence of resistance associated variants (RAVs)<br />

at positions L159, S282, C316, V321 und S368 of the NS5B<br />

polymerase by direct sequencing. The presence of these variants<br />

was tested for associations with treatment outcome. Results<br />

In total, 54 patients with HCV GT1 infection (24 with HCV<br />

GT1b) were treated with a SOF-based regimen (SOF+RBV:<br />

n=18; SOF+PEG-IFN+RBV: n=38). In patients with HCV GT1a<br />

and available follow-up data, 23/28 (82%) achieved a sustained<br />

virologic response (SVR) whereas only 14/23 (62%)<br />

patients with GT1b achieved an SVR. The C316N RAV was<br />

detected in 2 GT1a patients, both of whom did not achieve an<br />

SVR (1 relapse, 1 treatment discontinuation due to side effects).<br />

C316N was detected in 12 patients with GT1b, of whom 6<br />

achieved an SVR and 6 did not (all with relapse). In GT1b<br />

patients, SVR rates were significantly higher in patients without<br />

RAVs compared to patients in whom C316N was present<br />

(87.5% vs. 50%; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 727A<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

The following authors have nothing to disclose: Simone Susser, Christoph P. Berg,<br />

Caterina Berkowski<br />

1056<br />

Hepatocellular carcinoma development in hepatitis C<br />

virus patients who achieved sustained viral response by<br />

interferon therapy: a large scale, long-term cohort study<br />

Yuko Nagaoki, Hiroshi Aikata, Hiromi Kan, Hatsue Fujino, Tomoki<br />

Kobayashi, Takayuki Fukuhara, Keiichi Masaki, Takashi Nakahara,<br />

Tomokazu Kawaoka, Masataka Tsuge, Akira Hiramatu,<br />

Michio Imamura, Yoshiiku Kawakami, Hidenori Ochi, Kazuaki<br />

Chayama; Department of Gastroenterology and Metabolism, Hiroshima<br />

University, Hiroshima, Japan<br />

Background We assessed risk factors for the development of<br />

hepatocellular carcinoma (HCC) following successful eradication<br />

of hepatitis C virus (HCV) with interferon therapy in a longterm,<br />

large-scale cohort study. Methods We reviewed 2262<br />

consecutive patients with HCV infection who achieved sustained<br />

viral response (SVR) by interferon therapy between January<br />

1995 and September 2013 in our study group. Among them,<br />

1094 patients who met the following inclusion criteria were<br />

enrolled in this retrospective cohort study: 1) received HCC<br />

surveillance by tumor markers (AFP, DCP), ultrasonography<br />

and/or dynamic computed tomography at least biannually; 2)<br />

underwent histological examination with liver biopsy before<br />

IFN therapy; and 3) with no evidence of HCC development<br />

prior to HCV eradication and within a year following eradication.<br />

We assessed host, pre-IFN treatment, and post-treatment<br />

risk factors associated with HCC development among these<br />

SVR patients. Results 1) Patients included 585 males aged 11<br />

to 85 years (median 60) at the time of HCV eradication. 581<br />

patients had HCV genotype 1b, and 513 had genotypes 2a,<br />

2b or other genotypes. 936 patients had interferon lambda 3<br />

(IFNL3) SNP rs8099917 genotype TT, and 158 had genotypes<br />

GG or TG. 885 patients had DEPDC5 SNP rs1012068 genotype<br />

TT, and 209 had genotype GG or TG. Histological fibrosis<br />

stage before IFN treatment was as follows: F0 or F1, 492;<br />

F2, 39; F3, 174; and F4, 34. 2) During the observation period<br />

(median 50 months: range 13-224), 36 (3%) of 1094 patients<br />

developed HCC after HCV eradication. The median period<br />

from HCV eradication to HCC development was 37 months<br />

(range 17-141), and the cumulative rates of developing HCC<br />

at 5,10 and 15 years were 4%, 6% and 12%, respectively. 3)<br />

As independent risk factors for HCC development, multivariate<br />

analysis identified older age ≥ 60 years (HR, 3.1: 95%CI,<br />

1.3-6.6: P=0.009), male sex (HR, 11.9: 95%CI, 2.8-50.0:<br />

P


728A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Treatment Outcomes<br />

N* number of patients with available SVR12<br />

Disclosures:<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

David R. Nelson - Advisory Committees or Review Panels: Merck; Grant/Research<br />

Support: Abbot, BMS, Beohringer Ingelheim, Gilead, Genentech, Merck, Bayer,<br />

Idenix, Vertex, Jansen<br />

Adrian M. Di Bisceglie - Advisory Committees or Review Panels: Gilead, AbbVie,<br />

Novartis, Bayer, BTG; Grant/Research Support: Gilead, AbbVie<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

Alexander Kuo - Advisory Committees or Review Panels: Gilead; Grant/Research<br />

Support: Gilead; Speaking and Teaching: IAS-USA<br />

Joseph K. Lim - Consulting: Merck, Boehringer-Ingelheim, Gilead, Bristol Myers<br />

Squibb, Janssen; Grant/Research Support: AbbVie, Boehringer-Ingelheim, Bristol<br />

Myers Squibb, Gilead, Hologic, Janssen<br />

Jama M. Darling - Consulting: BristolMyers Squibb; Grant/Research Support:<br />

BristolMyers Squibb<br />

Paul J. Pockros - Advisory Committees or Review Panels: Janssen, Merck, BMS,<br />

Gilead, AbbVie; Consulting: Lumena, Beckman Coulter; Grant/Research Support:<br />

Intercept, Janssen, BMS, Gilead, Lumena, Beckman Coulter, AbbVie, RMS,<br />

Merck; Speaking and Teaching: AbbVie, Janssen, Gilead<br />

Lynn M. Frazier - Advisory Committees or Review Panels: Abbvie ; Speaking and<br />

Teaching: Jansen, Gilead<br />

Saleh Alqahtani - Advisory Committees or Review Panels: Gilead Sciences, Janssen<br />

Therapeutics; Consulting: Abbvie; Grant/Research Support: Merck & Co,<br />

Inc., Abbvie, Gilead<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Michael W. Fried - Consulting: Merck, Abbvie, Janssen, Bristol Myers Squibb,<br />

Gilead; Grant/Research Support: Merck, AbbVie, Janssen, Bristol Myers Squibb,<br />

Gilead; Patent Held/Filed: HCCPlex<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

The following authors have nothing to disclose: Giuseppe Morelli, Joseph S.<br />

Galati<br />

1058<br />

Daclatasvir plus sofosbuvir with or without ribavirin for<br />

the treatment of chronic HCV in patients coinfected with<br />

HIV: Interim results of a multicenter compassionate use<br />

program<br />

Jürgen K. Rockstroh 1 , Tania M. Welzel 2 , Patrick Ingiliz 3 , Joerg<br />

Petersen 4 , Marc van der Valk 5 , Kerstin Herzer 6 , Peter Ferenci 7 ,<br />

Michael Gschwantler 8 , Markus Cornberg 9 , Thomas Berg 10 , Ulrich<br />

Spengler 1 , Ola Weiland 11 , Hartwig Klinker 12 , Markus Peck-Radosavljevic<br />

7 , Yue Zhao 13 , Maria Jesus Jimenez Exposito 14 , Stefan<br />

Zeuzem 2 ; 1 Universitätsklinikum Bonn, Bonn, Germany; 2 Universitätsklinikum<br />

der Johann Wolfgang Goethe Universität, Frankfurt,<br />

Germany; 3 Praxis Driesener Strasse, Berlin, Germany; 4 IFI Institut<br />

für Interdisziplinäre Medizin, Hamburg, Germany; 5 Academic<br />

Medical Center, University of Amsterdam, Amsterdam, Netherlands;<br />

6 Universitätsklinikum Essen (AöR), Essen, Germany; 7 Medizinische<br />

Universität Wien, Vienna, Austria; 8 Wilhelminenspital,<br />

Vienna, Austria; 9 Medizinische Hochschule Hannover, Hannover,<br />

Germany; 10 Universitätsklinikum Leipzig, Leipzig, Germany;<br />

11 Karolinska Institutet, Karolinska University Hospital Huddinge,<br />

Stockholm, Sweden; 12 Universitätsklinikum Würzburg, Würzburg,<br />

Germany; 13 Global Biostatistics, Bristol-Myers Squibb, Hopewell,<br />

NJ; 14 Research and Development, Bristol-Myers Squibb, Princeton,<br />

NJ<br />

BACKGROUND: HCV-related liver disease is a leading cause<br />

of death among HIV/HCV coinfected patients. Interferon-free<br />

regimens that can be safely co-administered with antiretroviral<br />

(ARV) therapy are needed in this population. In clinical trials<br />

(ALLY-2 study), the pangenotypic, all-oral 12-weeks regimen of<br />

daclatasvir (DCV) and sofosbuvir (SOF) was well tolerated and<br />

achieved 97% SVR rates in HIV/HCV coinfected patients receiving<br />

a wide range of ARVs. We report here interim results on the<br />

combination DCV plus SOF, with or without ribavirin (RBV) in<br />

HIV/HCV coinfected patients enrolled in a compassionate use<br />

program (CUP). METHODS: This CUP enrolled adult patients<br />

with chronic HCV infection who were at a high risk of hepatic<br />

decompensation or death within 12 months if left untreated,<br />

and who had no available treatment options. Patients received<br />

DCV 60 mg + SOF 400 mg QD, with RBV added at the physician’s<br />

discretion. Dose adjustments for DCV were allowed for<br />

concomitant use of ARVs: 30 mg with ritonavir-boosted PIs; 90<br />

mg with NNRTIs, except rilpivirine. Recommended duration<br />

of treatment was 24 weeks. This interim analysis includes 55<br />

HCV-HIV coinfected patients enrolled in the CUP with available<br />

data on March 16, 2015, of whom 15 have available SVR12<br />

data to date. RESULTS: Most patients were male (82%), white<br />

(95%), with a median age of 52 years (31–67). HIV RNA<br />

was undetectable in 18/21 patients; median baseline HIV<br />

RNA: 20 copies/mL (0-100 copies/mL); 17/27 patients had<br />

a CD4 cell count was


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 729A<br />

patients in a real-world setting, DCV+SOF±RBV demonstrated<br />

good rates of SVR12 and was well tolerated in a HCV-HIV<br />

coinfected patient population containing a high proportion of<br />

cirrhotic patients.<br />

Disclosures:<br />

Jürgen K. Rockstroh - Advisory Committees or Review Panels: Abbvie, BI, BMS,<br />

Merck, Roche, Tibotec, Abbvie, Bionor, Tobira, ViiV, Gilead, Janssen; Consulting:<br />

Novartis; Grant/Research Support: Merck; Speaking and Teaching: Abbott,<br />

BI, BMS, Merck, Roche, Tibotec, Gilead, Janssen, ViiV<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

Marc van der Valk - Advisory Committees or Review Panels: gilead, msd, bms,<br />

abbvie, janssen cilag, viiV, roche<br />

Peter Ferenci - Advisory Committees or Review Panels: Idenix, Gilead, MSD,<br />

Janssen, Salix, AbbVie, BMS; Patent Held/Filed: Madaus Rottapharm; Speaking<br />

and Teaching: Gilead, Roche<br />

Michael Gschwantler - Advisory Committees or Review Panels: Janssen, BMS,<br />

Gilead, AbbVie; Speaking and Teaching: Janssen, BMS, Gilead, AbbVie<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Ola Weiland - Advisory Committees or Review Panels: Merk, BMS, Medivir, Gilead,<br />

AbbVie; Grant/Research Support; Speaking and Teaching: Merk, Roche,<br />

BMS, Novartis, Janssen, Medivir, Gilead, AbbVie<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Maria Jesus Jimenez Exposito - Employment: Bristol-Myers Squibb<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

The following authors have nothing to disclose: Patrick Ingiliz, Kerstin Herzer,<br />

Ulrich Spengler, Hartwig Klinker, Yue Zhao<br />

1059<br />

The Significance of Immunoallegic Adverse Events<br />

Developing during Dual Oral Therapy with Daclatasvir<br />

Plus Asunaprevir in Patients with Genotype 1b HCV<br />

Infection<br />

YOHEI FUJII, Yoshihito Uchida, Kayoko Sugawara, Mie Inao,<br />

Nobuaki Nakayama, Satoshi Mochida; SAITAMA MEDICAL UNI-<br />

VERSITY, Saitama, Japan<br />

Aim: Dual oral therapy with daclatasvir (DCV) plus asunaprevir<br />

(ASV) was approved in Japan in July 2014 for patients<br />

with genotype 1b HCV. We established a novel system to<br />

quantify NS5A-RAVs using cycling-probe real-time PCR (PLos<br />

One, 2014), and performed the therapy by priority for patients<br />

without such RAVs. Although SVR4-12 was achieved in 95% of<br />

DAA-naïve patients without baseline NS5A-RAVs, severe immunoallegic<br />

hepatitis resembling drug-induced hypersensitive syndrome<br />

(DIHS) developed in one patient (Hepatology 2015;<br />

61: 400). Thus, the significance of immunoalergic adverse<br />

events during DCV/ASV therapies were evaluated. Methods<br />

& Results: A total of 224 patients (98 men. 126 women; 23<br />

to 87 years old.) received DCV/ASV administrations for 24<br />

weeks since Semtember 2014. Adverse events were seen<br />

during the therapy in 97 patients (43%); liver injury in 33<br />

(15%), eosinophilia in 31 (14%), fever higher than 38C in 22<br />

(10%) and decompasation of cirrhosis in 3 (1%). Among them,<br />

the therapy was discontinued in 18 patients (8%) including 8<br />

patients showing grade-4 elevation of serum ALT levels (>300<br />

IU/L) and 1 patient with increase of both serum bilirubin concentrations<br />

and GGT and ALP levels. Fever occurred between<br />

7 and 14 days after the initiation of therapy simultaneously<br />

with serum CRP level elevation, but was attenuated after oral<br />

intakes of NSAIDs for between 3 and 7 days, and did not<br />

recur following discontinuation of NSAIDs intakes. Among 33<br />

patients showing liver injuries, serum CRP level elevation and/<br />

or eosinophilia were seen simultaneously in 4 patients; 3 and<br />

1 patient(s) showing elevation of serum ALT levels and bilirubin<br />

concentrations, respectively, and oral and/or intravenous glucocorticoid<br />

adminstrations were required in 3 patients, since<br />

the events were not attenuated following discontinuation of<br />

dual oral therapy. In remaining 29 patients showing serum ALT<br />

level elevation, the therapy was continued for 24 weeks in 21<br />

patients, while dairy doses of ASV were reduced from 200 mg<br />

to 100 mg in 8 patients to maintain serum ALT levels less than<br />

150 IU/mL. Conclusion: Immunoallegic adverse events such as<br />

fever, serum CRP level elevation and eosinophilia occurred in<br />

frequent during dual oral therapy with DCV plus ASV, but were<br />

tolerable by NSAIDs intakes. Also, liver injuries developed frequently<br />

in these patients probable due to toxic effects by ASV.<br />

It should be noted, however, there existed patients showing<br />

immunoallergic liver injuries, presenting elevation of serum ALT<br />

levels and/or bilirubin concentrations combined with increased<br />

serum CRP levels and eosinophilia, since glucocorticoid administrations<br />

may be required for such patients.<br />

Disclosures:<br />

Yoshihito Uchida - Patent Held/Filed: SRL Inc.<br />

Satoshi Mochida - Grant/Research Support: Chugai, MSD, Tioray Medical,<br />

BMS; Speaking and Teaching: MSD, Toray Medical, BMS, Tanabe Mitsubishi<br />

The following authors have nothing to disclose: YOHEI FUJII, Kayoko Sugawara,<br />

Mie Inao, Nobuaki Nakayama<br />

1060<br />

Sofosbuvir-based treatments for patients with hepatitis<br />

C virus (HCV) mono-infection and human immunodeficiency<br />

virus (HIV)-HCV co-infection with genotype 3 in<br />

clinical practice – Results from the GErman hepatitis C<br />

COhort (GECCO)<br />

Patrick Ingiliz 1 , Knud Schewe 2 , Thomas Lutz 3 , Christoph Boesecke 4 ,<br />

Stefan Mauss 5 , Heiner W. Busch 6 , Axel Baumgarten 1 , Marcel<br />

Schuetze 1 , Guenther Schmutz 5 , Stefan Christensen 6 , Karl-Georg<br />

Simon 7 , Dietrich Hueppe 8 ; 1 Medical Center for Infectious Diseases,<br />

Berlin, Germany; 2 ICH, Hamburg, Germany; 3 Infektiologikum,<br />

Frankfurt, Germany; 4 Department of Medicine I, University<br />

of Bonn, Bonn, Germany; 5 Center for HIV and Hepatogastroenterology,<br />

Duesseldorf, Germany; 6 Center for Interdisciplinary<br />

Medicine, Muenster, Germany; 7 Practice for Gastroenterology<br />

Leverkusen,, Leverkusen, Germany; 8 Center for Gastroenterology,<br />

Herne, Germany<br />

Introduction The first direct acting antivirals (DAA) were<br />

approved in Europe in 2014 based on limited study data.<br />

In particular, HIV-HCV co-infected patients and pre-treated<br />

patients had not been systematically studied. Here, we present<br />

real-life data on efficacy and safety of sofosbuvir (SOF)-based<br />

therapies from Germany in genotype (GT) 3 patients. Methods<br />

In this multicenter cohort, all patients who were started<br />

on sofosbuvir-based treatments were documented. For the current<br />

analysis, only patients with HCV genotype (GT) 3 on the<br />

following regimes were included: SOF/PegIFN/RBV (ribavirin)/12<br />

weeks, SOF/RBV/12 or 24 weeks, SOF/DCV (daclatasvir)/12<br />

or 24 weeks, and SOF/LDV (ledipasvir)/12 or 24<br />

weeks. All patients within the GECCO cohort are part of the


730A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

prospective German hepatitis C registry. Results Overall, 836<br />

patients on SOF-containing regimens have been enrolled so<br />

far, 629 (75%) are HCV-monoinfected and 207 (25%) HIV-<br />

HCV co-infected. Of those, 155 (19%) patients were infected<br />

with HCV genotype 3, 120/155 (77%) were male. The median<br />

age (IQR) was 50 years (41 to 55). 65 patients (41%) have<br />

been unsuccesfully treated before, mainly with dual therapy<br />

consisting of pegylated interferon and ribavirin. Liver cirrhosis<br />

was diagnosed in 55 patients (37%), two patients were Child-<br />

Pugh stage B. The sustained virological response (SVR12) rate<br />

in patients treated with SOF/PegIFN/RBV (n=40) was 94%<br />

(2/31). In patients treated with SOF/RBV (n=76), the SVR12<br />

rate was 91% (21/23). In patients treated with SOF/DCV<br />

(n=17) for 12 or 24 weeks or SOF/LDV for 12 or 24 weeks,<br />

only very few patients have reached the SVR12 timepoint but all<br />

responded (4/4). Prior treatment failure did not affect treatment<br />

responses. Patients with liver cirrhosis had a lower response<br />

rate of 86% vs 97% without cirrhosis (p=0.28). Conclusion: In<br />

this large cohort patients with genotype 3 response rates were<br />

promising. Patients with liver cirrhosis responded less well to<br />

the different regimens used. This real-life data underline that<br />

the existing drug combinations are effective tools against HCV<br />

genotype 3.<br />

Disclosures:<br />

Patrick Ingiliz - Consulting: Gilead, Abbvie, Janssen Cilag; Speaking and Teaching:<br />

BMS, MSD, Gilead, Abbvie<br />

Knud Schewe - Advisory Committees or Review Panels: abbvie, gilead, msd,<br />

bms, janssen cilag, viiv, hexal<br />

Thomas Lutz - Advisory Committees or Review Panels: Gilead, MSD, AbbVie,<br />

BMS, Janssen Cilag; Grant/Research Support: Gilead, GlaxoSmithKline, Roche,<br />

MSD, AbbVie, Janssen Cilag; Speaking and Teaching: ViiV, BMS<br />

Christoph Boesecke - Speaking and Teaching: MSD, Boehringer, Gilead, BMS,<br />

Roche, Abbvie, ViiV<br />

Stefan Mauss - Advisory Committees or Review Panels: BMS, AbbVie, Janssen,<br />

ViiV, Gilead; Speaking and Teaching: BMS, AbbVie, Janssen, Gilead, MSD<br />

Heiner W. Busch - Speaking and Teaching: Tibotec, MSD, Janssen, BMS, Gilead<br />

Stefan Christensen - Advisory Committees or Review Panels: BMS, Abbvie, Janssen,<br />

ViiV, Gilead, MSD; Speaking and Teaching: Gilead, MSD, Abbvie, BMS,<br />

ViiV, Reckitt Benckiser, Janssen<br />

Dietrich Hueppe - Advisory Committees or Review Panels: Roche, MSD, Novatis,<br />

Gilead, BMS, Janssen<br />

The following authors have nothing to disclose: Axel Baumgarten, Marcel Schuetze,<br />

Guenther Schmutz, Karl-Georg Simon<br />

1061<br />

A comparison of direct sequencing and PCR- invader<br />

assay for the detection of NS5A Y93H mutation and<br />

response to daclatasvir and asunaprevir therapy in<br />

patients with hepatitis C virus genotype 1b<br />

Kazuhiko Hayashi, Masatoshi Ishigami, Yoji Ishizu, Teiji Kuzuya,<br />

Takashi Honda, Yoshiki Hirooka, Hidemi Goto; Gastroenterology,<br />

Nagoya University, Nagoya, Japan<br />

OBJECTIVES: Interferon free regimen such as daclatasvir (DCV)<br />

and asunaprevir (ASV) was approved for patients with hepatitis<br />

C virus (HCV) genotype 1b. Virologic failure of DCV and<br />

ASV therapy was associated with baseline NS5A polymorphism<br />

such as Y93H mutation and screening for the presence<br />

of Y93H mutation at pretreatment is recommended. The direct<br />

sequencing analysis is standard method to identify Y93H mutation<br />

but needs a lot of efforts and costs. A simple assay based<br />

on the Q-Invader technology was developed to determine and<br />

quantify NS5A Y93H mutant strains with high sensitivity. This<br />

study sought to compare two methods used to detection of<br />

Y93H mutation and evaluate the effect of Y93H mutation on<br />

response to DCV and ASV therapy. METHODS: 121 patients<br />

with HCV genotype 1b were enrolled. All patients were examined<br />

the NS5A Y93H mutation by both direct sequencing<br />

analysis and PCR- invader assay. 45 patients were received<br />

DCV and ASV combination therapy for 24 weeks and were<br />

selected for virologic response study. RESULTS: 112 of 121<br />

(92.7%) patients were able to measure Y93H mutation by<br />

direct sequencing analysis and 109 of 121 (90.1%) patients<br />

were able to measure Y93H mutation by PCR- invader assay.<br />

There are no statistically significant differences of sensitivity<br />

in both methods. 21 of 112 (18.6%) patients were identified<br />

Y93H mutation strain by direct sequencing analysis and 18 of<br />

109 (16.5%) patients were able to detect Y93H mutation by<br />

PCR- invader assay. Twelve patients were defined with Y93H<br />

mutation by both methods and the percent HCV-RNA level of<br />

the Y93H mutant were 10, 24, 53, 55, 85, 89, 89, 95, 98,<br />

99, 99, and 99, respectively. Six patients were detected Y93H<br />

mutation by PCR- invader assay but failed to be detected by<br />

direct sequencing and the percent of the Y93H mutant were 4,<br />

5, 14, 30, 52, and 83, respectively. The difference in detection<br />

rate of Y93H mutation between two methods indicated that<br />

direct sequencing analysis could not measure less than 20%<br />

of minor strains. Nine cases were detected Y93H mutation by<br />

direct sequencing but failed to be detected by PCR- invader<br />

assay. One was Y93C, 3 were Y93H, and 5 were mixed type.<br />

The patients without Y93H mutation by both methods were<br />

achieved to SVR except for one case (97.1%). The other virologic<br />

failure were occurred in patients with Y93H mutation by<br />

either method. CONCLUSIONS: The sensitivity of PCR- invader<br />

assay was as same as that of direct sequencing analysis, but<br />

PCR- invader assay could measure minor strains not possible by<br />

direct sequencing analysis. A combination of direct sequencing<br />

and PCR- invader assay would improve prediction of the<br />

response to DCV and ASV therapy.<br />

Disclosures:<br />

Hidemi Goto - Grant/Research Support: MSD, Roche, Bayer, Bristol-Myers, Eisai,<br />

Ajinomoto, Otsuka, Astra, Tanabe, Takeda<br />

The following authors have nothing to disclose: Kazuhiko Hayashi, Masatoshi<br />

Ishigami, Yoji Ishizu, Teiji Kuzuya, Takashi Honda, Yoshiki Hirooka<br />

1062<br />

Prediction for poor virological response and clinical<br />

utility of resistance-associated variant detection in combination<br />

therapy with direct-acting antivirals for HCV<br />

genotype 1<br />

Norio Akuta, Fumitaka Suzuki, Yusuke Kawamura, Hitomi Sezaki,<br />

Yoshiyuki Suzuki, Tetsuya Hosaka, Masahiro Kobayashi, Satoshi<br />

Saitoh, Mariko Kobayashi, Yasuji Arase, Kenji Ikeda, Hiromitsu<br />

Kumada; Hepatology, Toranomon Hospital, Tokyo, Japan<br />

BACKGROUND: All-oral combinations of direct-acting antivirals<br />

improved treatment efficacy for patients with HCV infection,<br />

but prediction of poor virological response and clinical utility<br />

of resistance-associated variant (RAV) detection are not clear.<br />

METHODS: 405 interferon-ineligible/intolerant and 340 nonresponder<br />

patients with chronic HCV genotype 1b infection<br />

introduced daclatasvir 60 mg once daily plus asunaprevir 100<br />

mg twice daily for 24 weeks. 44 patients, who failed to triple<br />

therapy of peginterferon plus ribavirin with NS3/4A protease<br />

inhibitor (27 patients of telaprevir, 15 of simeprevir, and 2 of<br />

simeprevir after telaprevir), were included. Sustained virologic<br />

response 24 weeks after treatment (SVR24) and non-virological<br />

response (patients who do not achieve HCV-RNA negativity<br />

during treatment) were evaluated. Patients with virological failure<br />

were tested for RAVs (NS3 substitutions of Q80, D168,<br />

and NS5A substitutions of L31, Y93) by direct sequencing at<br />

the baseline. NS5A-Y93 mutant at baseline was also measured<br />

by comparative quantitative analysis of the Q-Invader assay.<br />

RESULTS: SVR24 was evaluated in 62 patients, and the rates


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 731A<br />

were 71% of interferon-ineligible/intolerant patients and 89%<br />

of prior non-response patients. Non-virological response was<br />

evaluated in 583 patients, and the rates were 3%. The higher<br />

levels of NS5A-Y93 mutant affected the higher rates of non-virological<br />

response, significantly. Especially, in patients of IL28B<br />

rs8099917 non TT with NS5A-Y93 mutant levels 16 on<br />

CES-D) recorded in 44%. Average medication adherence was<br />

97%. Among patients who used drugs in the past month and<br />

year, adherence was 92% and 96%, respectively. Medication<br />

adherence in those who drank alcohol in past year was 97%.<br />

Nonadherence was due to missing doses (67%) and potential<br />

double-dosing (33%). Clinic attendance to scheduled visits was<br />

95% (range: 87% - 100%). All 32 patients (100%) completed<br />

the full duration of 12 weeks of therapy. At 12-week EOT visit,<br />

97% had HCV RNA below the level of quantitation. Conclusion:<br />

Rates of medication adherence, clinic attendance, and<br />

persistence are extremely high in patients with SUDs/AUDs.<br />

These data provide no justification for withholding treatment<br />

from these patients due to adherence concerns. Clinicians<br />

and patients need to be aware of accidental double-dosing.<br />

Research is needed to determine the extent to which financial<br />

incentives are needed to optimize adherence and to identify a<br />

new adherence threshold that may jeopardize viral cure.<br />

Disclosures:<br />

The following authors have nothing to disclose: Donna M. Evon, Angela Edwards,<br />

Becky Straub, Christopher B. Hurt, Harsha Thirumurthy, David Wohl<br />

1064<br />

Frequent Emergences of Rare RAVs Showing Extreme<br />

Resistance to NS5A Inhibitors during Dual Oral Threpy<br />

with Daclatasvir Plus Asunaprevir in Patients Previously<br />

Receiving Triple Therapy with Simeprevir<br />

Yoshihito Uchida, Nobuaki Nakayama, Jun-ichi Kouyama, Kayoko<br />

Naiki, Hayato Uemura, Kayoko Sugawara, Mie Inao, Satoshi<br />

Mochida; Department of Gastroenterology and Hepatology,<br />

Saitama Medical University, Iruma-gun, Saitama, Japan<br />

Aim: Dual oral therapy with daclatasvir (DCV) plus asunaprevir<br />

(ASV) was approved in Japan for patients with genotype 1b<br />

HCV in July 2014. We established a novel system to quantify<br />

NS5A-RAVs using cycling-probe real-time PCR (PLos One<br />

2014), and performed the therapy by priority for these without<br />

baseline RAVs. However, rare RAVs showing extreme resistance<br />

to NS5A inhibitors developed frequently during DCV/<br />

ASV administrations especially in those with preveious triple<br />

therapy with simeprevir (SMV) despite that baseline RAVs were<br />

absent. We report on this serious problem. Methods: A total<br />

of 224 patients (98 men. 126 women; 23 to 87 years old.)<br />

received DCV/ASV for 24 weeks. Among them, 5 patients had<br />

a history of virologic failure after combined SMV/Peg-IFN/<br />

ribavirin therapy. RAVs in NS3 and NS5A regions at baseline<br />

and during the therapies were evaluated by cycling-probe realtime<br />

PCR combined with direct sequencing. Results: Baseline<br />

Y93H and L31M mutations were found in 27 (12%) and 6<br />

(2%) patients, respectively, while those with both mutations<br />

were absent. The therapy has already been completed in 156<br />

patients, and SVR4-12 was achieved in 134 patients (86%);<br />

117 (91%) of 129 without NS5A-RAVs, 11 (52%) of 21 with<br />

Y93H mutation and all of 6 with L31M mutations. Baseline<br />

NS5A-RAVs were absent in 5 patients with previous therapies<br />

including SMV, but virologic failure developed in all of them;<br />

DCV/ASV was discontinued at 4 weeks in 2 due to null-response,<br />

at 8 weeks in 2 due to virologic rebound and at 20<br />

weeks in 1 due to breakthrough. Among them, both Y93H and<br />

L31M mutations were detected only in 1 patient showing virologic<br />

rebound, while were absent in the remaining 4 patients.


732A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

In these patients, rare mutations in NS5A region, R30H, A92K,<br />

P29del and P32del, were detected simultaneously with NS3-<br />

D168E/V mutations in HCV straines obtained at virologic failure.<br />

Especially in 2 patients showing null-response, P29del<br />

and P32del mutations were found, and experiments using replicon<br />

system revealed HCV strains with such deletions showed<br />

>390,000-fold registance against DCV. Consequently, SVR4-<br />

12 rates in patients without baseline Y93H mutation were 95%<br />

when those receiving SMV were excluded. Deep-sequencing<br />

to evaluate the kinetics of HCV strains with P29del and P32del<br />

was currently in progress. Conclusion: Although high SVR rates<br />

were obtained after DCV/ASV therapy in patients without<br />

baseline NS5A-RAVs, rare RAVs extremely resistant to NS5A<br />

inhibitors emerged frequently in those receiving SMV previously.<br />

Such patients should be add to the lists of contraindication<br />

for DCV/ASV therapy even when baseline NS5A-RAVs<br />

were absent.<br />

Disclosures:<br />

Yoshihito Uchida - Patent Held/Filed: SRL Inc.<br />

Jun-ichi Kouyama - Patent Held/Filed: SRL Inc.<br />

Kayoko Naiki - Patent Held/Filed: SRL Inc.<br />

Satoshi Mochida - Grant/Research Support: Chugai, MSD, Tioray Medical,<br />

BMS; Speaking and Teaching: MSD, Toray Medical, BMS, Tanabe Mitsubishi<br />

The following authors have nothing to disclose: Nobuaki Nakayama, Hayato<br />

Uemura, Kayoko Sugawara, Mie Inao<br />

pts without cirrhosis received 12 weeks (wks) of 3D+RBV and<br />

cirrhotics received 12 or 24 wks of 3D+RBV. GT1b-infected<br />

pts without cirrhosis received 12 wks of 3D and cirrhotics<br />

received 12 wks of 3D+RBV. End of treatment response (EOTR)<br />

is reported; SVR12 will be presented. Results: 615 pts enrolled<br />

and received study drug. Baseline characteristics and treatment<br />

response are presented below. 98.0% (603/615) of pts<br />

achieved EOTR. 0.8% (5/615) of pts experienced on-treatment<br />

virologic failure, and 1.0% (6/615) discontinued treatment<br />

due to adverse events. The most common adverse events<br />

(occurring in >10% of pts) were fatigue, nausea, headache,<br />

insomnia, and pruritus. Serious adverse events considered at<br />

least possibly related to 3D treatment were reported in 0.5%<br />

(3/615) of pts. Conclusions: In the TOPAZ-II study, 3D+/-RBV<br />

achieved high rates of on-treatment response and has been<br />

well-tolerated with a low rate of discontinuation due to adverse<br />

events in a broad population of HCV GT1-infected pts.<br />

Baseline characteristics and efficacy<br />

1065<br />

Preliminary Safety and Efficacy Results from TOPAZ-II:<br />

A Phase 3b Study Evaluating Long-Term Clinical Outcomes<br />

in HCV Genotype 1-infected Patients Receiving<br />

Ombitasvir/Paritaprevir/r and Dasabuvir +/-Ribavirin<br />

Nancy Reau 1 , Fred Poordad 2 , Jeffrey Enejosa 3 , Asma Siddique 4 ,<br />

Humberto I. Aguilar 5 , Jacob P. Lalezari 6 , Franco Felizarta 7 , Peter<br />

J. Ruane 8 , Peter Varunok 9 , David Bernstein 10 , Douglas Dieterich 11 ,<br />

Gregory T. Everson 12 , Yiran Hu 3 , Greg Ball 3 , Tami Pilot-Matias 3 ,<br />

Nancy Shulman 3 , Sanjeev Arora 13 ; 1 University of Chicago Medical<br />

Center, Chicago, IL; 2 The Texas Liver Institute/University of<br />

Texas Health Science Center, San Antonio, TX; 3 AbbVie Inc.,<br />

North Chicago, IL; 4 Virginia Mason Medical Center, Seattle, WA;<br />

5 Louisiana Research Center, LLC, Shreveport, LA; 6 Quest Clinical<br />

Research, San Francisco, CA; 7 Private practice, Bakersfield, CA;<br />

8 Ruane Medical and Liver Health Institute, Los Angeles, CA; 9 Premier<br />

Medical Group of the Hudson Valley, PC, Poughkeepsie, NY;<br />

10 North Shore University Hospital, Manhasset, NY; 11 Mount Sinai<br />

School of Medicine, Icahn School of Medicine, New York, NY;<br />

12 University of Colorado Denver School of Medicine, Aurora, CO;<br />

13 University of New Mexico, Albuquerque, NM<br />

Background: The 3-direct-acting antiviral regimen of ombitasvir/paritaprevir/ritonavir<br />

(paritaprevir identified by AbbVie<br />

and Enanta, co-dosed with ritonavir) and dasabuvir (3D) +/-ribavirin<br />

(RBV) has resulted in high SVR12 rates with a favorable<br />

safety profile in HCV genotype (GT) 1-infected patients (pts),<br />

including those with compensated cirrhosis, in phase 3 trials.<br />

TOPAZ-I (ex-US) and TOPAZ-II (US) are evaluating the impact<br />

of SVR12 on progression of liver disease through 5 years<br />

post-treatment in a broad population of HCV GT1-infected pts<br />

receiving 3D+/-RBV. Here, we report on-treatment safety and<br />

efficacy of 3D+/-RBV among pts in the TOPAZ-II study. Methods:<br />

TOPAZ–II is a phase 3b, open-label, non-randomized,<br />

multicenter trial conducted in 48 community and academic<br />

centers in the US. Pts are treatment-naïve or interferon/RBV<br />

treatment-experienced, with or without compensated cirrhosis.<br />

There was no upper limit on age, BMI, or ALT/AST levels. Pts<br />

with creatinine clearance ≥30 mL/min, albumin ≥2.8 g/dL,<br />

and platelet count ≥25 × 10 9 /L were eligible. GT1a-infected<br />

*Data missing for 1 pt<br />

Disclosures:<br />

Nancy Reau - Advisory Committees or Review Panels: Jannsen, Merck, AbbVie,<br />

Intercept, Salix, BMS, Jannsen; Grant/Research Support: Merck, Gilead, BMS,<br />

AbbVie, Jannsen, BI<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Jeffrey Enejosa - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Humberto I. Aguilar - Speaking and Teaching: abbvie, gilead<br />

Franco Felizarta - Grant/Research Support: AbbVie, Gilead, Janssen, Merck,<br />

BMS, Boehringer-Ingelheim, Vertex, Roche; Speaking and Teaching: AbbVie,<br />

Gilead, Janssen, Merck<br />

Peter J. Ruane - Advisory Committees or Review Panels: Gilead, Boehringer,<br />

Jannsen, Viiv, Abbott; Consulting: Gilead, Jannsen, Abbott, BMS; Grant/<br />

Research Support: Gilead, Boehringer, Jannsen, Idenix, Abbott, BMS; Speaking<br />

and Teaching: Gilead, Merck, Boehringer, Jannsen, VIIV, Abbott, BMS; Stock<br />

Shareholder: Gilead<br />

David Bernstein - Advisory Committees or Review Panels: Gilead; Consulting:<br />

Abbvie, BMS, Merck, Janssen; Grant/Research Support: Gilead, Abbvie, BMS,<br />

Merck, Janssen, Genentech; Speaking and Teaching: Abbvie, BMS, Merck,<br />

Gilead<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Gregory T. Everson - Advisory Committees or Review Panels: Roche/Genentech,<br />

Abbvie, Galectin, Boehringer-Ingelheim, Eisai, Bristol-Myers Squibb, HepC<br />

Connection, BioTest, Gilead, Merck; Board Membership: HepQuant LLC, PSC<br />

Partners, HepQuant LLC; Consulting: Abbvie, BMS, Gilead, Bristol-Myers Squibb;<br />

Grant/Research Support: Roche/Genentech, Pharmassett, Vertex, Abbvie, Bristol-Myers<br />

Squibb, Merck, Eisai, Conatus, PSC Partners, Vertex, Tibotec, GlobeImmune,<br />

Pfizer, Gilead; Management Position: HepQuant LLC, HepQuant LLC;<br />

Patent Held/Filed: Univ of Colorado; Speaking and Teaching: Abbvie, Gilead<br />

Yiran Hu - Employment: AbbVie Inc.<br />

Tami Pilot-Matias - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Nancy Shulman - Employment: Abbvie<br />

The following authors have nothing to disclose: Asma Siddique, Jacob P. Lalezari,<br />

Peter Varunok, Greg Ball, Sanjeev Arora


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 733A<br />

1066<br />

Effect of Renal Function on the Pharmacokinetics of<br />

Ombitasvir/ Paritaprevir/Ritonavir, Dasabuvir and Ribavirin<br />

in Over 2000 Subjects with HCV GT1 Infection<br />

Akshanth R. Polepally 1 , Prajakta Badri 1 , Doerthe Eckert 2 , Sven<br />

Mensing 2 , Rajeev Menon 1 ; 1 Clinical Pharmacology and Pharmacometrics,<br />

AbbVie Inc.,, North Chicago, IL; 2 Clinical Pharmacology<br />

and Pharmacometrics, AbbVie Deutschland GmbH & Co.KG,<br />

Ludwigshafen, Germany<br />

Background and aims: Paritaprevir, a NS3/4A protease inhibitor<br />

(administered with low-dose ritonavir [RTV], paritaprevir/r),<br />

identified by AbbVie and Enanta, ombitasvir, a NS5A inhibitor,<br />

and dasabuvir, a NS5B polymerase inhibitor, are directly<br />

acting antivirals (DAAs) approved in combination (3D regimen)<br />

± ribavirin (RBV) for the treatment of HCV GT1 infection . The<br />

objective of this analysis was to evaluate the effect of renal function<br />

as estimated by creatinine clearance (CrCL) on the pharmacokinetics<br />

of the DAAs, RTV and RBV in HCV infected GT1<br />

subjects. Methods: Total exposure measured by area under the<br />

plasma concentration curve (AUC) was generated for DAAs,<br />

RTV and RBV using population pharmacokinetic modeling by<br />

pooling data from 6 phase 3 <strong>studies</strong> and 1 phase 2 study in ><br />

2000 HCV GT1 infected subjects. All subjects received ombitasvir/<br />

paritaprevir/r 25/150/100 mg QD and dasabuvir<br />

250 mg BID ± weight based RBV. DAA and RTV AUC values<br />

were available from 2093 subjects and RBV AUC values were<br />

available from 1584 subjects. The dataset included subjects<br />

with normal renal function (NF) (CrCl ≥ 90 mL/min, n = 1495),<br />

mild renal impairment (RI) (CrCL 60-89 mL/min, n = 576) and<br />

moderate RI (CrCL 30-59 mL/min, n = 22). The effect of CrCL<br />

on the AUC values of each DAA, RTV and RBV was evaluated,<br />

and adjusted for any significant subject-specific covariates (at<br />

a significance level of 0.05) including, age, sex, body weight<br />

(BW), cirrhosis (CRHS) and Asian ethnicity (ASN) in multiple<br />

linear regression (MLR) analysis (R 3.2.0). CrCL was retained<br />

in the models, regardless of its statistical significance, to determine<br />

the effect, if any, on the AUC values. Using the final<br />

MLR model, AUC values were predicted for subjects with NF<br />

(CrCL=105 mL/min), mild RI (CrCL=75 mL/min) and moderate<br />

RI (CrCL=45 mL/min). Results: CrCL was not a statistically<br />

significant predictor of DAAs and RTV AUC values (p > 0.05).<br />

Age, sex, CRHS were significant covariates for all DAAs/RTV<br />

while BW and ASN were for ombitasvir and dasabuvir. CrCL<br />

showed a significant relation with the RBVAUC values (p <<br />

0.05), which is consistent with RBV’s predominant renal excretion.<br />

Age, sex, BW and CRHS were significant covariates for<br />

RBV. The DAA AUC values were comparable (≤ 10% difference)<br />

amongst different levels of renal function, while RBV AUC<br />

values were up to 17% higher in mild/moderate RI compared<br />

to NF. Conclusions: In HCV GT1-infected subjects with or without<br />

cirrhosis, mild/moderate RI did not affect DAA and RTV<br />

exposures; thus, no dose-adjustments are needed for the 3D<br />

regimen. RBV doses should be adjusted for renal impairment<br />

as recommended in its label.<br />

Disclosures:<br />

Akshanth R. Polepally - Employment: AbbVie<br />

Prajakta Badri - Employment: Abbvie; Stock Shareholder: Abbvie<br />

Sven Mensing - Employment: AbbVie<br />

Rajeev Menon - Employment: AbbVie; Stock Shareholder: AbbVie<br />

The following authors have nothing to disclose: Doerthe Eckert<br />

1067<br />

Clinical Management of Ribavirin Dosing in HCV-Infected<br />

Patients with Anemia-Related Events Receiving<br />

Ombitasvir/Paritaprevir/r and Dasabuvir<br />

Jordan J. Feld 2 , David Bernstein 3 , Ziad Younes 4 , Hans Van Vlierberghe<br />

5 , Greg Ball 1 , Ronald D’Amico 1 , Peter Ferenci 6 ; 1 AbbVie,<br />

Inc., North Chicago, IL; 2 Toronto Centre for Liver Disease, University<br />

of Toronto, Toronto, ON, Canada; 3 North Shore University<br />

Hospital, Manhasset, NY; 4 GastroOne, Germantown, TN; 5 Ghent<br />

University Hospital, Gent, Belgium; 6 Medical University of Vienna,<br />

Vienna, Austria<br />

Objective: Some individuals infected with HCV genotype 1<br />

treated with the 3 direct-acting antiviral (3D) regimen of ombitasvir<br />

(OBV), paritaprevir (PTV, identified by AbbVie and Enanta),<br />

and dasabuvir (DSV) require ribavirin (RBV) to optimize the<br />

chance of achieving a sustained virologic response (SVR). We<br />

describe the management of anemia-related adverse events<br />

(AEs) due to RBV-associated hemolysis in phase 3 trials for<br />

patients receiving the 3D + RBV regimen. Methods: Data were<br />

pooled from the SAPPHIRE-I and -II, PEARL-II, -III, and -IV, and<br />

TURQUOISE-II <strong>studies</strong> including patients with HCV genotype<br />

1a (N = 856) and 1b (N = 691) infection. Patients without cirrhosis<br />

received the 3D + RBV regimen for 12 weeks, whereas<br />

those with cirrhosis received 12 or 24 weeks treatment. Ribavirin<br />

was initially dosed according to body weight with a total<br />

daily dose of 1000 mg if


734A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Ronald D’Amico - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Peter Ferenci - Advisory Committees or Review Panels: Idenix, Gilead, MSD,<br />

Janssen, Salix, AbbVie, BMS; Patent Held/Filed: Madaus Rottapharm; Speaking<br />

and Teaching: Gilead, Roche<br />

The following authors have nothing to disclose: Hans Van Vlierberghe, Greg Ball<br />

1068<br />

The healthcare cost burden of HCV-infected baby boomers<br />

in the US<br />

Ron Brookmeyer 2 , Timothy R. Juday 1 , Darius Lakdawalla 3,4 , Steven<br />

Marx 1 , Gigi Moreno 4 , Yuri Sanchez 1 ; 1 HEOR, AbbVie, Mettawa,<br />

ID; 2 David Geffen School of Medicine at UCLA, Los Angeles, CA;<br />

3 USC, Los Angeles, CA; 4 Precision Health Economics, Los Angeles,<br />

CA<br />

Background While the incidence of hepatitis-C virus (HCV)<br />

infection peaked more than 25 years ago, the untreated<br />

infected population is aging and experiencing expensive liver<br />

complications, raising questions about future financial burdens<br />

for Medicare. To shed light on this, we quantified healthcare<br />

costs for 50-64 year-old “baby boomers” with HCV by diagnosis<br />

and insurance status. Methods This study conducted a<br />

literature review and analyzed five waves of National Health<br />

and Nutrition Expenditures Survey (NHANES) data from 2003-<br />

2004 through 2011-2012. Insurance, HCV infection, and<br />

diagnosis status were obtained from NHANES. Respondents<br />

were identified as diagnosed if they tested positive for HCV,<br />

responded to the follow-up interview, and reported awareness<br />

of their HCV infection prior to NHANES. The analysis focused<br />

on respondents with HCV RNA, indicating chronic infection.<br />

The final sample contained 28,607 respondents. Healthcare<br />

costs for chronic HCV reported in the literature were applied<br />

to estimate total HCV costs. Results Among the 2.7 million<br />

Americans with chronic HCV, approximately 1 million are<br />

aged 50–64. Of these, about 319,003 are undiagnosed,<br />

and 161,609 are diagnosed but uninsured. Privately insured<br />

chronic HCV patients cost $17,176 annually if suffering from<br />

noncirrhotic disease (NCD), $22,752 if suffering from compensated<br />

cirrhosis (CC), and $59,995 if suffering from endstage<br />

liver disease (ESLD). This implies annual costs for the<br />

diagnosed, insured population of roughly $11.12 billion. If<br />

the diagnosed uninsured were treated like insured patients,<br />

this would add $3.23 billion. If the undiagnosed were diagnosed<br />

and treated like insured patients, this would add another<br />

$5.48 billion. Conclusions As baby boomers with HCV enter<br />

Medicare, the program may face substantial cost burdens.<br />

Proactively treating HCV patients before they suffer expensive<br />

long-term complications can help mitigate this burden to the US<br />

government.<br />

Gigi Moreno - Consulting: AbbVie<br />

Yuri Sanchez - Employment: AbbVie; Stock Shareholder: AbbVie<br />

1069<br />

An Actuarial View of the Aging Hepatitis C Virus (HCV)<br />

Epidemic<br />

Gabriela Dieguez 2 , Bruce Pyenson 2 , Ryan Cannon 2 , Steven<br />

Marx 1 , Yuri Sanchez 1 , Timothy R. Juday 1 ; 1 HEOR, AbbVie, Mettawa,<br />

ID; 2 Milliman, New York, NY<br />

PURPOSE In the coming years, the aging and treatment dynamics<br />

of people with chronic HCV infection may substantially shift<br />

the financial burden of health insurance coverage from private<br />

payers to Medicare. We examined shifts in the distribution of<br />

the HCV-infected population from 2015-2025 by US health<br />

insurance market, with a focus on people aging into Medicare<br />

and assuming that no patients would be cured of HCV infection.<br />

METHODS A stochastic actuarial model was developed to forecast<br />

the annual number of people with HCV insured by Medicare,<br />

Medicaid, commercial (including self-insured employers<br />

and Health Insurance Exchange enrollees), and military-associated<br />

insurance (including the Veteran Administration), as well<br />

as the number of uninsured HCV patients from 2015 to 2025.<br />

The current number of HCV patients in the major insurance<br />

markets and their annual probabilities of mortality, disability,<br />

and disease progression by age, gender, income, and HCV<br />

stage were estimated using data from Truven Health Market-<br />

Scan, Medicare 5% Sample, National Health and Nutrition<br />

Examination Survey, and Census Bureau. The expected impact<br />

of the Affordable Care Act on the HCV population’s health<br />

insurance status was considered. Prisoners were excluded from<br />

the analysis. Based on the HCV population forecast, the number<br />

of people with HCV who will gain Medicare eligibility<br />

through aging or disability in the next 10 years was calculated.<br />

RESULTS About 2.5 million non-institutionalized people in the<br />

US are estimated to be infected with HCV today. The majority<br />

of these patients are under age 65 with only 16% (405,000)<br />

covered by Medicare. This forecast predicts that, by 2025,<br />

38% (591,000) of people with HCV in the US are expected<br />

to be eligible for Medicare, and the number of HCV patients<br />

with commercial, Medicare, or military insurance is expected<br />

to decline to less than half of the current levels. In the next 10<br />

years, 635,000 more people with HCV will gain Medicare<br />

coverage. Furthermore, 921,000 people with HCV will die if<br />

left untreated. CONCLUSION Due to both age- and disease-related<br />

mortality, the HCV-infected population will decrease by<br />

almost 1 million over the next 10 years, assuming no treatment.<br />

By 2020, Medicare will displace commercial insurers as the<br />

largest payer for people who are currently infected with HCV.<br />

Sources: NHANES, Di Bisceglie, 2000, and Gordon 2012.<br />

Disclosures:<br />

Ron Brookmeyer - Consulting: Precision Health Economics<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Darius Lakdawalla - Consulting: AbbVie Pharmaceuticals; Stock Shareholder:<br />

Precision Health Economics<br />

Steven Marx - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Disclosures:<br />

Gabriela Dieguez - Consulting: Abbvie<br />

Bruce Pyenson - Consulting: AbbVie, which funded this research, Employed by<br />

Milliman, Inc., which consults to numerous payer, providers, and suppliers to the<br />

healthcare industry, PhRMA<br />

Ryan Cannon - Consulting: Abbvie, Milliman, Inc., PhRMA<br />

Steven Marx - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Yuri Sanchez - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 735A<br />

1070<br />

Early and persistent increase of serum cholesterol and<br />

low-density lipoprotein in chronic hepatitis C patients<br />

during IFN-free treatment<br />

Clarissa Freissmuth, Albert Stättermayer, Karin Kozbial, Sandra<br />

Beinhardt, Rafael Stern, Michael Eder, Petra E. Steindl-Munda,<br />

Michael Trauner, Peter Ferenci, Harald Hofer; Department of Internal<br />

Medicine III, Division of Gastroenterology and Hepatology,<br />

Medical University of Vienna, Vienna, Austria<br />

Background&Aims: Hepatitis C virus (HCV) is closely related<br />

to lipid metabolism. HCV genotype-(GT)-3 infection is associated<br />

with hepatic steatosis and low levels of serum cholesterol,<br />

which are reversible after successful HCV treatment. Moreover,<br />

in HCV GT-1 infection serum levels of low-density lipoprotein<br />

(LDL) particles increase on HCV clearance during sofosbuvir/<br />

ribavirin (SOF/RBV) therapy. This study investigated serum lipid<br />

parameters in patients with HCV-1 and 3 infection undergoing<br />

IFN-free treatment. Patients&Methods: Total serum cholesterol,<br />

high density lipoprotein (HDL), LDL and serum triglycerides (TG)<br />

were analyzed at baseline, during (initially weekly, afterwards<br />

every four weeks) and after antiviral therapy in 208 patients<br />

(GT-1: N=163, age: 57±11, BMI: 26±4kg/m 2 , male/female:<br />

99/64, F4: 108 (66%), Diabetes mellitus (DM): 35(21%)),<br />

(GT-3: N=45, age: 52±7, BMI: 26±4kg/m 2 , m/f: 32/13, F4:<br />

34 (76%), DM: 6 (13%)). Patients with GT-3 received SOF,<br />

400mg/d with RBV (N=24) or Daclatasvir (60mg/d) (DCV,<br />

N=21). Patients with GT-1 received SOF in combination with<br />

DCV (N=85), Simeprevir 150mg/d (N=56), 90mg/d Ledipasvir<br />

(N=20) or RBV (N=2). Results: Overall, serum cholesterol<br />

and LDL increased during treatment (table 1). This was<br />

observed after the first week of treatment (chol BL<br />

149±36<br />

vs. chol W1<br />

162±41 vs. chol W2<br />

169±40 mg/dL, mean±SD,<br />

p


736A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

results must be confirmed by analysis of the SVR12. Full result<br />

analysis will be presented at the congress. * Same contribution<br />

for the 2 first authors<br />

Disclosures:<br />

Eric Nguyen-Khac - Speaking and Teaching: Gilead, Abbvie, Janssen, Roche,<br />

MSD, BMS<br />

Alexandre Pariente - Board Membership: Mayoli spindler; Speaking and Teaching:<br />

Janssen, Roche, BMS, Abbvie<br />

Thong Dao - Board Membership: Schering-Plough, Schering-Plough, Schering-Plough,<br />

Schering-Plough; Speaking and Teaching: Roche, Gilead, Roche,<br />

Gilead, Roche, Gilead, Roche, Gilead<br />

Alexandre Pariente - Board Membership: Mayoli spindler; Speaking and Teaching:<br />

Janssen, Roche, BMS, Abbvie<br />

Patrick Delasalle - Speaking and Teaching: ROCHE, AXCAN, BMS, GILEAD,<br />

MERCK, JANSSEN, ROCHE, AXCAN, BMS, GILEAD, MERCK, JANSSEN<br />

Xavier Causse - Board Membership: Gilead, Janssen-Cilag; Grant/Research<br />

Support: MSD; Speaking and Teaching: Gilead, BMS, Janssen-Cilag<br />

Francois Bourhis - Consulting: ABBVIE, GILEAD; Speaking and Teaching: BMS<br />

Alexandra Heurgué-Berlot - Consulting: Abbvie; Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Bruno Lesgourgues, Remy Andre,<br />

Brigitte Bernard-Chabert, Isabelle Rosa, Damien Lucidarme, Christophe Renou,<br />

Christine Silvain, Gilles Macaigne, Thierry Fontanges, Matthieu Schnee, Hortensia<br />

Lison, Christophe Pilette, Jean Pierre Arput, Frédéric Heluwaert<br />

1073<br />

Safety And Efficacy Of The Combination Simeprevir-Sofosbuvir<br />

In HCV Genotype 1- And 4-Mono-Infected<br />

Patients From The French Observational Cohort Anrs<br />

Co22 Hepather*<br />

ANRS/AFEF HEPATHER CO22 study group; National Agency for<br />

Research on AIDS and Viral Hepatitis, Paris, France<br />

Background and aims. Real-life results of the Sofosbuvir/Simeprevir<br />

combination have been extensively reported from USA<br />

but there are few data from other geographical areas. In May<br />

2015, more than 3500 patients of the French observational<br />

cohort ANRS CO22 HEPATHER were given the new oral antivirals<br />

in 32 centers: we report the preliminary results of the Sofosbuvir/Simeprevir<br />

combination in Genotype 1- and 4-infected<br />

patients. Methods. Demographics, history of liver disease<br />

were collected at entry in the cohort. Clinical, adverse events,<br />

and virological data were collected throughout treatment and<br />

post-treatment follow-up. Results. 552 HCV (433 genotype 1-<br />

and 119 genotype 4-) mono-infected patients were given a<br />

combination of Sofosbuvir (SOF: 400 mg/d) and Simeprevir<br />

(SMV: 150 mg/d) without Ribavirin (n=491) or with Ribavirin<br />

(1-1.2 g/d, n=61). 54% of patients had cirrhosis (3% decompensated),<br />

75% were previously treated with PR (n=315) or<br />

PR + a first generation PI (n=25). Overall, the SVR12 rate was<br />

88%. Underlying cirrhosis did not impact SVR rates. In patients<br />

who received the 12-week combination SOF/SMV +/- RBV the<br />

SVR12 rate differed significantly according to the genotype<br />

(80% in GT1a, 92% in GT1b and 86% in GT4, p=0.004) and<br />

according to the biochemical markers of liver severity -MELD<br />

score, total and conjugated bilirubin, platelet count, Prothrombin<br />

time, ALT, AST-, p50 gms/day), active cocaine or crack use Patient characteristics-<br />

table to follow on the slide/poster They were divided into<br />

2 groups; Group A (n=14): SOF 400 mg + LDV 90 mg (daily<br />

once) - 4 weeks Group B (n=15): SOF 400 mg + SIM 150 mg<br />

(daily once) - 8 weeks Labs: Prior to therapy- HCV RNA 0 and<br />

3 weeks, sickle cell panel, LFT’s, CBC, SMA On therapy- HCV<br />

RNA on day 0, 1 week, 4 week, 6 week followed by 16 th and<br />

20 th week (SVR), LFT’s, CBC, SMA Results: Table Conclusion:<br />

This study demonstrates a high SVR with short-course DAA’s in<br />

acute hepatitis C with SVR (at 20 weeks) >90% in both groups.<br />

The drugs were well tolerated.<br />

Results<br />

Disclosures:<br />

The following authors have nothing to disclose: Patrick Basu, Niraj J. Shah, Nimy<br />

John, M. Aloysius, Robert Brown


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 737A<br />

1075<br />

Sofosbuvir, Ledipasvir in IBD treated patients with<br />

advanced biologics including Ribavirin eradicating<br />

Chronic Hepatitis C: SOLATAIRE C Trial. A multi-center<br />

clinical prospective pilot study<br />

Patrick Basu 1,2 , Niraj J. Shah 3 , Nimy John 2 , M. Aloysius 2 , Robert<br />

Brown 4 ; 1 Columbia University School of Physicians and Surgeons,<br />

Forest hills, NY; 2 King’s County Hospital Medical Center, NY,<br />

New York, NY; 3 James J. Peters VA Medical Center, Icahn School<br />

of Medicine at Mount Sinai, NY, New York, NY; 4 Weill Cornell<br />

Medical College, New York, NY<br />

Background: The prevalence and concomitant treatment of<br />

chronic HCV and IBD has not been elucidated. IBD and CHC<br />

of >20 years duration may have more fibrosis due to immune<br />

suppression. Treatment of IBD often requires biological therapies<br />

to achieve longer disease-free remission, which further<br />

accelerates higher HCV replication. Thus effective therapy for<br />

IBD patients requiring immunosuppressive therapy is needed<br />

Aim: To evaluate the role and efficacy of new NS5A + NS5B<br />

inhibitors with and without RBV in treating CHC in moderate<br />

to severe IBD requiring biologic therapy Methods: 35 patients<br />

were recruited from the Kings County Hospital Medical Center,<br />

Brooklyn. Inclusion criteria CHC with IBD, Age: >18, HCV Viral<br />

Load: > 400,000IU/mL, Genotype 1. Fibrotic Score: Metavir<br />

F1 to F4 Primary End Point: SVR12 Secondary End Point: IBD<br />

activity index, sustained control of symptoms SVR 24 and complete<br />

IBD remission at 30 weeks (endoscopy at 30 weeks)<br />

The patients were divided into two groups: Group A (n=17)-<br />

RBV 1000 mg + LDV and SOF; for 8 weeks Group B (n=18):<br />

LDV and SOF; for 12 weeks All will have base line RAV and<br />

ETRAV’s and SVR 24. RAVs 5A polymorphism by Quest. Viral<br />

loads 0, 7, 14 days; 4 th , 8 th , 12 th and 24 th weeks Patient characteristics:<br />

list to follow in the poster/oral presentation Results:<br />

See table. 24 th week RAV- results are pending Conclusion: This<br />

study demonstrates that DAA’s for HCV appear to have similar<br />

efficacy and safety in the presence of active IBD therapy with-<br />

TNF Alfa antagonists while keeping the IBD in uninterrupted<br />

remission<br />

Results<br />

Disclosures:<br />

The following authors have nothing to disclose: Patrick Basu, Niraj J. Shah, Nimy<br />

John, M. Aloysius, Robert Brown<br />

1076<br />

Sofosbuvir plus Ribavirin in the Treatment of Egyptian<br />

Patients with Chronic Genotype 4 HCV Infection<br />

Wahid H. Doss 1 , Peter J. Ruane 2 , Gamal Shiha 3 , Dani Ain 2 ,<br />

Reham Soliman 3 , Marwa Khairy 4 , Mohamad Hassany 1 , Joseph<br />

Riad 2 , Waleed Samir 3 , Rabab F. Omar 4 , Radi Hammad 1 , Raymond<br />

Meshrekey 2 , Mostafa Gamil 4 , Deyuan Jiang 5 , Benedetta<br />

Massetto 5 , Steven J. Knox 5 , Kathryn Kersey 5 , John G. McHutchison<br />

5 , Gamal E. Esmat 4 ; 1 National Hepatology and Tropical Medicine<br />

Research Institute, Cairo, Egypt; 2 Ruane Medical and Liver<br />

Health Institute, Los Angeles, CA; 3 Egyptian Liver Research Institute<br />

and Hospital, Mansoura, Egypt; 4 University of Cairo, Cairo,<br />

Egypt; 5 Gilead Sciences, Inc., Foster City, CA<br />

Background: HCV infection is a major public health burden in<br />

Egypt, with an estimated 6 million people chronically infected.<br />

The combined safety and efficacy data for sofosbuvir (SOF)<br />

plus ribavirin (RBV) from 2 <strong>studies</strong> in Egyptian patients with<br />

chronic genotype (GT) 4 HCV infection are summarized here.<br />

Methods: An integrated analysis was conducted of data from<br />

treatment-naïve and treatment-experienced patients enrolled in<br />

Study GS-US-334-0114 in the USA (n=60) and Study GS-US-<br />

334-0138 in Egypt (n=103). In each study, subjects were randomized<br />

1:1 to receive either 12 or 24 weeks of SOF (400<br />

mg daily) + RBV (1000-1200 mg daily). Exact logistic regression<br />

analyses were used to explore the relationships between<br />

SVR12 rates and baseline characteristics in the combined dataset.<br />

Results: Overall, 67% of patients were male, mean age<br />

(range) was 49 years (19-75), 42% had BMI ≥30 kg/m 2 , 19%<br />

had cirrhosis, 82% had IL28B non-CC genotype, 56% had<br />

HCV RNA ≥800,000 IU/mL at baseline, and 53% were treatment-experienced.<br />

SVR12 rates were 73% (61/83) and 91%<br />

(73/80) for SOF+RBV administered for 12 and 24 weeks,<br />

respectively. Multivariate logistic regression analysis identified<br />

male sex, IL28B non-CC genotype, and baseline HCV RNA<br />

≥800,000 IU/mL as statistically significant factors associated<br />

with worse treatment outcome while 24 weeks of treatment<br />

increased the likelihood of achieving SVR12. In treatment-naïve<br />

patients without cirrhosis, SOF+RBV for 12 or 24 weeks<br />

resulted in SVR12 rates of 88% (29/33) and 94% (30/32),<br />

respectively. The most common adverse events were headache,<br />

fatigue, insomnia, cough and pruritus. Most AEs were mild or<br />

moderate in severity and none resulted in treatment discontinuation;<br />

5 patients experienced SAEs of which 1 (dyspnea)<br />

was considered treatment-related. Conclusions: Integrated data<br />

from these 2 <strong>studies</strong> show SOF+RBV was well tolerated with<br />

an AE profile consistent with that of RBV. SOF+RBV for 24<br />

weeks resulted in a 91% SVR12 rate in Egyptian patients with<br />

HCV GT4 infection. The combined safety and efficacy data<br />

support the use of this interferon-free regimen. In addition, the<br />

data suggest a shorter treatment duration of 12 weeks may be<br />

of benefit for some patients, such as treatment naïve patients<br />

without cirrhosis.<br />

Disclosures:<br />

Peter J. Ruane - Advisory Committees or Review Panels: BMS; Consulting: Gilead,<br />

Abbvie, Janssen; Grant/Research Support: BMS, Gilead, Merck, Abbvie, Idenix,<br />

Idenix, Janssen, Viiv; Speaking and Teaching: Gilead, Merck, Abbvie, Abbvie,<br />

Janssen; Stock Shareholder: Gilead, Gilead<br />

Mohamad Hassany - Grant/Research Support: Gilead Sc<br />

Radi Hammad - Grant/Research Support: Gilead Sciences,Inc, Janssen Pharmaceuticals,<br />

Inc.<br />

Deyuan Jiang - Employment: Gilead Sciences<br />

Benedetta Massetto - Employment: Gilead Sciences, Inc.; Stock Shareholder:<br />

Gilead Sciences, Inc<br />

Steven J. Knox - Employment: Gilead Sciences<br />

Kathryn Kersey - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc


738A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Gamal E. Esmat - Advisory Committees or Review Panels: MSD &BMS companies,<br />

MSD &BMS companies, Abbvie; Grant/Research Support: Gilead Sc;<br />

Speaking and Teaching: Roche & GSK companies, Roche & GSK companies<br />

The following authors have nothing to disclose: Wahid H. Doss, Gamal Shiha,<br />

Dani Ain, Reham Soliman, Marwa Khairy, Joseph Riad, Waleed Samir, Rabab<br />

F. Omar, Raymond Meshrekey, Mostafa Gamil<br />

1077<br />

The TOSCAR Study: Sofosbuvir and Daclatasvir Therapy<br />

for Decompensated HCV Cirrhosis with MELD scores ≥<br />

15: What is the point of no return?<br />

Geoff McCaughan 1 , Stuart K. Roberts 2 , Simone I. Strasser 1 , Paul<br />

Gow 3 , Alan J. Wigg 4 , Caroline Tallis 5 , Gary P. Jeffrey 6 , Jacob<br />

George 7 , Alex J. Thompson 8 , Susan Mason 9 , Joanne Mitchell 2 ,<br />

Peter W. Angus 3 ; 1 ANLTU, Royal Prince Alfred Hospital/Univeristy<br />

of Sydney, Sydney, NSW, Australia; 2 Alfred Health, Melbourne,<br />

VIC, Australia; 3 VLTU, Melbourne, VIC, Australia; 4 SALTU, Adelaide,<br />

SA, Australia; 5 QLTU, Brisbane, QLD, Australia; 6 WALTU,<br />

Perth, WA, Australia; 7 Westmead Hospital, Sydney, NSW, Australia;<br />

8 St Vincent’s Hospital, Melbourne, VIC, Australia; 9 AW Morrow<br />

Gastroenterology & Liver Centre, Royal Prince Alfred Hospital,<br />

Sydney, NSW, Australia<br />

Background: Interferon free therapies for chronic HCV have<br />

resulted in greater than 95% SVRs. However, there are few<br />

data that examine the efficacy in patients with decompensated<br />

chronic liver disease due to HCV. We report such data in a<br />

well defined patient population. Aims: To examine virological<br />

responses in patients with decompensated HCV liver disease<br />

treated with an interferon free regime To examine the effect of<br />

viral clearance on severity and outcomes of liver disease. Methods:<br />

A compassionate access Australia-wide protocol using<br />

Sofosbuvir and Daclatasvir for 24 weeks without Ribavirin to<br />

treat patients with HCV decompensated liver disease with a<br />

MELD score of ≥ 15 was introduced in November 2014. Data<br />

was collected from multiple sites at tertiary level liver centres<br />

throughout Australia. Results: 92 patients have been analysed.<br />

The average age was 55 years. 73% were male. 39% had<br />

genotype 1a, 9% genotype1b and 27% genotype 3. The average<br />

Child Pugh Score (CPS) was 9.5 with an average MELD<br />

of 17.4. 23 patients (25%) had MELD scores > 20 (maximum<br />

29). 76% of patients had ascites and 65% had encephalopathy.<br />

35 patients were listed and 10 underwent liver transplantation<br />

during this period .6 patients died before transplantation<br />

(5 within 8 weeks of starting treatment). End of treatment (EOT)<br />

data was available in 15 patients. All but 1 patient was HCV<br />

PCR negative. The single patient who remained positive had<br />

undergone a surgical procedure with treatment disruption. This<br />

patient had also had a previous relapse from a 12 week course<br />

of Sofosbuvir and Ledipasvir. 8 of 15 patients (approx 50%)<br />

had a ≥ 2 improvement in MELD scores whilst 4 (approx 25%)<br />

had an increase in MELD score of ≥ 2. Improvement in MELD<br />

≥ 2 was only seen in patients with a pre-treatment MELD of <<br />

20. The improvements in CPS ≥ 2 were seen in 6/15 (40%).<br />

SVR12 data and post-transplant HCV RNA status for all patients<br />

will be presented. Conclusion: Sofosbuvir/Daclatasvir therapy<br />

in patients with significantly decompensated HCV liver disease<br />

results in on-treatment virological control in almost all patients.<br />

Despite this, EOT improvements in MELD scores were only seen<br />

in 50% of patients. Improvements were limited to patients with<br />

MELD scores < 20. SVRs and further assessment of liver disease<br />

severity are currently being analysed.<br />

Disclosures:<br />

Geoff McCaughan - Advisory Committees or Review Panels: Gilead<br />

Stuart K. Roberts - Board Membership: AbbVie, Gilead<br />

Simone I. Strasser - Advisory Committees or Review Panels: Janssen, AbbVie,<br />

Roche Products Australia, MSD, Bristol-Myers Squibb, Gilead, Norgine, Bayer<br />

Healthcare; Speaking and Teaching: Bayer Healthcare, Bristol-Myers Squibb,<br />

MSD, Roche Products Australia, Gilead, Janssen, Abbvie<br />

Jacob George - Advisory Committees or Review Panels: Roche, BMS, MSD,<br />

Gilead, Janssen, Abbvie; Grant/Research Support: MSD<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

Susan Mason - Advisory Committees or Review Panels: Janssen, MSD, BMS;<br />

Grant/Research Support: Janssen, MSD; Speaking and Teaching: ASHM<br />

Peter W. Angus - Advisory Committees or Review Panels: Gilead Sciences, BMS;<br />

Grant/Research Support: Gilead sciences<br />

The following authors have nothing to disclose: Paul Gow, Alan J. Wigg, Caroline<br />

Tallis, Gary P. Jeffrey, Joanne Mitchell<br />

1078<br />

(THE SOFGER TRIAL) Sofosbuvir-based treatment under<br />

real life conditions in Germany<br />

Peter Buggisch 1 , Christoph Sarrazin 2 , Stefan Mauss 3 , Holger<br />

Hinrichsen 4 , Karl-Georg Simon 5 , Dietrich Hueppe 6 , Johannes Vermehren<br />

2 , Barbara Seegers 4 , Joerg Petersen 1 ; 1 IFI, Liver Center<br />

Hamburg, Hamburg, Germany; 2 1 Medizinische Klinik, Johann<br />

Wolfgang Goethe University Hospital, Frankfurt, Germany; 3 Zentrum<br />

für HIV und Hepatogastroenterologie, Duesseldorf, Germany;<br />

4 Gastroenterolgische Praxis, Kiel, Germany; 5 Gastroenterologisch-hepatologisches<br />

Zentrum, Leverkusen, Germany; 6 Gastroenterologische<br />

Gemeinschaftspraxis, Herne, Germany<br />

Background and Aims: After the approval of direct-acting antivirals<br />

(DAA) in Germany 2014 starting with Sofosbuvir(SOF)<br />

Jan., Simeprevir(SIM) May, Daclatasvir(DCV) August and Ledipasvir(LDV)<br />

November, recommendations in Germany (BNG/<br />

DGVS) were rapidly changed accordingly. Rules for reimbursement<br />

changed over time,too, but there was no limitation to<br />

advanced fibrosis stages. Aim of this study was to evaluate<br />

the change of treatment modalities following availability of<br />

treatment options, implementation of recommendation/reimbursement<br />

rules as well as the safety and efficacy of the SOF<br />

based therapies under real life conditions. Methods: In this<br />

non-interventional, prospective, multi-center study conducted<br />

by the Association of German Gastroenterologists in private<br />

practice (BNG) (5 center) and one academic based center<br />

(Frankfurt), SOF-based therapies from Jan.2014 to February<br />

2015 are being evaluated in Germany. Demographic, clinical,<br />

virology data and adverse events are collected throughout<br />

treatment and post-treatment. This interim analysis shows data<br />

about pts.with SOF-based treatments. Results: 790 patients<br />

(471 male, 318 female) received SOF-based treatments. 553<br />

pts.with genotype 1(275 1a, 278 1b), 61 genotype 2, 128<br />

genotype 3, 48 genotype 4. 393 (49,8% had previous treatment<br />

(65 with DAAs), 417 of pat.had fibrosis F3-4. In 118 pat.<br />

no exact fibosis stage was evalueted, but 45% of those were<br />

clinically cirrhotic. 93 pat. received SOF/Riba (84% Genotype<br />

2(12 w) and 3(24 w) 160 received PEG/RIBA/SOF 12 w<br />

(all before Sep.2014) 118 SIM/SOF, 221 SOF+DCV ±Riba<br />

(75% F3/4, 198 SOF/LDV (58% F0-2). No IFN-based treatment<br />

started after 9/2014. SVR 4 for SIM/SOF (116 pat.) is<br />

87% SVR 12 (90 pat.) 86 %. SVR 4 for PEG/RIBA/SOF (157<br />

pat.) is 83%, SVR 12 (144 pat.)79%. SVR 4 for SOF/RIBA is<br />

78%(88 pat.) SVR 12 75% (82 pat). SVR 4 for SOF/DCV is<br />

85% (163 pat.) SVR 12 (97 pat.) 87%. SVR 4 for SOF/LDV<br />

is 96% (110 pat.). There was a clear trend to use SOF/LDV<br />

in lower fibrosis stages (8 or 12 weeks). Anemia (Riba-based<br />

therapy),headache, sleeping disorders and fatigue were the<br />

most reported AEs. No severe side effects occurred. 6 deaths<br />

were reported due to complications of cirrhosis. More SVR


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 739A<br />

12 data and detailed results for the different treatments will<br />

be available at the meeting. Conclusions: SOF-based triple<br />

therapies under real life conditions seems to have comparable<br />

results to clinical trials and other real life cohorts.There was a<br />

fast switch to IFN-free therapies in Germany following the recommendations.<br />

Advanced patients were treated first, but now<br />

almost 40% were treated without advanced fibrosis. Adverse<br />

events and treatment discontinuations were low.<br />

Disclosures:<br />

Peter Buggisch - Advisory Committees or Review Panels: Janssen, AbbVie, BMS;<br />

Speaking and Teaching: Roche, MSD, Gilead, Merz Pharma<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

Stefan Mauss - Advisory Committees or Review Panels: BMS, AbbVie, Janssen,<br />

ViiV, Gilead; Speaking and Teaching: BMS, AbbVie, Janssen, Gilead, MSD<br />

Holger Hinrichsen - Advisory Committees or Review Panels: BMS, Janssen, Gilead,<br />

Abbvie; Speaking and Teaching: MSD<br />

Karl-Georg Simon - Advisory Committees or Review Panels: AbbVie, BMS,<br />

JANSSEN; Speaking and Teaching: AbbVie, BMS, FALK, GILEAD, JANSSEN,<br />

NORGINE<br />

Dietrich Hueppe - Advisory Committees or Review Panels: Roche, MSD, Novatis,<br />

Gilead, BMS, Janssen<br />

Johannes Vermehren - Advisory Committees or Review Panels: Abbott; Speaking<br />

and Teaching: BMS, Covidien<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

The following authors have nothing to disclose: Barbara Seegers<br />

1079<br />

Quality-Adjusted Cost of Care for Treatment Naive (TN)<br />

Patients with Genotype 1 (GT1) Chronic Hepatitis C (CH-<br />

C): An Assessment of Innovation Cost of Drug Regimens<br />

versus the Value of Health Gains to the Society<br />

Zobair M. Younossi 1,2 , Haesuk Park 3 , Douglas Dieterich 4 , Sammy<br />

Saab 5 , Aijaz Ahmed 6 , Stuart C. Gordon 7 ; 1 Center For Liver<br />

Disease, Department of Medicine, Inova Fairfax Hospital, Falls<br />

Church, VA; 2 Betty and Guy Beatty Center for Integrated Research,<br />

Inova Health System, Falls Church, VA; 3 University of Florida,<br />

Gainesville, FL; 4 Mount Sinai Medical Center, New York City, NY;<br />

5 University of California Los Angeles, Los Angeles, CA; 6 Stanford<br />

University, Standford, CA; 7 Henry Ford Hospital, Detroit, MI<br />

Objectives: New HCV regimens have increased cure rates and<br />

treatment costs. Quality-adjusted cost of care (QACC) provides<br />

an approach for assessing cost growth of regimens in the<br />

context of value of health [quality-adjusted life-years (QALY)]<br />

improvements delivered by these regimens (Lakdawalla, Health<br />

Affairs 2015). Our aim was to assess QACC for regimens<br />

approved for GT1 CH-C in the US. Methods: A decision analytic<br />

Markov compared treatment of TN GT1 CHC patients<br />

with approved regimens including pegylated interferon and<br />

ribavirin (PR), 1st generation triple (boceprevir+PR and telaprevir+PR),<br />

2nd generation triple (sofosbuvir+PR and simeprevir+PR)<br />

and all-oral regimens (ledipasvir/sofosbuvir and<br />

ombitasvir+paritaprevir/ritonavir+dasabuvir±ribavirin). Sustained<br />

virologic response (SVR), transition probabilities, utility,<br />

and mortality were based on literature review and consensus<br />

by hepatologists. Drug costs were from Redbook using wholesale<br />

acquisition costs. QACC was defined as the increase in<br />

treatment price minus the increase in the patient’s QALYs multiplied<br />

by the value of a QALY ($50,000). Since this threshold<br />

may be outdated, we also estimated the value of a QALY to<br />

be between $100,000 and $300,000. Results: Average drug<br />

costs increased by $41,311, $54,249, and $52,352 for 1st,<br />

2nd generation triples and all-oral therapies compared to PR<br />

($35,331). Health improvements were 0.121, 0.692, and<br />

1.366 QALYs, valued at $6,057, $34,622, and $68,318<br />

for 1st, 2nd generation triples and all-oral therapies compared<br />

to PR. Compared to PR, 1st and 2nd generation triples could<br />

cost the society $35,254 and $19,627. On the other hand,<br />

all-oral therapy could save the society $15,966. By increasing<br />

the value threshold for a QALY up to $300,000, all-oral regimens<br />

save the society up to $357,555. Conclusions: Despite<br />

the higher cost of all-oral regimens, this QACC analysis shows<br />

substantial value to the society.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Haesuk Park - Consulting: Gilead Science<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Sammy Saab - Advisory Committees or Review Panels: BMS, Gilead, Merck,<br />

Janssen; Grant/Research Support: Gilead; Speaking and Teaching: BMS, Gilead,<br />

Merck, Janssen, Salix, Onyx, Bayer, Janssen; Stock Shareholder: Achillion,<br />

Johnson and Johnson, BMS, Gilead<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

1080<br />

Asian Patients with Chronic Genotype 2 HCV Infection<br />

Achieve 98% Sustained Virologic Response Following<br />

12 Week administration of Sofosbuvir in Combination<br />

with Ribavirin: Integrated analysis of Phase 3 Multicenter<br />

Studies<br />

Masao Omata 1 , Young-Suk Lim 2 , Wan-Long Chuang 3 , Jia-Horng<br />

Kao 4 , Bing Gao 5 , Shampa De-Oertel 5 , Jenny C. Yang 5 , Hongmei<br />

Mo 5 , Diana M. Brainard 5 , Steven J. Knox 5 , John G. McHutchison<br />

5 , Masashi Mizokami 6 , Kwang-Hyub Han 7 ; 1 Yamanashi Prefectural<br />

Hospital Organization, Yamanashi, Japan; 2 Asan Medical<br />

Center, University of Ulsan College of Medicine, Seoul, Korea<br />

(the Republic of); 3 Kaohsiung Medical University, Kaohsiung, Taiwan;<br />

4 National Taiwan University College of Medicine, Taipei,<br />

Taiwan; 5 Gilead Sciences, Inc, Foster City, CA; 6 National Center<br />

for Global Health and Medicine, Tokyo, Japan; 7 Yonsei University<br />

College of Medicine, Seoul, Korea (the Republic of)<br />

Introduction: Chronic hepatitis C (CHC) infection presents a<br />

significant burden on public health in Asia. Among patients


740A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

with CHC, 20-30% in Japan and 50% in Korea and Taiwan<br />

are infected with hepatitis C virus (HCV) genotype (GT) 2.<br />

Sofosbuvir (SOF), an NS5B polymerase inhibitor, in combination<br />

with ribavirin (RBV) is the first all-oral regimen for<br />

treatment of HCV GT2 infection. The aim of this integrated<br />

analysis is to characterize the efficacy and safety of SOF+RBV<br />

in a large cohort of Asian patients with HCV GT2 infection.<br />

Methods: This integrated analysis combines results from two<br />

Phase 3 trials, GS-US-334-0118 (Japan) and GS-US-334-<br />

0115 (Korea and Taiwan). In both <strong>studies</strong>, treatment-naïve<br />

and treatment-experienced adults with chronic GT2 HCV infection<br />

received SOF (400mg) combined with RBV (weight-based<br />

dosing) for 12 weeks. Eligibility criteria: no upper age limit<br />

restriction, inclusion of subjects with compensated cirrhosis,<br />

platelet count ≥50,000/mL and no restriction on neutrophil<br />

count. The primary efficacy endpoint was Sustained Virologic<br />

Response measured 12 weeks after the last dose of study drug<br />

(SVR12). Results: Overall, 369 patients were enrolled (n=129<br />

Korea, n=87 Taiwan, and n=153 Japan), of which 64%<br />

(238/369) were treatment naïve and 36% (131/369) were<br />

treatment experienced. Mean age (range) was 55 (22-82)<br />

years, 44% male (164/369), 82.1% (303/369) IL28B-CC,<br />

and 12% (43/369) had cirrhosis. The overall SVR12 rate was<br />

98% (360/369). Similar high rates of SVR12 were seen in<br />

treatment-experienced (98%, 128/131), patients ≥65 years<br />

old (96%, 73/76), and those with cirrhosis (98%, 42/43).<br />

Among patients who did not achieve SVR12 (n=9), 1 was a<br />

partial responder, 6 relapsed, and 2 were lost to follow up.<br />

Adverse events were generally mild to moderate. The most<br />

common adverse events were nasopharyngitis (17%, 63/369),<br />

headache (12%, 45/369), and pruritis (11%, 41/369). Laboratory<br />

abnormalities were infrequent and consistent with the<br />

safety profile of RBV. No AEs led to treatment discontinuation.<br />

Conclusions: Treatment-naïve and treatment-experienced Asian<br />

patients in Japan, Korea, and Taiwan with chronic GT2 HCV<br />

infection, including those with compensated cirrhosis, achieved<br />

high rates of SVR12 with 12 weeks of an IFN-free, all-oral<br />

regimen of SOF+RBV. The regimen was safe and well-tolerated<br />

with no treatment discontinuations due to AE and the overall<br />

AE profile was consistent with that observed with RBV. The<br />

data suggest that SOF+RBV may offer an improved, IFN-free<br />

treatment for Asian patients with chronic GT2 HCV infection.<br />

Disclosures:<br />

Masao Omata - Advisory Committees or Review Panels: Boehringer Ingelheim;<br />

Speaking and Teaching: Otsuka Pharmaceutical, Bayer<br />

Young-Suk Lim - Advisory Committees or Review Panels: Bayer Healthcare, Gilead<br />

Sciences; Grant/Research Support: Bayer Healthcare, BMS, Gilead Sciences,<br />

Novartis<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Bing Gao - Employment: Gilead; Stock Shareholder: Gilead<br />

Shampa De-Oertel - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Jenny C. Yang - Employment: Gilead Sciences, Inc<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Steven J. Knox - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

The following authors have nothing to disclose: Jia-Horng Kao, Masashi<br />

Mizokami, Kwang-Hyub Han<br />

1081<br />

Sofosbuvir and ledipasvir for 8 weeks in patients with<br />

hepatitis C virus (HCV) mono-infection and human<br />

immunodeficiency virus (HIV)-HCV co-infection with<br />

genotype 1 and 4 in clinical practice – Results from the<br />

GErman hepatitis C COhort (GECCO)<br />

Stefan Christensen 1 , Stefan Mauss 2 , Dietrich Hueppe 3 , Knud<br />

Schewe 4 , Thomas Lutz 5 , Jürgen K. Rockstroh 6 , Marcel Schuetze 7 ,<br />

Guenther Schmutz 2 , Karl-Georg Simon 8 , Heiner W. Busch 1 , Patrick<br />

Ingiliz 7 , Axel Baumgarten 7 ; 1 Center for interdisciplinary Medicine<br />

(CIM), Muenster, Germany; 2 Center for HIV and Hepatogastroenterology,<br />

Duesseldorf, Germany; 3 Practice for Gastroenterology,<br />

Herne, Germany; 4 Infektionsmedizinisches Centrum Hamburg,<br />

Hamburg, Germany; 5 Infektiologikum, Frankfurt, Germany; 6 Internal<br />

Medicine I, University of Bonn, Bonn, Germany; 7 Medizinisches<br />

Infektiologie Zentrum Berlin, Berlin, Germany; 8 Practice<br />

for Gastroenterology Leverkusen, Leverkusen, Germany<br />

Introduction The first direct acting antivirals (DAA) were<br />

approved in Europe in 2014 based on limited study data. In<br />

particular, sofosbuvir (SOF) and ledipasvir (LDV) for 8 weeks in<br />

HCV-mono and HIV co-infected patients had not been systematically<br />

studied. Here, we present real-life data on efficacy and<br />

safety of SOF/ LDV for 8 weeks from Germany. Methods In this<br />

multicenter cohort, all patients who were started on sofosbuvir-based<br />

treatments were documented. For the current analysis,<br />

only patients with HCV genotype (GT) 1 and 4 on SOF/LDV for<br />

8 weeks were included. All patients within the GECCO cohort<br />

are part of the prospective German hepatitis C registry. Results<br />

Until May 2015, 836 patients on SOF-containing regimens<br />

have been enrolled. Of those, 629 (75%) are HCV-monoinfected<br />

and 207 (25%) HIV-HCV co-infected. 95/836 (92 GT 1<br />

and 3 GT 4) patients were treated with SOF/LDV for 8 weeks.<br />

19/95 patients (19%) had been unsuccessfully treated before,<br />

mainly with dual therapy consisting of pegylated interferon and<br />

ribavirin. 16/95 (17%) were HIV coinfected, 14 of those with a<br />

HIV-RNA < 15 cop/ml. 40/95 (42%) were male, median age<br />

(IQR) was 53 years (23-81). Median HCV-RNA was 809.900<br />

IU/ml (IQR 5.000-14.100.000). 6/95 (6%) had a HCV-RNA<br />

at baseline of > 6 mio IU/ml. 59/95 (62%) had liver stiffness<br />

measurement by transient elastography (Fibroscan®), with a<br />

median of 6,1kPa. Liver cirrhosis defined by a value above<br />

12.5kPa was present in 2 cases. 44/95 have reached post<br />

treatment week 4. All (44/44) had a negative HCV-RNA at<br />

post treatment week 4 (SVR 4). No premature discontinuations<br />

or lost to follow up are documented. Conclusion: In this this<br />

real life patient population with mainly low HCV viral load and<br />

without significant fibrosis we found high response rates in GT<br />

1 and GT 4 patients treated with SOF/LDV for 8 weeks. Even<br />

in pretreated patients, patients with advanced fibrosis and in<br />

patients with a HCV-RNA > 6 mio IU/ml, all not candidates<br />

for a short course SOF/LDV according to the German HCV<br />

treatment guidelines, 8 weeks of SOF/LDV seems to be safe<br />

and effective.<br />

Disclosures:<br />

Stefan Christensen - Advisory Committees or Review Panels: BMS, Abbvie, Janssen,<br />

ViiV, Gilead, MSD; Speaking and Teaching: Gilead, MSD, Abbvie, BMS,<br />

ViiV, Reckitt Benckiser, Janssen<br />

Stefan Mauss - Advisory Committees or Review Panels: BMS, AbbVie, Janssen,<br />

ViiV, Gilead; Speaking and Teaching: BMS, AbbVie, Janssen, Gilead, MSD<br />

Dietrich Hueppe - Advisory Committees or Review Panels: Roche, MSD, Novatis,<br />

Gilead, BMS, Janssen<br />

Knud Schewe - Advisory Committees or Review Panels: abbvie, gilead, msd,<br />

bms, janssen cilag, viiv, hexal<br />

Thomas Lutz - Advisory Committees or Review Panels: Gilead, MSD, AbbVie,<br />

BMS, Janssen Cilag; Grant/Research Support: Gilead, GlaxoSmithKline, Roche,<br />

MSD, AbbVie, Janssen Cilag; Speaking and Teaching: ViiV, BMS


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 741A<br />

Jürgen K. Rockstroh - Advisory Committees or Review Panels: Abbvie, BI, BMS,<br />

Merck, Roche, Tibotec, Abbvie, Bionor, Tobira, ViiV, Gilead, Janssen; Consulting:<br />

Novartis; Grant/Research Support: Merck; Speaking and Teaching: Abbott,<br />

BI, BMS, Merck, Roche, Tibotec, Gilead, Janssen, ViiV<br />

Karl-Georg Simon - Advisory Committees or Review Panels: AbbVie, BMS,<br />

JANSSEN; Speaking and Teaching: AbbVie, BMS, FALK, GILEAD, JANSSEN,<br />

NORGINE<br />

Heiner W. Busch - Speaking and Teaching: Tibotec, MSD, Janssen, BMS, Gilead<br />

The following authors have nothing to disclose: Marcel Schuetze, Guenther<br />

Schmutz, Patrick Ingiliz, Axel Baumgarten<br />

1082<br />

Hepatitis C virus NS5A L31V plus Y93H variants emerging<br />

after treatment failure with daclatasvir/asunaprevir<br />

are resistant to ledipasvir and sofosbuvir-like nucleotide<br />

polymerase inhibitor GS-558093 in human hepatocyte<br />

chimeric mice<br />

Yugo Kai, Hayato Hikita, Tomohide Tatsumi, Kazuhiro Murai, Yuto<br />

Shiode, Yasutoshi Nozaki, Yuki Makino, Tasuku Nakabori, Yoshinobu<br />

Saito, Naoki Morishita, Satoshi Tanaka, Ryotaro Sakamori,<br />

Naoki Hiramatsu, Tetsuo Takehara; Department of Gastroenterology<br />

and Hepatology, Osaka University Graduate School of Medicine,<br />

Suita, Japan<br />

Background: Resistance-associated variants (RAVs) emerge<br />

in hepatitis C virus (HCV) infection upon treatment failure of<br />

direct-acting antivirals (DAAs). Treatment failure of the asunaprevir<br />

(ASV) and daclatasvir (DCV) regimen is associated<br />

with emergence of RAVs at both L31 and Y93 in the HCV<br />

NS5A region. In the present study, we used human hepatocyte<br />

chimeric mice to examine the therapeutic effect of ledipasvir<br />

(LDV) and GS-558093 (NS5B NI), a nucleotide NS5B polymerase<br />

inhibitor like sofosbuvir, as a re-treatment option for<br />

L31 and Y93 treatment emergent RAVs. Methods: Thymidine<br />

kinase transgenic NOG (TK-NOG) mice with humanized liver<br />

were inoculated with serum from treatment-naïve and treatment-experienced<br />

patients with chronic hepatitis C under the<br />

approval of the institutional ethics committee. After developing<br />

persistent HCV infection, mice were administrated DAAs orally.<br />

HCV RNA levels were measured by TaqMan PCR, and population<br />

and deep sequencing were performed. Results: Human<br />

hepatocyte chimeric mice were inoculated with serum from<br />

either a patient who had failed to achieve sustained virologic<br />

response (SVR) by the ASV/DCV therapy or from a patient who<br />

had not experienced DAA therapy. After inoculation, the mice<br />

developed persistent HCV infection with HCV RNA levels up to<br />

5-7 LC/ml in serum. Population and deep sequencing revealed<br />

that the formers carried NS3/4A D168V, NS5A L31V plus<br />

Y93H triple mutant virus while the latter carried wild-type virus<br />

at these positions. Mice in each group were divided into 2<br />

treatment groups: one received 4 weeks of ASV/DCV administration<br />

and the other received 4 weeks of LDV/ NS5B NI.<br />

Serum HCV RNA levels in mice with wild-type HCV rapidly<br />

declined and sustained undetectable levels after ASV/DCV<br />

or LDV/ NS5B NI. In contrast, serum HCV RNA levels in mice<br />

with mutant virus reduced 1 to 2 LC/mL or 2 to 3 LC/mL from<br />

baseline by ASV/DCV or LDV/ NS5B NI therapy, respectively,<br />

but did not achieve undetectable levels at the end-of-treatment.<br />

Although mice infected with mutant virus failed to eradicate<br />

HCV by both therapies, no additional substitutions including<br />

NS5B S282T were detected. These mice were re-treated by<br />

combined administration of NS5B NI and telaprevir (TVR), of<br />

which resistance profile differs from ASV. Serum HCV RNA<br />

levels rapidly reached to undetectable levels. Discussion/Conclusion:<br />

The triple mutant virus emerging upon the treatment<br />

failure of ASV and DCV is relatively resistant to LDV/ NS5B<br />

NI. Other DAA regimens that do not share cross-resistance<br />

at NS3/4A D168, NS5A L31 and Y93 positions may be a<br />

choice of re-treatment for such case.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yugo Kai, Tomohide Tatsumi,<br />

Kazuhiro Murai, Yuto Shiode, Yasutoshi Nozaki, Yuki Makino, Tasuku Nakabori,<br />

Yoshinobu Saito, Naoki Morishita, Satoshi Tanaka, Ryotaro Sakamori, Naoki<br />

Hiramatsu<br />

1083<br />

Sofosbuvir and Ribavirin for Six Weeks Is Not Effective<br />

Among People with Acute and Recently Acquired HCV<br />

Infection: The DARE-C II Study<br />

Marianne Martinello 1 , Edward J. Gane 2 , Margaret Hellard 3 , Joe<br />

Sasadeusz 4 , David Shaw 5 , Kathy Petoumenos 1 , Tanya L. Applegate<br />

1 , Jason Grebely 1 , Laurence Maire 1 , Philippa Marks 1 , David<br />

Cooper 1 , Gregory Dore 1 , Gail Matthews 1 ; 1 Kirby Institute, UNSW<br />

Australia, Sydney, NSW, Australia; 2 Auckland Hospital, Auckland,<br />

New Zealand; 3 Burnet Institute, Melbourne, VIC, Australia;<br />

4 Royal Melbourne Hospital, Melbourne, VIC, Australia; 5 Royal<br />

Adelaide Hospital, Adelaide, SA, Australia<br />

Background: Sofosbuvir (SOF) and ribavirin (RBV) are effective<br />

and well tolerated when administered for 12–24 weeks in genotype<br />

(GT) 1-4 chronic hepatitis C virus (HCV) infection. The<br />

aim of this study was to assess the efficacy of SOF and RBV for<br />

six weeks in individuals with acute or recently acquired HCV<br />

infection. Methods: In this multicentre study conducted in Australia<br />

and New Zealand, adults with recent HCV (duration of<br />

infection


742A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ever, efficacy was suboptimal. Further research is needed to<br />

determine whether more potent DAA regimens with or without<br />

RBV will allow treatment duration to be shortened in acute or<br />

recently acquired HCV infection.<br />

Disclosures:<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Joe Sasadeusz - Grant/Research Support: Gilead Sciences, BMS, Roche, Janssen;<br />

Speaking and Teaching: Gilead Sciences, Roche, BMS<br />

Kathy Petoumenos - Grant/Research Support: Gilead Sciences<br />

Jason Grebely - Advisory Committees or Review Panels: Merck, Gilead; Grant/<br />

Research Support: Merck, Gilead, Abbvie, BMS<br />

Gregory Dore - Board Membership: Gilead, Merck, Abbvie, Bristol-Myers<br />

Squibb; Grant/Research Support: Gilead, Merck, Abbvie, Bristol-Myers Squibb;<br />

Speaking and Teaching: Gilead, Merck, Abbvie, Bristol-Myers Squibb<br />

Gail Matthews - Advisory Committees or Review Panels: gilead; Consulting: Viiv;<br />

Grant/Research Support: Gilead Sciences, janssen; Speaking and Teaching:<br />

BMS, MSD<br />

The following authors have nothing to disclose: Marianne Martinello, Margaret<br />

Hellard, David Shaw, Tanya L. Applegate, Laurence Maire, Philippa Marks,<br />

David Cooper<br />

patients, including 6 with GT1b infection, received 3D+RBV<br />

for 24 weeks, and 13 GT1b patients received treatment without<br />

RBV for 24 weeks. SVR4 was achieved in 39/40 (98%)<br />

patients. The most common adverse events (AEs) were fatigue,<br />

nausea, headache, rash, anemia, and asthenia. Two serious<br />

AEs were reported, neither of which was attributed to study<br />

drugs. One patient with a history of prior LT rejection discontinued<br />

treatment on day 86 due to low grade rejection revealed<br />

by liver biopsy, but achieved SVR12. Hemoglobin declines<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 743A<br />

1085<br />

Utilizing claims data to understand the relationships<br />

between diagnosing, prescribing, and dispensing patterns<br />

for HCV patients in the pre and post sofosbivir<br />

eras<br />

Michel F. Denarie; Real world evidence Solutions, IMS Health,<br />

Yardley, PA<br />

UTILIZING CLAIMS DATA TO UNDERSTAND THE RELATION-<br />

SHIPS BETWEEN DIAGNOSING, PRESCRIBING, AND DIS-<br />

PENSING PATTERNS FOR HCV PATIENTS IN THE PRE AND<br />

POST SOFOSBUVIR ERAS BACKGROUND AND AIMS: The<br />

treatment of Hepatitis C (HCV) patients has undergone profound<br />

change since the introduction of sofosbuvir in late 2013.<br />

Many patients were warehoused until the launch of sofosbuvir<br />

because it was the first directly acting antiviral (DAA) to offer an<br />

interferon-free, short duration, highly tolerable and efficacious<br />

regimen. The study also determined the amount and duration<br />

of patient warehousing and the variations between ethnicities.<br />

The intent was to utilize claims data to demonstrate how HCV<br />

treatment dynamics have changed in the pre and post sofosbuvir<br />

era. DESIGN/METHODS: This 2 year retrospective study<br />

utilized US IMS claims data. Patients were selected for HCV<br />

diagnosis for a period of six months, then followed for a period<br />

of 12 months to determine the date of prescription, and then<br />

followed for an additional 6 months to ascertain whether they<br />

were actually dispensed the prescribed product(s). We limited<br />

the analysis to patients prescribed telaprevir, boceprevir, sofosbuvir<br />

and simeprevir; breaking into two patient cohorts: pre-sofosbuvir<br />

(Period 1/Jan 2012 - Nov 2013) and post-sofosbuvir<br />

(Period 2/Dec 2013 - Aug 2014) launch. Using these cohorts,<br />

we were able to determine an index diagnosis date, an index<br />

written date, and an index dispensed date. RESULTS: 11,256<br />

newly diagnosed HCV patients (40% females and 60% males)<br />

were included in the study. Results demonstrate the amount and<br />

length of patient warehousing increased dramatically in the 12<br />

months preceding the introduction of sofosbuvir and simeprevir<br />

(see Chart 1), most notably for patients prescribed sofosbuvir<br />

and simeprevir. During Period 1, 93% of simeprevir and 81%<br />

of sofosbuvir patients were dispensed their medication more<br />

than 12 months after the initial HCV diagnosis was recorded,<br />

while only 45% of simeprevir and 36% of sofosbuvir patients<br />

were dispensed their medication more than 12 months after<br />

initial diagnosis in Period 2. The time between the WRx (written<br />

scripts) to the DRx (dispensed script) did not materially change.<br />

Chart 1 CONCLUSIONS: This study displayed the drastic differences<br />

in patient warehousing between Periods 1 and 2,<br />

most notably for sofosbuvir and simeprevir treatments. Many<br />

newly diagnosed HCV patients waited for the new DAAs to be<br />

introduced to receive a prescription. Once prescribed, despite<br />

the high cost of treatment, the time between the WRx and the<br />

DRx did not materially increase, and remained mostly within<br />

one month indicating the benefits outweighed the cost.<br />

Disclosures:<br />

The following authors have nothing to disclose: Michel F. Denarie<br />

1086<br />

Long-Term Efficacy of Ombitasvir/Paritaprevir/r and<br />

Dasabuvir With or Without Ribavirin in HCV Genotype<br />

1-Infected Patients With or Without Cirrhosis<br />

Stefan Zeuzem 2 , Ira M. Jacobson 3 , Jordan J. Feld 4 , Heiner Wedemeyer<br />

5 , Xavier Forns 6 , Pietro Andreone 7 , Massimo Colombo 8 ,<br />

David Bernstein 9 , Fred Poordad 10 , Christophe Hezode 11 , Thomas<br />

Podsadecki 1 , Wangang Xie 1 , Tami Pilot-Matias 1 , Regis A.<br />

Vilchez 1 , John M. Vierling 12 ; 1 AbbVie, North Chicago, IL; 2 J.W.<br />

Goethe University, Frankfurt, Germany; 3 Weill Cornell Medical<br />

College, New York, NY; 4 Toronto Centre for Liver Disease, University<br />

of Toronto, Toronto, ON, Canada; 5 Medizinische Hochschule<br />

Hannover, Hannover, Germany; 6 Liver Unit, Hospital Clinic,<br />

IDIBAPS and CIBEREHD, Barcelona, Spain; 7 Dipartimento di Scienze<br />

Mediche e Chirurgiche, University of Bologna, Bologna, Italy;<br />

8 Fondazione IRCCS Cà Granda Ospedale Maggiore Policlinico,<br />

Università degli Studi di Milano, Milan, Italy; 9 North Shore University<br />

Hospital, Manhasset, NY; 10 The Texas Liver Institute/University<br />

of Texas Health Science Center, San Antonio, TX; 11 Henri Mondor<br />

University Hospital, AP-HP, Universite Paris-Est, Cr’eteil, France;<br />

12 Baylor College of Medicine, St. Luke’s Advanced Liver Therapies,<br />

Houston, TX<br />

Introduction: Treatment with the 3 direct-acting antiviral (3D)<br />

regimen of ombitasvir (OBV), paritaprevir (PTV, boosted with<br />

ritonavir [r]; identified by AbbVie and Enanta) and dasabuvir<br />

(DSV), with or without ribavirin (RBV), achieved sustained virologic<br />

response (SVR) rates between 95% and 100% across a<br />

broad range of hepatitis C virus (HCV) genotype (GT) 1-infected<br />

patients. In this analysis, we examined the efficacy through<br />

post-treatment week 48 of the 3D regimen in HCV GT1-infected<br />

patients with or without cirrhosis. Methods: Treatment-naïve and<br />

-experienced HCV GT1a- or GT1b-infected patients with and<br />

without compensated cirrhosis received the recommended 3D<br />

regimen of co-formulated OBV/PTV/r (25mg/150mg/100mg<br />

once daily) and DSV (250mg twice daily), ± weightbased<br />

RBV, for 12 or 24 weeks in six phase 3 <strong>studies</strong>. GT1b-infected<br />

patients without cirrhosis did not receive RBV as part of their<br />

treatment regimen. SVR at post-treatment week 12 (SVR12;<br />

HCV RNA


744A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

Xavier Forns - Consulting: Jansen, Abbvie; Grant/Research Support: Jansen,<br />

Gilead<br />

Pietro Andreone - Advisory Committees or Review Panels: Janssen-Cilag, Gilead,<br />

MSD/Schering-Plough, Abbvie; Speaking and Teaching: Gilead, BMS<br />

Massimo Colombo - Advisory Committees or Review Panels: BRISTOL-MEY-<br />

ERS-SQUIBB, SCHERING-PLOUGH, ROCHE, GILEAD, BRISTOL-MEYERS-SQUIBB,<br />

SCHERING-PLOUGH, ROCHE, GILEAD, Janssen Cilag, Achillion; Grant/<br />

Research Support: BRISTOL-MEYERS-SQUIBB, ROCHE, GILEAD, BRISTOL-MEY-<br />

ERS-SQUIBB, ROCHE, GILEAD; Speaking and Teaching: Glaxo Smith-Kline,<br />

BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH, ROCHE, NOVARTIS, GILEAD,<br />

VERTEX, Glaxo Smith-Kline, BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH,<br />

ROCHE, NOVARTIS, GILEAD, VERTEX, Sanofi<br />

David Bernstein - Advisory Committees or Review Panels: Gilead; Consulting:<br />

Abbvie, BMS, Merck, Janssen; Grant/Research Support: Gilead, Abbvie, BMS,<br />

Merck, Janssen, Genentech; Speaking and Teaching: Abbvie, BMS, Merck,<br />

Gilead<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Christophe Hezode - Speaking and Teaching: Roche, BMS, MSD, Janssen, abbvie,<br />

Gilead<br />

Thomas Podsadecki - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Wangang Xie - Employment: AbbVie<br />

Tami Pilot-Matias - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Regis A. Vilchez - Employment: AbbVie Inc.<br />

John M. Vierling - Advisory Committees or Review Panels: Abbvie, Bristol-Meyers-Squibb,<br />

Gilead, Hyperion, Intercept, Janssen, Novartis, Merck, Sundise,<br />

HepQuant, Salix, Immuron, Exalenz, Chronic Liver Disease Foundation; Board<br />

Membership: Clinical Research Centers of America, LLC; Grant/Research Support:<br />

Abbvie, Bristol-Meyers-Squibb, Eisai, Gilead, Hyperion, Intercept, Janssen,<br />

Novartis, Merck, Sundise, Ocera, Mochida, Immuron, Exalenz, Conatus; Speaking<br />

and Teaching: GALA, Chronic Liver Disease Foundation, ViralEd, Chronic<br />

Liver Disease Foundation, Clinical Care Options<br />

in lifetime risks of developing advanced liver disease (i.e. CC,<br />

DCC or HCC) and an 80% reduction in risks of LrD, regardless<br />

of treatment history or cirrhosis status. The lifetime risks of<br />

liver morbidity and mortality are also substantially lower in the<br />

overall coinfected patients treated with 3D±R than those treated<br />

with SOF+PR or SOF+R. In the naïve non-cirrhotic coinfected<br />

patients, treatment with 3D±R offers comparable reductions<br />

in lifetime risks of liver morbidity and mortality compared to<br />

SOF+LDV. Conclusion Compared with former or other current<br />

standards of care for treating GT1 HCV and HIV coinfected<br />

patients in the US, 3D±R offers favorable or comparable reductions<br />

in lifetime risks of liver morbidity and mortality.<br />

Lifetime risks of liver morbidity and mortality in GT1 HCV and HIV<br />

coinfected patients<br />

* Clinical trial data not available<br />

Disclosures:<br />

Sammy Saab - Advisory Committees or Review Panels: BMS, Gilead, Merck,<br />

Janssen; Grant/Research Support: Gilead; Speaking and Teaching: BMS, Gilead,<br />

Merck, Janssen, Salix, Onyx, Bayer, Janssen; Stock Shareholder: Achillion,<br />

Johnson and Johnson, BMS, Gilead<br />

Suchin Virabhak - Consulting: AbbVie<br />

Scott J. Johnson - Consulting: AbbVie; Employment: Medicus Economics<br />

Hélène Parisé - Consulting: MedicusEconomics LLC, AbbVie ; Employment: Statlog<br />

Consulting Inc<br />

Yuri Sanchez - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie<br />

The following authors have nothing to disclose: Alice Wang<br />

1087<br />

Lifetime risks of liver morbidity and mortality in patients<br />

with chronic genotype (GT) 1 hepatitis C virus (HCV)<br />

and HIV coinfection treated with 3D±R (ombitasvir/<br />

paritaprevir/ ritonavir, dasabuvir ± ribavirin) vs other<br />

standards of care in the US<br />

Sammy Saab 2 , Suchin Virabhak 3 , Scott J. Johnson 3 , Hélène<br />

Parisé 3 , Yuri Sanchez 1 , Alice Wang 1 , Timothy R. Juday 1 ; 1 HEOR,<br />

AbbVie, Mettawa, ID; 2 Pfleger Liver Institute, UCLA, Los Angeles,<br />

CA; 3 Medicus Economics Inc., Boston, MA<br />

Objective This study evaluated the lifetime risks of liver morbidity<br />

and mortality in patients with GT1 HCV and HIV coinfection<br />

treated with 3D±R for 12 or 24 weeks compared to other standards<br />

of care in the United States (US): sofosbuvir plus peg-interferon<br />

and ribavirin for 12 weeks (SOF+PR) and SOF+R for<br />

24 weeks in the overall coinfected population, and sofosbuvir<br />

plus ledipasvir for 12 weeks (SOF+LDV) in the treatment-naïve<br />

non-cirrhotic population. Methods A Markov model based on<br />

the natural history of liver disease progression estimated the<br />

risks of liver-related morbidity and mortality over a lifetime<br />

horizon for a cohort of patients with GT1 HCV and HIV coinfection.<br />

Baseline population characteristics and efficacy data<br />

were obtained from published clinical trials. Lifetime risks of<br />

compensated cirrhosis (CC), decompensated cirrhosis (DCC),<br />

hepatocellular carcinoma (HCC) and liver-related death (LrD)<br />

were analyzed. Treatment strategies included ‘no treatment’<br />

(NT), as well as regimens that have been studied in Phase II or<br />

III clinical trials of coinfected patients, and either recommended<br />

by the latest guidelines or approved by the Food and Drug<br />

Administration. Results Compared to untreated GT1 HCV and<br />

HIV coinfected patients, 3D±R offers at least a 59% reduction<br />

1088<br />

Effect of Chronic Kidney Disease on the Pharmacokinetics<br />

of Ombitasvir, Paritaprevir, Ritonavir and Dasabuvir<br />

in Subjects with HCV Genotype 1 Infection<br />

Diana L. Shuster, Rajeev Menon, Daniel E. Cohen, Amit Khatri;<br />

AbbVie, North Chicago, IL<br />

Background and Objective: The all-oral interferon-free, 3<br />

direct acting antiviral (3-DAA) regimen of ombitasvir (OBV)<br />

(NS5A inhibitor) + paritaprevir (NS3/4A protease inhibitor<br />

identified by AbbVie and Enanta), coadministered with ritonavir,<br />

a pharmacokinetic enhancer (PTV/r), + dasabuvir (DSV)<br />

(NS5B non-nucleoside polymerase inhibitor) ± ribavirin (RBV)<br />

is being evaluated in HCV genotype (GT) 1-infected subjects<br />

with chronic kidney disease (CKD). No meaningful alterations<br />

in exposures were seen when the 3 DAAs were administered<br />

to HCV-uninfected subjects with renal impairment. The present<br />

analysis examines the effect of CKD Stage 4 and Stage<br />

5 on the pharmacokinetics of OBV, PTV/r and DSV in HCV<br />

GT1-infected subjects. Methods: Pharmacokinetic data from<br />

a Phase 2 study (N=38) in subjects with normal or mild renal<br />

impairment (subjects “without kidney disease”) were combined<br />

with preliminary data from a Phase 3b study in subjects with<br />

CKD Stage 4 (N=5) or Stage 5 on hemodialysis (N=14). In<br />

both <strong>studies</strong> subjects received OBV/PTV/r 25/150/100 mg<br />

QD and DSV 250 mg BID ± RBV for 12 weeks. Pharmacokinetic<br />

parameters and steady-state exposures of the 3-DAA<br />

regimen were estimated using population pharmacokinetic<br />

models. Results: CKD was not a significant covariate in the<br />

population pharmacokinetic analyses, and the safety profile<br />

of the 3 DAAs was similar in subjects with or without CKD.<br />

PTV and DSV exposures were comparable (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 745A<br />

between subjects without kidney disease and CKD Stage 4.<br />

OBV and ritonavir exposures were ~80% and 200% higher,<br />

respectively, in CKD Stage 4 subjects in the limited number of<br />

subjects in this analyses. OBV and PTV exposures were comparable<br />

(7 log 10<br />

IU/mL. Reasons for ineligibility for ACTG<br />

A5327 were treatment initiated >24 weeks after HCV diagnosis<br />

(n=6); peak ALT


746A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

pre- or post-ACTG A5327 enrolment period (n=3) (1 man had<br />

2 reasons). All 12 men who were treated completed 12 weeks<br />

of SOF+RBV and 12 weeks of follow-up monitoring. Eleven<br />

(92%) achieved SVR 12. The one man with treatment failure,<br />

viral relapse between weeks 4 and 9 post-treatment, had a<br />

very low peak ALT (103 U/mL) and baseline VL higher than<br />

the median (6.12 log 10<br />

IU/mL). Irritability and insomnia were<br />

commonly reported but did not limit treatment. Conclusions<br />

SOF+RBV for 12 weeks was highly effective in the treatment of<br />

acute HCV in HIV-infected men in this “real-world” setting using<br />

broad enrollment criteria. Larger <strong>studies</strong> would be needed to<br />

confirm these findings, but with the subsequent availability of<br />

new oral agents, efforts should probably instead be directed<br />

toward developing shorter and RBV-free regimens.<br />

Disclosures:<br />

Daniel S. Fierer - Stock Shareholder: Gilead<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

The following authors have nothing to disclose: Zachary Barbati, Andrew L.<br />

Foster, Tristan Morey, Samuel Turner<br />

1091<br />

Sofosbuvir in combination with simeprevir +/- ribavirin<br />

in genotype 4 hepatitis C patients with advanced fibrosis<br />

or cirrhosis: real-life experience from Belgium<br />

Christophe Moreno 1 , Luc Lasser 2 , Jean Delwaide 3 , Peter Starkel 4 ,<br />

Wim Laleman 5 , Philippe Langlet 6 , Hendrik Reynaert 7 , Stefan<br />

Bourgeois 8 , Sergio Negrin Dastis 9 , Thierry Gustot 1 , Sven M.<br />

Francque 10 , Anja M. Geerts 11 , Christophe Van Steenkiste 12 ,<br />

Chantal de Galocsy 13 , Collins Assene 13 , Hans Orlent 14 , Marcel<br />

Nkuize 15 , Delphine Degré 1 ; 1 Department of Gastroenterology,<br />

Hepatopancreatology and Digestive oncology, CUB Hôpital<br />

Erasme, Brussels, Belgium; 2 Department of Hepatogastroenterology,<br />

CHU Brugmann, Brussels, Belgium; 3 Department of Hepatogastroenterology,<br />

CHU Sart Tilman, Liège, Belgium; 4 Department<br />

of Gastroenterology, Hepatopancreatology and Digestive oncology,<br />

Cliniques Universitaires Saint-Luc, Brussels, Belgium; 5 Department<br />

of Liver and Biliopancreatic disorders, University Hospitals<br />

Leuven, Leuven, Belgium; 6 Department of Gastroenterology, Hepatopancreatology<br />

and Digestive oncology, CHIREC, Brussels,<br />

Belgium; 7 Department of Hepatogastroenterology, UZ Brussel,<br />

Brussels, Belgium; 8 Stuivenberg, ZNA, Antwerp, Belgium; 9 Grand<br />

hôpital de Charleroi, Charleroi, Belgium; 10 Department of Gastroenterology<br />

and Hepatology, University Hospital Antwerp, Edegem,<br />

Belgium; 11 Department of Hepatogastroenterology, Ghent University<br />

Hospital, Ghent, Belgium; 12 AZ Maria-Middelares, Ghent,<br />

Belgium; 13 Hôpitaux Iris Sud, Brussels, Belgium; 14 Department of<br />

Gastroenterology and Hepatology, AZ St Jan, Brugge, Belgium;<br />

15 Department of Hepatogastroenterology, CHU Saint-Pierre, Brussels,<br />

Belgium<br />

Background: All-oral, interferon-free regimens that combine<br />

direct-acting antiviral drugs have significantly advanced the<br />

treatment of hepatitis C (HCV), especially for genotype 1(G1)<br />

patients. However, efficacy and safety data of interferon-free<br />

regimens in HCV genotype 4 (G4) patients are scarce. In Belgium,<br />

Sofosbuvir (SOF) and Simeprevir (SMV) treatment is<br />

available since January 2015 for G4 patients with advanced<br />

fibrosis (F3-F4 METAVIR) for 12 weeks. Methods: analysis of<br />

HCV G4 patients receiving SOF and SMV treatment in Belgium.<br />

The aim of the study was to evaluate the safety and efficacy<br />

of the treatment. Results: 73 G4 patients were enrolled in<br />

this data collection including 32 (43.8%) patients with severe<br />

fibrosis F3 and 41(56.2%) cirrhotic patients. The study population<br />

comprised 58.9% male, 77.8% treatment experienced<br />

patients. Median age was 59 [51-66] years and 5 patients<br />

were HCV/HIV co-infected. 24 patients received the treatment<br />

associated with ribavirin, 11/32 (34.37%) of patients with<br />

advanced fibrosis and 13/41 (31.71%) of cirrhotic patients. In<br />

cirrhotic patients, median MELD and Child-Pugh score were 9<br />

[7-12.5] and 5 [5-6], 46.2% had platelet below 100.000/mm<br />

and 28.6% had albumin below 35 g/L. W4 HCV RNA was<br />

undetectable in 31.25% (15/48). 9 of the 15 patients with<br />

undetectable W4 HCV RNA received RBV. At W12, 100%<br />

(23/23) had HCV RNA below the limit of quantification, with<br />

6/23 still detectable. All SVR12 data will be available at the<br />

time of presentation. No patient experienced serious adverse<br />

event. Conclusions: these preliminary results in difficult-to-treat<br />

G4 HCV patients show that SOF/SIM +/- RBV treatment is safe<br />

and seems promising, in line with that was observed in G1<br />

HCV patients.<br />

Disclosures:<br />

Christophe Moreno - Consulting: Abbvie, Janssen, Gilead, BMS; Grant/Research<br />

Support: Janssen, Gilead, Roche, Astellas<br />

Hendrik Reynaert - Advisory Committees or Review Panels: MSD, Gillead, Janssen,<br />

BMS, Abbvie, Norgine; Grant/Research Support: Roche<br />

Stefan Bourgeois - Advisory Committees or Review Panels: AbbVIe, Gilead; Consulting:<br />

Roche, MSD; Speaking and Teaching: Janssen, BMS<br />

The following authors have nothing to disclose: Luc Lasser, Jean Delwaide, Peter<br />

Starkel, Wim Laleman, Philippe Langlet, Sergio Negrin Dastis, Thierry Gustot,<br />

Sven M. Francque, Anja M. Geerts, Christophe Van Steenkiste, Chantal de<br />

Galocsy, Collins Assene, Hans Orlent, Marcel Nkuize, Delphine Degré<br />

1092<br />

Efficacy and safety of IFN-free, DAA-based treatment-regimens<br />

in patients with HCV-recurrence after<br />

Liver Transplantation: Real-life experience from Austria<br />

Sandra Beinhardt 1 , Ramona Al Zoairy 2 , Clarissa Freissmuth 1 , Karin<br />

Kozbial 1 , Stephanie Hametner 3 , Rafael Stern 1 , Rudolf E. Stauber 5 ,<br />

Andreas Maieron 3 , Katharina Staufer 4 , Michael P. Strasser 6 , Heinz<br />

M. Zoller 2 , Ivo Graziadei 7 , Wolfgang Vogel 8 , Markus Peck-Radosavljevic<br />

1 , Michael Trauner 2 , Peter Ferenci 1 , Harald Hofer 1 ;<br />

1 Gastroenterology and Hepatology, Medical University of Vienna,<br />

Vienna, Austria; 2 Gastroenterology and Hepatology, Department<br />

of Internal Medicine II, Innsbruck, Austria; 3 Hosital Elisabethinen,<br />

Linz, Austria; 4 Department of Transplant Surgeries, Medical University<br />

of Vienna, Vienna, Austria; 5 Department of Gastroenerology<br />

and Hepatology, Medical University of Graz, Graz, Austria;<br />

6 Department of Gastroenerology and Hepatology, Paracelsus<br />

Medical University Salzburg, Salzburg, Austria; 7 Departmen of<br />

Internal Medicine, LKH Hall/Tirol, Hall/Tirol, Austria; 8 Department<br />

of Internal Medicine II, Division of Gastroenterology and Hepatology,<br />

Medical University of Innsbruck, Innsbruck, Austria<br />

Background&Aims: Patients after orthotopic liver transplantation<br />

(OLT) and hepatitis C virus-recurrence have high need<br />

of antiviral therapy (TX) due to accelerated progression of<br />

disease. IFN-free regimens revealed promising efficacy- and<br />

safety-data in clinical <strong>studies</strong>. This study aimed to evaluate<br />

real-life efficacy- and safety-data of DAA-regimens in patients<br />

with HCV-recurrence after OLT. Patients&Methods: 96 OLT-patients<br />

(10:liver-/kidney-transplantation; male:79/82.3%;<br />

age:59±14 years [mean±SD]; 34-78 [range]; GT-1:69<br />

[1a:17; 1b:49; 1g:1; undetermined:2]; GT3a:17; GT-4:8;<br />

missing:2) started all-oral DAA-regimens. Time from OLT to<br />

TX was 67±6 (mean±SD; range: -8–304) months. 8 patients<br />

had fibrosing cholestatic hepatitis (FHC), 49 cirrhosis (CPS-<br />

A:33; CPS-B/C:10; FIB4-Score/BL:6.9±8.7 [mean±SD]; TX-experienced:65/68%).<br />

Immunosuppressive-TX was in 37/39%<br />

patients based on calcineurin- and 55/57% on m-TOR-inhibitors.<br />

DAA-regimens were: Sofosbuvir(SOF)/ribavirin(RBV):26,<br />

SOF/Daclatasvir(DCV):43, SOF/Simeprevir(SIM):12, SOF/<br />

Ledipasvir(LDV)±RBV:10; 3D-combination (Paritaprevir/r&Om-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 747A<br />

bitasvir&Dasabuvir):4; SIM/DCV:1. HCV-RNA-quantification<br />

was performed with Abbott RealTime HCV ([ART], lower limit of<br />

quantification [LLOQ]:12IU/ml) or Roche COBAS AmpliPrep/<br />

COBAS TaqMan assay (LLOQ:15IU/ml). Results: At present<br />

4/4.2% patients reached TX-week4, 14/15% end of treatment<br />

(EoT), 78/81% end of follow-up (12 weeks after TX). Overall<br />

SVR12-rate was 83% (65/78); table). 2 FCH-patients died<br />

at week 8 and 20, respectively; 2 interrupted treatment; one<br />

was transplanted at week10, the other hospitalized for hepatic<br />

encephalopathy; treatment was stopped at week8, the patient<br />

achieved SVR12. No episode of graft-rejection was observed.<br />

In SVR12-patients transaminases decreased compared to<br />

baseline (BL/SVR12:ALT:117±93 vs. 56±35; AST:113±95<br />

vs. 36±35,p


748A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Christine Jacomet - Advisory Committees or Review Panels: ANSM; Consulting:<br />

CONVERGENCE; Grant/Research Support: JANSSEN, ROCHE, MSD; Speaking<br />

and Teaching: ABBOTT, BMS, MSD, JANSSEN, VIIV, GILEAD<br />

Claudine C. Duvivier - Grant/Research Support: Gilead Sciences<br />

Laurent Cotte - Grant/Research Support: MSD, ViiV<br />

The following authors have nothing to disclose: Isabelle Poizot-Martin, Alissa<br />

Naqvi, Véronique Obry-Roguet, Eric Billaud, Antoine Chéret, David Rey, Pascal<br />

Pugliese, Pierre Pradat<br />

1094<br />

Sofosbuvir Plus Ribavirin Without Interferon For Treatment<br />

of Acute Hepatitis C Virus Infection in HIV-1<br />

infected Individuals (SWIFT-C)<br />

Susanna Naggie 1 , Kristen M. Marks 2 , Michael Hughes 3 , Daniel S.<br />

Fierer 4 , Arthur Y. Kim 5 , Kimberly Hollabaugh 3 , Jennifer Kiser 6 , Jhoanna<br />

Roa 3 , Bill Symonds 7 , Diana M. Brainard 8 , John G. McHutchison<br />

8 , Marion G. Peters 9 , Raymond T. Chung 5 ; 1 Duke University<br />

Medical Center, Durham, NC; 2 Weill Cornell, New York, NY;<br />

3 Harvard, Cambridge, MA; 4 Mount Sinai School of Medicine,<br />

New York, NY; 5 Massachusetts General Hospital, Boston, MA;<br />

6 University of Colorado, Denver, CO; 7 Roivant, New York, NY;<br />

8 Gilead Sciences, Inc, Foster City, CA; 9 University of California,<br />

San Francisco, San Francisco, CA<br />

Background: The role of interferon-free, oral direct acting antivirals<br />

(DAA) for the treatment of acute hepatitis C virus (HCV)<br />

infection in HIV-infected persons has not been explored. Methods:<br />

SWIFT-C is a multicenter, single arm trial of the AIDS Clinical<br />

Trials Group (ACTG) investigating the safety and efficacy<br />

of 12 weeks of sofosbuvir (400mg/day) and weight based<br />

ribavirin (1000mg


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 749A<br />

dL (IQR: 0.5 – 0.9 mg/dL). Of the patients with data available<br />

12-weeks after the end-of-treatment, 43/48 (90%) achieved<br />

SVR12. Of the 5 patients who failed, 4 relapsed within the first<br />

4 weeks after the end-of-treatment. One patient was sent to hospice<br />

after a hospitalization for acute cholangitis with elevated<br />

total bilirubin, rectal bleeding, and worsening renal function.<br />

The average cost of pharmaceuticals per patient was $90,234.<br />

The average cost-per-SVR was $100,727. Compared to treatment<br />

with SMV/SOF, the cost-per-SVR with LDV/SOF was on<br />

average $68,337 lower. Conclusions: We observed high SVR<br />

rates (90%) with LDV/SOF in standard clinical practice. The<br />

cost-per-SVR was $100,000 and was approximately $68,000<br />

less than previous treatments. Future investigations of real-world<br />

SVR rates are needed.<br />

Disclosures:<br />

Kian Bichoupan - Consulting: Janssen, Gilead<br />

Alyson Harty - Advisory Committees or Review Panels: Gilead; Consulting: Gilead,<br />

Jannsen, Acaria Pharmacy, Abbvie<br />

David B. Motamed - Advisory Committees or Review Panels: Gilead Pharmaceuticals<br />

Viktoriya Khaitova - Advisory Committees or Review Panels: Gilead, Johnson<br />

and Johnson<br />

Charissa Y. Chang - Consulting: Gilead, Vertex, Onyx<br />

Jennifer Leong - Advisory Committees or Review Panels: Gilead; Consulting:<br />

Bayer<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

Albert Min - Consulting: Bristol Myers Squibb, Gilead, Janssen, Merck; Grant/<br />

Research Support: Bristol Myers Squibb, Gilead; Speaking and Teaching: Bristol<br />

Myers Squibb, Gilead<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

The following authors have nothing to disclose: Daniel J. Waintraub, Neal M.<br />

Patel, Sweta Chekuri, Joshua Hartman, Alicia Stivala, Angela Woody, Meena<br />

B. Bansal, Priya Grewal, Ritu Agarwal, Gene Y. Im, Lawrence Liu, Joseph A.<br />

Odin, Nancy Bach, Thomas D. Schiano, Ponni V. Perumalswami, Lan S. Wang,<br />

Jahnavi Naik<br />

1096<br />

IFN-λ inhibits miR-122 transcription through a Stat3-<br />

HNF4a inflammatory feedback loop in an IFN-α resistant<br />

HCV cell culture system<br />

Fatma Aboulnasr 1 , Sidhartha Hazari 1 , Partha K. Chandra 1 , Pauline<br />

Ferraris 1 , Rajesh Panigrahi 1 , Srinivas Chava 1 , Ramazan Kurt 1 ,<br />

Kyoungsub Song 1 , Asha Dash 1 , Luis A. Balart 2 , Tong Wu 1 , Srikanta<br />

Dash 1,2 ; 1 Pathology, Tulane University, New Orleans, LA;<br />

2 Medicine, Division of Gasteroenterology and Hepatology, Tulane<br />

University, New Orleans, LA<br />

Background: HCV replication in persistently infected cell culture<br />

remains resistant to IFN-α /RBV combination treatment,<br />

whereas IFN-λ1 induces viral clearance. The antiviral mechanisms<br />

by which IFN-λ1 induces sustained HCV clearance<br />

have not been determined. Aim: To investigate the mechanisms<br />

by which IFN-λ clears HCV replication in an HCV cell<br />

culture model. Methods: IFN-α sensitive (S3-GFP) and resistant<br />

(R4-GFP) cells were treated with equivalent concentrations of<br />

either IFN-λ1 or IFN-α.The relative antiviral effects of IFN-α<br />

and IFN-λ1 were compared by measuring the HCV replication,<br />

quantification of HCV-GFP expression by flow cytometry, and<br />

viral RNA levels by real time RT-PCR. Activation of Jak-Stat<br />

signaling, interferon stimulated gene (ISG) expression, and<br />

miRNA-122 transcription in S3-GFP and R4-GFP cells were<br />

examined. Results: We have shown that IFN-λ1 induces HCV<br />

clearance in IFN-α resistant and sensitive replicon cell lines<br />

in a dose dependent manner through Jak-Stat signaling, and<br />

induces STAT1 and STAT2 activation, ISRE-luciferase promoter<br />

activation and ISG expression. Stat3 activation is also involved<br />

in IFN-λ1 induced antiviral activity in HCV cell culture. IFNλ1<br />

induced Stat3 phosphorylation reduces the expression of<br />

hepatocyte nuclear factor 4 alpha (HNF4α) through miR-24 in<br />

R4-GFP cells. Reduced expression of HNF4α is associated with<br />

decreased expression of miR-122 resulting an anti-HCV effect.<br />

Northern blot analysis confirms that IFN-λ1 reduces miR-122<br />

levels in R4-GFP cells. Our results indicate that IFN-λ1 activates<br />

the Stat3-HNF4a feedback inflammatory loop to inhibit miR-<br />

122 transcription in HCV cell culture.<br />

Disclosures:<br />

Luis A. Balart - Advisory Committees or Review Panels: abbvie, Genentech;<br />

Grant/Research Support: Merck, Genentech, Bayer, conatus, Ocera, Hyperion,<br />

Gilead Sciences, Bristol Myers Squibb, abbvie, Vertex, Merck, Genentech,<br />

Bayer, Conatus, Ocera, Hyperion, Gilead Sciences, Bristol Myers Squibb,<br />

Mochida, Eisai, Vertex, takeda, GI Dynamics, tobira; Speaking and Teaching:<br />

Merck, Merck, Merck, Merck, Abbvie, janssen<br />

The following authors have nothing to disclose: Fatma Aboulnasr, Sidhartha<br />

Hazari, Partha K. Chandra, Pauline Ferraris, Rajesh Panigrahi, Srinivas Chava,<br />

Ramazan Kurt, Kyoungsub Song, Asha Dash, Tong Wu, Srikanta Dash<br />

1097<br />

Efficacy of Daclatasvir/Asunaprevir According to Resistance<br />

Associated Variants in Chronic Hepatitis C Genotype1b<br />

Etsuko Iio 1 , Noritomo Shimada 2 , Hiroshi Abe 3 , Masanori Atsukawa<br />

4 , Kai Yoshiza 5 , Koichi Takaguchi 6 , Yuichiro Eguchi 7 , Hideyuki<br />

Nomura 8 , Tomoyuki Kuramitsu 9 , Jong-Hon Kang 10 , Takeshi<br />

Matsui 10 , Noboru Hirashima 11 , Atsunori Kusakabe 12 , Yasuhito<br />

Tanaka 1 ; 1 Nagoya City University Graduate School of Medical<br />

Sciences, Nagoya, Japan; 2 Ootakanomori Hospital, Kashiwa,<br />

Japan; 3 Jikei University School of Medicine Katsushika Medical<br />

Center, Tokyo, Japan; 4 Nippon Medical School Chiba Hokusoh<br />

Hospital, Chiba, Japan; 5 Machida Municipal Hospital, Tokyo,<br />

Japan; 6 Kagawa Prefectural Central Hospital, Takamatsu, Japan;<br />

7 Saga University Hospital, Saga, Japan; 8 Shin-Kokura Hospital,<br />

Kitakyushu, Japan; 9 Kuramitsu Clinic, Akita, Japan; 10 Teine Keijinkai<br />

Hospital, Sapporo, Japan; 11 National Hospital Organization<br />

Nagoya Medical Center, Nagoya, Japan; 12 Nagoya Red<br />

Cross Hospital, Nagoya, Japan<br />

[Background/Aims] Interferon (IFN)-free daclatasvir (DCV)<br />

and asunaprevir (ASV) therapy for 24 weeks improved efficacy<br />

and safety outcomes for patients with hepatitis C virus<br />

(HCV) infection. The aim of the present study is to assess the<br />

treatment outcome according to resistance associated variants<br />

(RAVs) in NS3/NS5A regions among patients with chronic<br />

HCV genotype 1b infection (HCV-1b). [Methods] 612 patients<br />

with HCV-1b were enrolled at multi centers in Japan. This<br />

study protocol was approved by the appropriate institutional<br />

ethics review committees and written informed consents were<br />

observed. Direct sequencing in NS5A region was performed<br />

for all 612 patients before DCV/ASV therapy. 393 chronic<br />

hepatitis (CH) and 219 liver cirrhosis (LC) patients. The median<br />

age was 71 (30-87) and male was 267 (43.6%). The median<br />

platelet count was 12.6×10 4 /μL and HCV-RNA was 6.1 log<br />

IU/mL. 36 patients had experience of Peg-IFN/RBV/Protease<br />

inhibitors (PI) combination therapy, and 27 with simeprevir<br />

(SMV). [Results] Direct sequencing analysis of 612 patients<br />

administered DCV/ASV therapy showed the existence of 31M<br />

(2.4%), 93H (4.1%), and 31M/93H (0.3%) variants in NS5A<br />

region. The 531 patients received over the 4 weeks of DCV/


750A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ASV; 75.0% (398/531) achieved rapid virologic response<br />

(RVR) and 97.0% (450/482) became undetectable until 8<br />

week. There was no significant differences of RVR among<br />

IL28B genotypes. 76.9% (380/494) with wild-type NS5A-Y93<br />

achieved RVR, whereas only 54.2% (13/24) with Y93H substitution<br />

achieved RVR. In 93Y group, there were 22 viral breakthrough<br />

and 3 relapsers, Among these patients with treatment<br />

failure, 40.0% (10/25) had SMV treatment history. On the<br />

other hand, among patients with Y93H at baseline, 5 patients<br />

(20.1%) had virologic failure and 1 patient experienced<br />

relapse, resulting in emergence of NS5A-Y93 and NS3-D168<br />

variants. Finally, SVR (sustained virologic response) 4 was<br />

89.6% in Y93, whereas 45.5% in Y93H, suggesting that RVR<br />

and SVR were higher in 93Y than 93H group. Interestingly,<br />

multiple RAVs such as L31M/Q54H/Y93H in NS5A were<br />

emerged by treatment failure. The number of RAVs were 0/ 1/<br />

2/ 3=57/ 36/ 7/ 0 (%) at baseline, and 0/ 14/ 57/ 29 (%)<br />

at the time of virologic failure; patients with multiple RAVs in<br />

NS5A region increased from 7% to 86%. [Conclusion] History<br />

of SMV therapy and pre-existing NS5A-Y93H substitution were<br />

associated with breakthrough during DCV/ASV therapy, resulting<br />

in emergence of multiple RAVs. As the multi-RAVs influence<br />

SVR rate by DAAs therapy including NS5A inhibitors, patients<br />

with RAVs at baseline will be required assessment for optimizing<br />

future DAAs therapies.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Etsuko Iio, Noritomo Shimada,<br />

Hiroshi Abe, Masanori Atsukawa, Kai Yoshiza, Koichi Takaguchi, Yuichiro Eguchi,<br />

Hideyuki Nomura, Tomoyuki Kuramitsu, Jong-Hon Kang, Takeshi Matsui,<br />

Noboru Hirashima, Atsunori Kusakabe<br />

1098<br />

Ribavirin levels at treatment weeks (TW) 4-6 can predict<br />

sustained viriological response (SVR12) in HCV cirrhotic<br />

patients treated with all oral direct–acting antiviral therapy<br />

(DAAs).<br />

Suman Verma 1 , Ivana Carey 1 , Kate E. Childs 1 , Andrew Ayers 1 ,<br />

Phillip Morgan 1 , Aisling B. Considine 1 , Kath Oakes 1 , Michael<br />

A. Heneghan 1 , Graham R. Foster 2 , Kosh Agarwal 1 ; 1 Institute of<br />

liver <strong>studies</strong>, Kings College Hospital, London, United Kingdom;<br />

2 The Blizard Institute, Queen Marys University of London, London,<br />

United Kingdom<br />

Background Ribavirin remains an important adjunct to DAAs<br />

in end-stage cirrhotic patients (Child–Pugh(CP) B7 and above).<br />

However ribavirin associated side-effects can be challenging<br />

to manage in this population (CUPIC). Hence this study aimed<br />

to investigate whether plasma ribavirin level monitoring in<br />

this advanced liver disease population can optimise SVR12.<br />

Method 46 patients, with CP B7-C10 cirrhosis, receiving ribavirin,<br />

Sofosbuvir and either Ledipasvir or Daclatasvir had<br />

serial trough plasma ribavirin measurements at TW2, 4, 6 and<br />

8 using a validated liquid chromatography assay. At TW0,<br />

a median of 1000mg/day (range 1000mg/day-1200mg/<br />

day) ribavirin was administered to all patients. Despite high<br />

inter-patient variability in ribavirin levels with the same dose,<br />

there was low intra-patient variation with steady-state achieved<br />

between TW4-TW6 and the median dose remained 1000mg/<br />

day. Hence levels at this timepoint were compared between<br />

those acheiving SVR12 vs. relapsing post treatment. All statistical<br />

analysis was performed using SPSS. Results 37 patients<br />

had SVR12 and 9 relapsed, but no significant difference in CP,<br />

UKELD, MELD, prescribed ribavirin dose (median 1000mg/<br />

day) or ribavirin dose/kg at TW0 and TW4-6 was identified<br />

(Table 1). 25 patients with SVR12 and 2 relapsers had ribavirin<br />

dose reduction during 12 weeks of therapy but none<br />

required erythropoetin or blood transfusion support. Age, CP,<br />

UKELD, MELD, and genotype showed no significant correlation<br />

with SVR12 or relapse. However, the median plamsa ribavirin<br />

level was higher in the SVR12 cohort vs. relapsers (2.4mg/L<br />

and 1.4mg/L respectively (p=0.003) and levels ≥1.7mg/L at<br />

TW4-TW6 correlated with SVR12 achievement (PPV 89.19%;<br />

NPV 77.18%)(p=0.001 Fisher’s exact test). Ribavirin dose<br />

reduction thereafter did not influence SVR12 development.<br />

Conclusion Ribavirin levels at 4-6 weeks of a 12 week DAA<br />

regimen are predictive of SVR12 in this advanced liver disease<br />

cohort. Ribavirin plasma measurements can be a useful adjunct<br />

in delivering optimal SVR12 and tolerability in this advanced<br />

challenging population.<br />

SVR12 and relapser cohort demographics (median and ranges<br />

shown)<br />

Disclosures:<br />

Ivana Carey - Grant/Research Support: Gilead, Roche; Speaking and Teaching:<br />

BMS<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

Kosh Agarwal - Advisory Committees or Review Panels: Gilead, Novartis,<br />

Abbott; Grant/Research Support: Roche, MSD; Speaking and Teaching: BMS,<br />

Astellas, Janssen<br />

The following authors have nothing to disclose: Suman Verma, Kate E. Childs,<br />

Andrew Ayers, Phillip Morgan, Aisling B. Considine, Kath Oakes, Michael A.<br />

Heneghan<br />

1099<br />

Frequency of Renal Impairment in Patients With Hepatitis<br />

C Infection Treated With Sofosbuvir-based Antiviral<br />

Regimens<br />

Saeed Almarzooqi 1 , Jagpal S. Klair 2 , Joel G. Karkada 1 , Raoel<br />

Maan 1 , Orlando Cerocchi 1 , Matthew Kowgier 1 , Sherrie M. Harrell<br />

2 , Kimberly Rhodes 2 , Harry L. Janssen 1 , Jordan J. Feld 1 , Andres<br />

Duarte-Rojo 2 ; 1 Toronto Center for Liver Disease, Toronto, ON, Canada;<br />

2 University of Arkansas for Medical Sciences, Little Rock, AR<br />

BACKGROUND: Renal impairment (RI) was a relevant adverse<br />

event noted in 5-7% of hepatitis C patients treated with the<br />

protease inhibitors boceprevir/telaprevir (BOC/TPV), although<br />

it was infrequent with peginterferon/ribavirin (PR). RI has been<br />

very rarely reported in trials with newer direct-acting antivirals<br />

such as simeprevir (SMV) and sofosbuvir (SOF). However, SOF<br />

is renally cleared and it might cause RI in patients with pre-existing<br />

kidney disease. We aimed to examine the rate of RI in<br />

patients treated with SOF-based therapies, and to compare it<br />

to that of BOC/TPV. METHODS: Patients treated with any SOFbased<br />

or BOC/TPV-based therapies at two university-based<br />

referral centers were included. Demography, presence of cirrhosis,<br />

ascites, original MELD score, SVR12, comorbidities,<br />

use of nephrotoxic drugs and on-treatment changes in creatinine<br />

(Cr) were obtained from medical records. Posttransplant


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 751A<br />

patients were excluded. RI was defined as an increase in Cr<br />

≥50% from baseline while on-treatment. Mann-Whitney U,<br />

Kruskal-Wallis, chi-square, and logistic regression were used<br />

for analysis. RESULTS: A total of 456 patients (55±9 years,<br />

61% males, cirrhotic 50%, HIV 3%) were included (G1: 78%,<br />

G2: 13%, G3: 7%, G4&6: 2%). BOC/TPV was used in 224<br />

patients (50%); SOF+PR in 76 (17%); SOF+R in 75 (17%);<br />

SMV/SOF±RBV in 66 (15%); and LDV/SOF in 11 (2%).<br />

Patients on PR-based therapies were less likely cirrhotic (44% vs<br />

60%, p=0.001) and had lower MELD scores when compared<br />

to those on all-oral therapies (4 [2-6] vs 6 [3-9], p=0.0001). In<br />

total, 95% had a baseline eGFR ≥60 mL/min (Cr 0.83±0.16<br />

mg/dL). Among them, RI was noted in 7% of patients (n=15)<br />

on BOC/TPV, 5% of SOF-PR (n=3), and 4% of all-oral regimens<br />

(n=5), (p=0.4). On-treatment CrMax in patients on BOC/<br />

TPV was 1.4 (1.2-2) mg/dL, on SOF-PR 2 (1.2-2), and on alloral<br />

regimens 2 (1.35-3.2) (p=0.04). Although cirrhosis and<br />

MELD were linked to RI on univariate analysis, only ascites<br />

(OR=3.16 [1.14-8.92]) and preexisting proteinuria (OR=5.74<br />

[2.04-16.15]) remained significant in multivariate models. In<br />

all cases, serum Cr returned to baseline after stopping therapy.<br />

SVR12 did not differ between those who did or did not<br />

develop RI (88% vs 86%, p=0.9). CONCLUSIONS: Reversible<br />

RI in SOF-based therapies was seen in 4-5% in this cohort with<br />

a high prevalence of cirrhosis, similar to rates reported with<br />

BOC/TPV-based regimens. RI was more frequent in patients<br />

with advanced liver disease, mainly in the presence of ascites<br />

and in those with preexisting kidney injury. Monitoring of<br />

renal function and standard nephroprotective measures are<br />

suggested when considering SOF-based regimens.<br />

Disclosures:<br />

Raoel Maan - Consulting: AbbVie<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Jordan J. Feld - Advisory Committees or Review Panels: Merck, Janssen, Gilead,<br />

AbbVie, Theravance, Bristol Meiers Squibb; Grant/Research Support: AbbVie,<br />

Boehringer Ingelheim, Janssen, Gilead, Merck<br />

Andres Duarte-Rojo - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Vital Therapies; Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Saeed Almarzooqi, Jagpal S.<br />

Klair, Joel G. Karkada, Orlando Cerocchi, Matthew Kowgier, Sherrie M. Harrell,<br />

Kimberly Rhodes<br />

1100<br />

Prospective Study for The Efficacy of Sofosbuvir and<br />

Simeprevir ± Ribavirin in Hepatitis C Genotype 1 and<br />

4 Compensated Cirrhotic Patients. Single Center Study<br />

and Real Life Experience<br />

Zeid Kayali 1,2 , Cory Amador 2 , Andrew Lowe 2 , Besher Ashouri 1,2 ,<br />

Katherine Lam 2 , Warren N. Schmidt 3 ; 1 Inland Empire Liver Foundation,<br />

Rialto, CA; 2 Arrowhead Regional Medical Center, Colton,<br />

CA; 3 University of Iowa Hospital and Clinics, Iowa City, IA<br />

Sofosbuvir (SOF) and simeprevir (SMV) ± ribavirin (RBV)<br />

have been used widely to treat hepatitis C genotype 1 and 4<br />

patients and showed 94% SVR12 in patients with stage 3 and<br />

4 fibrosis (COSMO study). The efficacy of this regimen, however,<br />

has never been validated in a large number of cirrhotic<br />

patients from a community, multiracial setting. The goal of this<br />

study was to investigate the performance of this regimen in a<br />

large, diverse group of difficult to treat cirrhotic patients. Methods:<br />

This was a prospective, open-label single center study.<br />

Enrolled subjects were consecutive patients with cirrhosis, genotype<br />

1 and 4, including both naïve and treatment experienced<br />

patients. Cirrhosis (Stage 4 fibrosis) was diagnosed based on<br />

biopsy and/or Fibrosure test. Decompensated cirrhotics, HIV<br />

co-infected, hepatocellular carcinoma, and active substance<br />

abuse patients were excluded. Duration of treatment was either<br />

24 weeks without RBV or 12 weeks with ribavirin. Primary<br />

end point was SVR12. Secondary end points were safety and<br />

relapse. Results: One hundred eight patients were enrolled.<br />

Mean age 54 year, 63 (58%) were male, 36 (33%) Caucasians,<br />

25 (21%) Black and 50 (46%) Hispanics. mean BMI 33,<br />

86 (79.6%) patients had HCV GT1a, and 7 (7%) had GT4,<br />

mean baseline viral load was 2.2 x 10 6 IU/ml, 59 (54%) were<br />

treatment naive. All patients completed treatment. No Grade 4<br />

adverse events were reported. HCV RNA was undetected in 90<br />

patients (83%) at end of treatment. Eighty-four patients (77%)<br />

achieved SVR12 and 24 patients (23%) relapsed. There was<br />

no change in MELD score or worsening decompensation at end<br />

of treatment. Univariate regression analysis showed that BMI<br />

(33) and Black race were independent factors associated with<br />

relapse (p=0.004,95%CI 1-1.22) and (p=0.025,95%CI 0.03-<br />

0.23) respectively. Duration of treatment, sex, age, baseline<br />

viral load, Geno1 subtypes and MELD score were not independent<br />

factors that predict relapse. Conclusion: In this large<br />

study of difficult to treat compensated cirrhotics patients, SOF<br />

and SMV ± RBV regimen was well tolerated but had lower SVR<br />

12 than previously reported. Surprisingly Black race and BMI<br />

were independent factors for relapse and these variables need<br />

to be considered when deciding the best regimen for difficult<br />

to treat patients.<br />

Disclosures:<br />

Zeid Kayali - Consulting: Abbvie; Grant/Research Support: Merck, Gilead<br />

Warren N. Schmidt - Consulting: gilead<br />

The following authors have nothing to disclose: Cory Amador, Andrew Lowe,<br />

Besher Ashouri, Katherine Lam<br />

1101<br />

Sofosbuvir + Ledispasvir Combination Therapy for<br />

Recurrent Hepatitis C in Liver Transplant Recipients: A<br />

Real-Life Multicenter Experience<br />

Ryan M. Kwok 1 , Suzanne Robertazzi 1 , Helen S. Te 2 , Joshua Wiegel<br />

3 , Joseph Ahn 3 , Janet Gripshover 4 , Darryn R. Potosky 4 , Amber<br />

Tierney 5 , Mohamed A. Hassan 5 , Rohit Satoskar 1 , Coleman I.<br />

Smith 1 ; 1 Transplant Institute, Medstar Georgetown University Hospital,<br />

Washington, DC; 2 Medicine, University of Chicago, Chicago,<br />

IL; 3 Oregon Health and Science University, Portland, OR;<br />

4 Division of Gastroenterology and Hepatology, University of Maryland<br />

School of Medicine, Baltimore, MD; 5 Gastroenterology Division,<br />

University of Minnesota Medical School, Minneapolis, MN<br />

Background Recurrent hepatitis C (HCV) after liver transplant<br />

(LT) is associated with significant morbidity and mortality.<br />

Second generation direct acting antiviral medications such as<br />

sofosbuvir / ledipasvir (SOF/LDV) have a high rate of sustained<br />

virologic response in treating non-cirrhotic and pretransplant<br />

HCV. Data on the efficacy and safety of SOF/LDV in the post-LT<br />

patient are limited. Aim: To evaluate the safety and efficacy of<br />

the SOF/LDV therapy in patients with HCV recurrence after LT.<br />

Methods All post-LT patients with recurrent HCV . previously or<br />

currently being treated with SOF/LDV with or without ribavirin<br />

for 12-24 weeks were identified at five centers. Sustained<br />

virologic response at 12 weeks (SVR12) after treatment was<br />

determined. On-treatment response, graft function, changes in<br />

immunosuppression, adverse events (AE), and survival were<br />

recorded. Results 128 patients were included: 76% male,<br />

65% Caucasian with a mean age of 60.9 ±6.8 years. 70% of<br />

patients were genotype (GT) 1a, 23% GT 1b, 1 patient was<br />

a GT 2, 2 patients were GT 4, 1 patient had both GT 1a/4.<br />

10% had kidney transplant, 52% had failed prior therapy, and


752A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

18% had F3-4 fibrosis. The median time from LT to SOF/LDV<br />

was 1398 days (15-7593 days). Tacrolimus was the primary<br />

immunosuppression in 86% of patients. 59% (44/74) of treatment<br />

week 4 (TW4) and 100% (51/51) of TW12 patients had<br />

undetectable viral load. 100% (12/12) achieved SVR12. 15%<br />

required changes in immunosuppression during or after treatment<br />

with 1 patient experiencing mild rejection at TW4 that<br />

was controlled with an increase in oral immunosuppression.<br />

Adverse events were reported in 15% of patients with constitutional<br />

symptoms being the most common (9%) and cough,<br />

rash, and headaches in 2%. In the 44 patients who received<br />

ribavirin, 50% developed anemia with only 3 patients requiring<br />

blood transfusion or erythropoeisis-stimulating agents. No<br />

serious adverse events were reported. Conclusions SOF / LDV<br />

is effective and safe in post-LT patients with recurrent HCV infection<br />

across a spectrum of genotypes. No significant changes<br />

in immunosuppression or graft function were noted. Adverse<br />

events were mild and were predominantly associated with the<br />

use of ribavirin. Additional SVR data will be reported. Further<br />

large, prospective trials are required to confirm these results.<br />

Disclosures:<br />

Helen S. Te - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Abbvie, Conatus, BMS<br />

Joseph Ahn - Advisory Committees or Review Panels: gilead; Grant/Research<br />

Support: bms<br />

Janet Gripshover - Advisory Committees or Review Panels: Gilead, Abbvie; Consulting:<br />

Abbvie; Speaking and Teaching: Gilead, Abbvie<br />

Darryn R. Potosky - Advisory Committees or Review Panels: Gilead, Abbvie<br />

Mohamed A. Hassan - Speaking and Teaching: GILEAD<br />

Coleman I. Smith - Advisory Committees or Review Panels: Vertex, Gilead, Janssen;<br />

Grant/Research Support: Gilead, Abbvie, Janssen, Salix, BMS, Merck, Intercept<br />

Pharma, Lumena Pharma; Speaking and Teaching: Merck, Vetex, Gilead,<br />

Bayer/Onyx, BMS, Abbvie, Janssen<br />

The following authors have nothing to disclose: Ryan M. Kwok, Suzanne Robertazzi,<br />

Joshua Wiegel, Amber Tierney, Rohit Satoskar<br />

1102<br />

Hepatitis C treatment with Sofosbuvir and NS5A inhibitor<br />

in patients with an addiction<br />

Jean-Baptiste Trabut 5 , Camille Barrault 1 , Hélène Charlot 2 , Damien<br />

Carmona 3 , Willy Kini-Matondo 2 , Franck Questel 4 , Christophe<br />

Hezode 6 ; 1 Addictologie, CHIC, Créteil, France; 2 Pharmacie,<br />

Hôpitaux Universitaires Henri Mondor, Créteil, France; 3 Centre<br />

Epice, Créteil, France; 4 Addictologie, Hôpital Fernand Widal,<br />

Paris, France; 5 Addictologie, Hôpitaux Universitaires Henri Mondor,<br />

Créteil, France; 6 Hépatologie, Hôpitaux Universitaires Henri<br />

Mondor, Créteil, France<br />

Background: Hepatitis C virus (HCV) infection is highly prevalent<br />

in patients with an addiction. Access to HCV treatment<br />

is reduced in those patients due to the poor tolerance of interferon-based<br />

therapies. There are very few data on the use of<br />

“interferon-free” oral combinations in that population. Methods:<br />

patients infected with HCV and advanced fibrosis followed in<br />

four addictions care centres were considered for a treatment<br />

combining sofosbuvir and an NS5A inhibitor (daclatasvir or<br />

ledipasvir) during 12 weeks (24 weeks for cirrhotic patients<br />

with genotype 3). A multidisciplinary team including addiction<br />

specialists, hepatologists and pharmacists reviewed all indications<br />

and decided the practical implementation of the treatments.<br />

Results: Treatment has been introduced in 39 patients<br />

(Male: 87%, Median age: 51, cirrhotic: 71%). Two third of<br />

the patients received a substitution for opioid dependence and<br />

78% had an alcohol use disorder. Most patients (62%) suffered<br />

from some forms of psychosocial vulnerability. Treatment has<br />

been completed to date in 22 patients with an end of treatment<br />

viral clearance in all cases. No virological relapse has been<br />

reported so far and sustained virological response has been<br />

established in 15 patients. Updated results will be presented<br />

at the congress. Conclusion: Our preliminary results suggest<br />

that the use of new oral combinations for the treatment of HCV<br />

infection could be highly effective in the addiction care setting.<br />

Disclosures:<br />

Christophe Hezode - Speaking and Teaching: Roche, BMS, MSD, Janssen, abbvie,<br />

Gilead<br />

The following authors have nothing to disclose: Jean-Baptiste Trabut, Camille<br />

Barrault, Hélène Charlot, Damien Carmona, Willy Kini-Matondo, Franck Questel<br />

1103<br />

Determination of on-treatment pharmacokinetics of<br />

sofosbuvir and ledipasvir in patients with decompensated<br />

cirrhosis - the need for real world PK <strong>studies</strong><br />

Omar El-Sherif 1,6 , Diarmaid D. Houlihan 2 , Stephen Stewart 3 ,<br />

Colm J. Bergin 4,6 , Liam J. Fanning 8 , Susan McKiernan 1,6 , Orla<br />

M. Crosbie 7 , Suzanne Norris 1,6 , Saye H. Khoo 5 ; 1 Department of<br />

Hepatology, St. James’s Hospital, Dublin, Ireland; 2 Liver Unit, St<br />

Vincents Hospital, Dublin, Ireland; 3 Mater Hospital, Dublin, Ireland;<br />

4 Department of Infectious Diseases, St James’s Hospital, Dublin,<br />

Ireland; 5 University of Liverpool, Liverpool, United Kingdom;<br />

6 Trinity College, Dublin, Ireland; 7 Cork University Hospital, Cork,<br />

Ireland; 8 University College Cork, Cork, Ireland<br />

Introduction: With the availability of interferon-free direct acting<br />

antiviral therapy (DAA) for hepatitis C, patients with decompensated<br />

cirrhosis can now be considered for antiviral therapy.<br />

The combination of sofosbuvir-ledipasvir with ribavirin was<br />

associated with high sustained virological response rates (SVR)<br />

in decompensated cirrhotics in the SOLAR-1 trial, although real<br />

world outcomes are awaited. It is unclear whether interindividual<br />

pharmacokinetic variability affects treatment outcome.<br />

Methods: Plasma trough samples were prospectively collected<br />

in 35 decompensated cirrhotic patients receiving 12 weeks<br />

of sofosbuvir-ledipasvir and ribavirin therapy for HCV genotype<br />

1 infection. GS-331007 (major circulating metabolite of<br />

sofosbuvir) and ledipasvir plasma trough samples were collected<br />

at treatment days 7, 14, 28, 84 and measured using<br />

validated LC-MS/MS. Mean age was 53.6 + 10.3 and 62%<br />

were male. The median MELD score was 10 + 3.3. Trough<br />

GS-331007 and ledipasvir concentrations were compared in<br />

patients achieving a rapid virologic response (RVR) vs no RVR<br />

and SVR4 vs non-SVR4. Results: All 35 patients achieved an<br />

on-treatment virologic response, and the SVR4 rate per protocol<br />

is 82%. GS-331007 C min<br />

ranged from 47 – 1097 ng/ml,<br />

and ledipasvir C min<br />

ranged from 36.1 – 366ng/ml. There was<br />

no relationship between trough GS-331007 levels and RVR<br />

or SVR. A trend towards higher mean day 28 ledipasvir C min<br />

levels in patients achieving SVR4 compared to early relapsers<br />

(182ng/ml vs 102.47ng/ml – p=0.062) was observed. Conclusion:<br />

The relationship between ledipasvir C min<br />

and treatment<br />

response needs to be examined in larger number of patients.<br />

These data highlight the potential utility of “real world” pharmacokinetic<br />

<strong>studies</strong> in optimising therapy and predicting treatment<br />

reponse in patients with advanced liver disease.<br />

Disclosures:<br />

Colm J. Bergin - Advisory Committees or Review Panels: Janssen, MSD, BMS,<br />

Pfizer; Grant/Research Support: MSD, Janssen, GSK, Abbott<br />

Saye H. Khoo - Grant/Research Support: Merck, Janssen, Gilead, ViiV<br />

The following authors have nothing to disclose: Omar El-Sherif, Diarmaid D.<br />

Houlihan, Stephen Stewart, Liam J. Fanning, Susan McKiernan, Orla M. Crosbie,<br />

Suzanne Norris


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 753A<br />

1104<br />

Real Life Experience of Direct Acting Anti-Viral Therapy<br />

for Hepatitis C infection in North East of Scotland<br />

Pauline Dundas, Shirley English, Lorna Bailey, Ashis Mukhopadhya,<br />

Balasubramaniam Vijayan, Andrew Fraser, Lindsay<br />

McLeman; Gastroenterology, Aberdeen Royal Infirmary, Aberdeen,<br />

United Kingdom<br />

There have been several clinical trials reported for new direct<br />

acting antiviral (DAA) drugs for hepatitis C (HCV) infection,<br />

but very few reports on ‘real life data’. We report our single<br />

centre experience of 140 patients initiated on the current generation<br />

of DAA agents in the North East of Scotland between<br />

October 2014 and May 2015. The choice and duration of<br />

regimen was decided according to local guidelines taking into<br />

consideration genotype, stage of disease, viral load, HIV coinfection,<br />

drug interaction, previous treatment and drug licences.<br />

Regimens included Simeprevir, Sofosbuvir, Daclatasvir and<br />

Sofosbuvir/Ledipasvir combinations. Sustained viral response<br />

(SVR) 4 week data will be available by September 2015 and<br />

12 week by November 2015. The regimens and patient characteristics<br />

are described in table 1. We report on 142 patient<br />

episodes of which 116 (82%) were initiated on an IFN free<br />

regimen. The mean age was 47 years (SD 10.9, range 20-79)<br />

and 109 (77%) were male. Eight (6%) patients were co-infected<br />

with HIV and 3 (2%) had an Orthotopic Liver Transplant<br />

(OLT). Prior to initiation of therapy; mean MELD score was 9<br />

(SD 4.96, range 6-29) and 53 (37%) had a Fibroscan score of<br />

> 12.5 kpa. The majority (91,64%,) had genotype 1 infection<br />

and 75 (53%) were treatment naïve. Patients who had failed<br />

first generation protease inhibitors were included. Opiate substitution<br />

therapy was prescribed in 42 (30%) patients. To date<br />

101/109 (93%) planned patient episodes are complete and<br />

33 remain on treatment. Of the 8 patient episodes where treatment<br />

was incomplete, 2 patients receiving peg/simeprevir/<br />

ribavirin switched to sofosbuvir/ledipasvir due to intolerable<br />

IFN related side effects however completed the all oral regimen,<br />

2 receiving daclatasvir/sofosbuvir died from end stage<br />

liver disease and 2 developed non resectable/transplantable.<br />

Treatment was incomplete in 2 patients who received sofosbuvir/ledipasvir<br />

+ ribavirin, one due to suicidal ideation after 1<br />

week, the other developed renal failure after 3 weeks. The new<br />

treatment regimens are well tolerated; to date 93% of patients<br />

have completed their planned course; the majority who failed<br />

to complete had advanced liver disease. No-one has been lost<br />

to follow up. We would hope that SVR rates will be comparable<br />

to those reported in clinical trials.<br />

TABLE 1<br />

Disclosures:<br />

The following authors have nothing to disclose: Pauline Dundas, Shirley English,<br />

Lorna Bailey, Ashis Mukhopadhya, Balasubramaniam Vijayan, Andrew Fraser,<br />

Lindsay McLeman<br />

1105<br />

Asian Patients with Genotype 1 HCV Infection Achieve<br />

99% Sustained Virologic Response with 12 Weeks of<br />

Ledipasvir/Sofosbuvir Single Tablet Regimen: Integrated<br />

analysis of Phase 3 Multicenter Studies<br />

Masashi Mizokami 2 , Wan-Long Chuang 3 , Kwang-Hyub Han 4 ,<br />

Young-Suk Lim 5 , Jenny C. Yang 1 , Shampa De-Oertel 1 , Bing<br />

Gao 1 , Hongmei Mo 1 , Phillip S. Pang 1 , Steven J. Knox 1 , John G.<br />

McHutchison 1 , Masao Omata 6 , Jia-Horng Kao 7 ; 1 Gilead Sciences,<br />

Foster City, CA; 2 National Center for Global Health and Medicine,<br />

Tokyo, Japan; 3 Kaohsiung Medical University, Kaohsiung, Taiwan;<br />

4 Yonsei University College of Medicine, Seoul, Korea (the Republic<br />

of); 5 University of Ulsan College of Medicine, Asan Medical Center,<br />

Seoul, Korea (the Republic of); 6 Yamanashi Prefectural Hospital<br />

Organization, Yamanashi, Japan; 7 National Taiwan University<br />

College of Medicine, Taipei, Taiwan<br />

Introduction: Similar to the United States and Europe, the majority<br />

of patients with chronic HCV infection in Japan, Korea, and<br />

Taiwan are infected with HCV genotype (GT) 1. However,<br />

important differences in viral and host characteristics exist<br />

between the infected populations in these regions, including<br />

age, BMI, IL28B genotype and HCV GT1 subtype. The aim of<br />

this integrated analysis is to evaluate the efficacy and safety<br />

of LDV/SOF in a large cohort of Asian patients with chronic<br />

GT1 HCV infection. Methods: This analysis combines data<br />

from subjects enrolled in two Phase 3 trials: GS-US-337-0113<br />

(Japan) and GS-US-337-0131 (Korea and Taiwan) evaluating<br />

12 weeks of LDV/SOF (90mg/400mg) in treatment-naïve and<br />

treatment-experienced adults with chronic GT1 HCV infection.<br />

Eligibility criteria: no upper age limit restriction, inclusion of<br />

subjects with compensated cirrhosis, platelet count ≥50,000/<br />

mL and no restriction on neutrophil values. HCV NS5A and<br />

NS5B resistance associated variants (RAV) were evaluated by<br />

deep sequencing (cutoff of 1%). The primary efficacy endpoint<br />

was SVR12. Results: Overall, 349 subjects were enrolled in<br />

Korea, Taiwan, and Japan, 67 (19%) had cirrhosis. The majority<br />

were female (58%), treatment-experienced (51%), GT1b<br />

infected (94%), and IL28B CC (62%). The mean age (range)<br />

was 57 (18-80) years old, BMI 24 (17-38) kg/m 2 , and HCV<br />

RNA was 6.6 (3.7-7.6) log 10<br />

IU/mL. HCV NS5A RAVs were<br />

detected in 23% (80/343) of subjects at baseline. The overall<br />

SVR12 rate was 99% (346/349); 2 subjects relapsed and 1<br />

subject prematurely discontinued treatment. All treatment-experienced<br />

subjects with cirrhosis (45/45) achieved SVR12.<br />

NS5A RAVs were detected at the time of relapse but no NS5B<br />

RAVs were detected. Serious AE and treatment discontinuations<br />

were rare (


754A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Steven J. Knox - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Masao Omata - Advisory Committees or Review Panels: Boehringer Ingelheim;<br />

Speaking and Teaching: Otsuka Pharmaceutical, Bayer<br />

The following authors have nothing to disclose: Masashi Mizokami, Kwang-Hyub<br />

Han, Jia-Horng Kao<br />

1106<br />

Efficacy, Change in MELD Score, and Safety by Baseline<br />

MELD Score in Patients With Compensated Cirrhosis<br />

Receiving Ombitasvir/Paritaprevir/r and Dasabuvir Plus<br />

Ribavirin in the Phase 3 TURQUOISE-II Trial<br />

Ira M. Jacobson 1 , Tania M. Welzel 2 , Hugo E. Vargas 3 , Marcos<br />

C. Pedrosa 4 , Norah Terrault 5 , Douglas Dieterich 6 , Fredric D.<br />

Gordon 7 , Kris V. Kowdley 8 , Guy Neff 4 , Ran Liu 4 , Juan-Carlos<br />

Lopez-Talavera 4 , Stefan Zeuzem 2 ; 1 Weill Cornell Medical College,<br />

New York, NY; 2 J.W. Goethe University, Frankfurt, Germany;<br />

3 Mayo Clinic Arizona, Phoenix, AZ; 4 AbbVie Inc., North Chicago,<br />

IL; 5 University of California San Francisco, San Francisco,<br />

CA; 6 Mount Sinai School of Medicine, Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 7 Lahey Hospital & Medical Center,<br />

Burlington, MA; 8 Liver Care Network, Swedish Medical Center,<br />

Seattle, WA<br />

Purpose: Curing HCV infection reduces the risk of complications<br />

such as liver decompensation and hepatocellular carcinoma.<br />

The 3 direct-acting antiviral (3D) regimens of co-formulated<br />

ombitasvir/paritaprevir/r (paritaprevir identified by AbbVie<br />

and Enanta, dosed with ritonavir [r]) and dasabuvir +/- ribavirin<br />

(RBV) have achieved high SVR12 rates among adults with<br />

chronic genotype 1 HCV, including those with compensated<br />

cirrhosis. Model for end-stage liver disease (MELD) scores<br />

assess liver disease severity. In this analysis, we evaluated<br />

efficacy and safety of 3D+RBV and changes in MELD score by<br />

baseline MELD score in the phase 3 TURQUOISE-II trial. Methods:<br />

Treatment-naïve or peginterferon/RBV treatment-experienced<br />

patients with compensated cirrhosis were randomized<br />

to receive either 12 or 24 weeks of 3D+RBV in TURQUOISE-II.<br />

SVR12, change in MELD score, and adverse event rates are<br />

reported for subgroups of patients with baseline MELD scores<br />

of 6-9, 10-13, and >14. Results: 380 patients were randomized<br />

and received either 12 or 24 weeks of study drug. SVR12<br />

rates were 94.2% (322/342), 93.9% (31/33), and 80.0%<br />

(4/5) in patients with baseline MELD scores of 6-9, 10-13, and<br />

>14, respectively. The 1 patient with MELD >14 not achieving<br />

SVR12 discontinued study drug prematurely. Changes in MELD<br />

score in patients who achieved SVR12 is in the table. Rate of<br />

serious adverse events was 5.3% (18/342), 6.1% (2/33),<br />

and 0% (0/5) for the 6-9, 10-13, and >14 subgroups, respectively.<br />

Rate of discontinuation due to adverse events was 2.0%<br />

(7/342), 0% (0/33), and 20.0% (1/5) for the 6-9, 10-13, and<br />

>14 subgroups, respectively. Conclusions: In TURQUOISE-II,<br />

3D+RBV demonstrated high SVR12 rates and favorable safety<br />

in cirrhotic patients regardless of baseline MELD score. While<br />

this analysis is limited by the small number of patients, the data<br />

suggest improvements in MELD scores in patients with higher<br />

baseline MELD scores.<br />

Table. Changes in MELD score for patients who achieved SVR12.<br />

*PTW48 data were not available for all patients who achieved<br />

SVR12. Only data for those patients with data at PTW48 is<br />

shown. PTW, post-treatment week.<br />

Disclosures:<br />

Ira M. Jacobson - Consulting: AbbVie, Achillion, Alnylam, Bristol Myers Squibb,<br />

Enanta, Gilead, Janssen, Merck; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead, Janssen, Merck, Tobira; Speaking and Teaching: AbbVie, Bristol<br />

Myers Squibb, Gilead, Janssen<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Fredric D. Gordon - Advisory Committees or Review Panels: Gilead, AbbVie;<br />

Grant/Research Support: BMS, Vertex, Gilead, AbbVie<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Ran Liu - Employment: Abbvie<br />

Juan-Carlos Lopez-Talavera - Employment: Abbvie; Stock Shareholder: BMS<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

The following authors have nothing to disclose: Hugo E. Vargas, Marcos C.<br />

Pedrosa, Guy Neff<br />

1107<br />

Efficacy and safety of ombitasvir/paritaprevir/r and<br />

dasabuvir +/- ribavirin in HCV genotype 1-infected<br />

patients with a history of bleeding disorders: Results<br />

from phase 3 trials<br />

Giustino Parruti 1 , Guilherme Macedo 2 , Axel Baumgarten 3 ,<br />

Frederik Nevens 4 , Jordan J. Feld 5 , Christophe Hezode 6 , Lois<br />

M. Larsen 7 , Nancy Shulman 7 , Regis A. Vilchez 7 , Heiner Wedemeyer<br />

8 ; 1 Ospedale Civile Spirito Santo, Pescara, Italy; 2 Centro<br />

Hospitalar de São João, Porto, Portugal; 3 Medical Center for Infectious<br />

Diseases, Berlin, Germany; 4 University Hospitals KU, Leuven,<br />

Belgium; 5 Toronto Centre for Liver Disease, University of Toronto,<br />

Toronto, ON, Canada; 6 Henri Mondor University Hospital, AP-HP,<br />

Université Paris-Est, Créteil, France; 7 AbbVie Inc., North Chicago,<br />

IL; 8 Medizinische Hochschule Hannover, Hannover, Germany<br />

Background: HCV infection has been prevalent in patients with<br />

bleeding disorders due to the high frequency of contamination<br />

of blood-derived clotting factor products prior to the 1990s.<br />

The 3 direct-acting antiviral (3D) regimens of co-formulated<br />

ombitasvir/paritaprevir/ritonavir (OBV/PTV/r, PTV identified<br />

by AbbVie and Enanta) and dasabuvir (DSV) +/- ribavirin<br />

(RBV) have achieved high SVR12 rates among adults with<br />

chronic genotype 1 HCV including those with compensated<br />

cirrhosis. In this analysis, we examined the efficacy and safety<br />

of 3D+/-RBV in patients with a history of bleeding disorders in<br />

phase 3 trials. Methods: Treatment-naïve or peginterferon/RBV<br />

treatment-experienced, cirrhotic or non-cirrhotic HCV genotype<br />

1-infected patients enrolled in phase 3 trials and received<br />

3D+/-weight-based RBV. The efficacy (SVR12 rate) and safety<br />

for patients with a history of bleeding disorders is reported.<br />

Results: Thirty-five patients with a history of bleeding disorders<br />

(29 of whom had hemophilia) were enrolled and received<br />

active study drug (32 patients, 3D+RBV; 3 patients, 3D). Baseline<br />

characteristics and efficacy results are in the table. Fifteen


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 755A<br />

patients were peginterferon/RBV treatment-experienced and 9<br />

patients were cirrhotic. SVR12 rate was 100% (35/35, 95%<br />

CI: 90.1%, 100%) among patients with a history of bleeding<br />

disorders. Among the 35 patients, there were no discontinuations<br />

due to adverse events. The most common adverse events<br />

were nausea, fatigue, and headache. One patient (2.9%) had<br />

a decline in hemoglobin to


756A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1109<br />

Lower 25-OH Vitamin D Levels are Associated With<br />

Advanced Fibrosis in Chronic Hepatitis C Infection but<br />

Pretreatment or Change in Level with Directing Acting<br />

Antiviral Therapy Do Not Predict Sustained Virologic<br />

Response<br />

David W. Backstedt 1 , Mark Pedersen 2 , Bobby Kakati 1 , Myunghan<br />

Choi 4 , Anil B. Seetharam 3 ; 1 Gastroenterology, Banner University<br />

Medical Center, Phoenix, AZ; 2 Internal Medicine, Banner University<br />

Medical Center, Phoenix, AZ; 3 University of Arizona College<br />

of Medicine, Banner Transplant and Advanced Liver Disease Center,<br />

Phoenix, AZ; 4 Arizona State University College of Nursing and<br />

Health Innovation, Phoenix, AZ<br />

Rationale: Recent reports suggest association between low<br />

25-OH Vitamin D (Vit D) levels and advanced fibrosis in Hepatitis<br />

C Virus (HCV) infection. Evaluation of the influence and<br />

predictive value of pretreatment Vit D level on sustained virologic<br />

response (SVR) with direct acting antiviral (DAA) therapy<br />

is needed. Aim: Evaluate association of pre-treatment Vit<br />

D level with stage of fibrosis in an HCV infected cohort and<br />

assess change in level with DAA therapy. Secondary aim was<br />

to assess predictive value of pre-treatment or change in Vit D<br />

level on SVR12. Methods: Prospective study of chronic HCV<br />

patients initiated on DAA. Consecutively enrolled patients<br />

received: 1) pegylated interferon alfa 2a (IFN) + sofobuvir<br />

(SOF) + ribavirin (RBV); 2) SOF + RBV; 3) SOF + Simeprevir<br />

(SMV); and 4) Ledipasvir (LDV) + SOF; 5) LDV + SOF + RBV in<br />

accordance with AASLD guidelines. Vit D levels were assessed<br />

prior to initiation of treatment and at end of treatment response<br />

(ETR). Pearson correlation coefficients were assessed followed<br />

by repeated measures ANOVA to determine significant Vit D<br />

level changes over time controlling for covariates: age, gender,<br />

and BMI. Multivariate logistic regression was performed<br />

to identify predictors of SVR12. Results: 218 patients completing<br />

DAA regimen: IFN + SOF + RBV (n=37); SOF + RBV<br />

(n=117); SMV + SOF (n=51), LDV+SOF (n=11), LDV +SOF<br />

+RIBA (n=2). 61.5% genotype 1 and 56% cirrhosis. 64%<br />

treatment naïve and 36% treatment experienced (26% non-responders,<br />

10% relapsers). 98% (213/218) achieved ETR with<br />

79% (172/218) achieving SVR12 (19% relapsers). Frequency<br />

of low pre-treatment Vit D level (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 757A<br />

1111<br />

Sofosbuvir and Ledipasvir for the Treatment of HCV<br />

GT-1, Cirrhotics and Non-Cirrhotics: Real-World Effectiveness<br />

Kirat Gill, Garrett Fante, Shirin Nafisi, Ann Moore, Ashley Hepner,<br />

Justin A. Reynolds, Karen Arendt, Yvette Cummings, Robert Gish,<br />

Richard A. Manch, Anita Kohli; Hepatology, St. Joseph’s Hospital<br />

and Medical Center, Creighton University School of Medicine,<br />

Phoenix, AZ<br />

The interferon and ribavirin (RBV) free combination of sofosbuvir<br />

(SOF) and ledipasvir (LDV), a hepatitis C nucleotide NS5B<br />

and N5A inhibitor respectively, have demonstrated high rates<br />

of sustained virologic response (SVR) in clinical trials of both<br />

treatment naïve and experienced patients, as well as cirrhotic<br />

and non-cirrhotic patients, with chronic hepatitis C (CHC). Real<br />

world non-clinical trial data for patients with CHC treated with<br />

SOF/LDV ± RBV is lacking. AIM: To evaluate the effectiveness<br />

of SOF/LDV for treatment of HCV genotype 1 (GT1) in a realworld<br />

patient population. METHODS: We performed a retrospective<br />

analysis of patients with CHC GT1 treated between<br />

October 2014-June 2015 (n=197) at a community based clinic<br />

within an academic medical center in Phoenix, AZ or by ECHO<br />

telemedicine to rural clinic sites. Demographics, CHC treatment<br />

regimens, liver fibrosis, virologic data and psychiatric<br />

history were collected throughout treatment and post-treatment<br />

from patient charts. RESULTS: Since October 2014, 197 GT1<br />

patients started treatment for CHC and are included in the<br />

current analysis. 81.7% of patients were Caucasian, 11.2%<br />

of patients were Hispanic and 1.5% Native American. 65% of<br />

patients were men, the mean age was 59 yrs. (range: 18-82<br />

yrs.) and mean HCV viral load 2,359,054 IU/mL (range:<br />

1,107-18,285,438 IU/mL). 75% of patients were GT 1a and<br />

24% were GT1b. 53.3% of patients had cirrhosis by imaging,<br />

Fibrosure or liver biopsy. 29.4% of patients had a concomitant<br />

psychiatric illness. 45.2% (n=89) of patients were previously<br />

treated. 50% (n=100) of patients were treated with the SOF/<br />

LDV for 12 weeks, 31.0% (n=61) with SOF/LDV for 24 weeks,<br />

18.3% (n=36) with SOF/LDV for 8 weeks and 1.0% (n=2)<br />

treated with SOF/LDV/RBV for 12 weeks. Overall, 87.5% of<br />

patients treated with SOF/LDV for 12 weeks and who have<br />

reached the SVR12 timepoint (n= 16) have achieved SVR12,<br />

and 12.5% have relapsed (pending for 84 patients). SVR12<br />

data is pending for 60 patients treated with 24 weeks with<br />

SOF/LDV with one patient having relapsed. 100% of patients<br />

treated with SOF/LDV for 8 weeks and who have reached<br />

the SVR12 timepoint (n=10) have achieved SVR12 (pending<br />

for 26 patients). SVR12 is pending in the two patients treated<br />

with SOF/LDV/RBV. Full results of SVR 12 will be presented<br />

and stratified by cirrhosis and viral genotype. CONCLUSION:<br />

Treatment for hepatitis C GT1 using SOF/LDV based therapies<br />

has been well tolerated with high rates of SVR to date. SVR<br />

results in this real world population, enriched for patients with<br />

advanced liver disease, and treated by hepatologists as well as<br />

primary care providers, will be presented.<br />

Disclosures:<br />

Robert Gish - Advisory Committees or Review Panels: Gilead, AbbVie, Arrowhead;<br />

Consulting: Eiger, Isis, Genentech; Speaking and Teaching: Gilead, Abb-<br />

Vie; Stock Shareholder: Arrowhead<br />

Richard A. Manch - Speaking and Teaching: Gilead, AbbVie, Bayer, Salix, BMS<br />

The following authors have nothing to disclose: Kirat Gill, Garrett Fante, Shirin<br />

Nafisi, Ann Moore, Ashley Hepner, Justin A. Reynolds, Karen Arendt, Yvette<br />

Cummings, Anita Kohli<br />

1112<br />

Telaprevir in the Treatment of Acute HCV infection in<br />

HIV-infected Men: Final Results<br />

Daniel S. Fierer 2 , Douglas Dieterich 1 , Michael P. Mullen 2 , Andrea<br />

D. Branch 1 , Alison J. Uriel 1,6 , Damaris C. Carriero 1,7 , Wouter van<br />

Seggelen 2,3 , Rosanne M. Hijdra 2,3 , David Cassagnol 2,5 , Tristan<br />

Morey 2,4 ; 1 Liver Diseases, Icahn School of Medicine at Mount<br />

Sinai, New York, NY; 2 Infectious Diseases, Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 3 Amsterdam Medical Center,<br />

Amsterdam, Netherlands; 4 James Cook University, Cairns, QLD,<br />

Australia; 5 Weill Cornell Medical College, New York, NY; 6 North<br />

Manchester General Hospital, Manchester, United Kingdom;<br />

7 Columbia University School of Medicine, New York, NY<br />

Background Treatment of HCV with pegylated interferon (pIFN)<br />

and ribavirin (RBV) is significantly more effective in the acute<br />

phase than in the chronic phase. Incorporating telaprevir (TVR)<br />

into the treatment of chronic HCV both improves the efficacy<br />

and shortens the duration of treatment, and our preliminary evidence<br />

suggested it can do the same in acute HCV. We present<br />

here the final results of our study of TVR + pIFN + RBV for acute<br />

HCV in HIV-infected men. Methods This is an IRB-approved,<br />

open-label, consecutive enrollment study of TVR (TID or BID)<br />

+ pIFN-α 180 μg/week + weight-based RBV for 12 weeks in<br />

HIV-infected men with acute genotype 1 HCV infection. Stopping<br />

rule was HCV VL > 1,000 IU/mL at week 4. Allowed<br />

ARVs were tenofovir+emtricitabine, efavirenz (with TVR dose<br />

adjustment), rilpivirine, atazanavir/ritonavir, and raltegravir.<br />

HCV VL was measured by transcription-mediated amplification<br />

(TMA, lower limit of detection 5 IU/mL). The comparator<br />

group was HIV-infected men with acute genotype 1 HCV either<br />

treated during the 3 years prior to the FDA approval of TVR or<br />

those ineligible for TVR. Results Thirty-six men in the TVR group<br />

and 50 men in the comparator group completed treatment and<br />

SVR 12 assessment. The two treatment groups did not differ in<br />

proportion of genotype 1a (84% vs 90%, respectively), IL28B<br />

CC (45% vs 42%, respectively), age, or race/ethnicity. Thirty-two<br />

(89%) achieved SVR 12 in the TVR group compared to<br />

32 (64%) in the comparator group (p = 0.01). Most (88%) in<br />

the TVR group received ≤12 weeks total treatment, while all<br />

in the comparator group received ≥24 weeks total treatment.<br />

TVR treatment resulted in faster VL kinetics among those who<br />

achieved SVR, with median time to HCV VL 75%. With the availability now of interferon-free regimens,<br />

though, efforts should move toward developing these less toxic<br />

regimens. Clinicians who do not have access to interferon-free<br />

regimens, however, should be alert to detect acute HCV infection<br />

in HIV-infected men to take advantage of this highly-effective<br />

TVR-based therapy.<br />

Disclosures:<br />

Daniel S. Fierer - Stock Shareholder: Gilead<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Michael P. Mullen - Advisory Committees or Review Panels: GILEAD; Speaking<br />

and Teaching: GILEAD<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

Damaris C. Carriero - Advisory Committees or Review Panels: Abbvie, BMS;<br />

Consulting: Gilead; Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Alison J. Uriel, Wouter van Seggelen,<br />

Rosanne M. Hijdra, David Cassagnol, Tristan Morey


758A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1113<br />

Transient renal dysfunction during INF-free therapy in<br />

decompensated HCV cirrhosis patients<br />

Ivana Carey, Suman Verma, Anna Mrzljak, Aisling B. Considine,<br />

Sarah Knighton, Kath Oakes, Kate Childs, Matthew J. Bruce, Mary<br />

Horner, Abid Suddle, Kosh Agarwal; Institute of Liver Studies,<br />

Kings College School of Medicine at King’s College Hospital, London,<br />

United Kingdom<br />

Transient renal dysfunction occurs in 5-7% HCV patients<br />

treated with triple antiviral peg-IFN based therapy. Similarly<br />

Coilly at al reported transient decline in estimated glomerular<br />

filtration (eGFR) in cohort of post-transplant HCV patients<br />

treated with IFN-free (sofosbuvir (SOF) + daclatasvir (DCV) +/-<br />

RBV). On-treatment changes in specific renal markers cystatin<br />

C (glomerular function) and neutrophil gelatinase-associated<br />

lipocalin (NGAL) (tubular function), used in diagnosis of acute<br />

kidney injury, enhanced therapy delivery in patients on triple<br />

antiviral therapy. There are no data on cystatin C and NGAL<br />

levels changes during IFN-free therapy in HCV patients with<br />

decompensated cirrhosis (Child-Pugh B ≥7). Aim: Retrospective<br />

cross-sectional single centre study to investigate on-treatment<br />

changes in cystatin C and NGAL levels and correlate<br />

with changes in eGFR and presence/absence of proteinuria<br />

during IFN-free therapy for 12 weeks (SOF/ledipasvir+RBV or<br />

SOF+DCV+RBV) in HCV patients with decompensated cirrhosis.<br />

Patients: 52 patients with HCV decompesnated cirrhosis<br />

(median age 57years, 33 males, median MELD 11) underwent<br />

12 weeks of therapy with IFN-free DAA regimen (86% SOF/<br />

ledipasvir+ RBV and 14% SOF+DCV+RBV) and completed 12<br />

weeks post-therapy follow-up. Material and Methods: Plasma<br />

levels of cystatin C, NGAL and renin at baseline, treatment<br />

week 4 (TW4), end of therapy (EoT), follow-up week 4 (FUW4)<br />

and follow-up week 12 (FUW12) were measured by ELISA<br />

and compared with eGFR and urinary protein creatinine ratio<br />

(uPCR) at the same time-points. Patients were stratified according<br />

to pre-existing renal risks (RR) (hypertension, diabetes,<br />

diuretics, tacrolimus). The results are presented as medians.<br />

Results: 50% patients had renal risk; 16 hypertension, 10 diabetes,<br />

26 therapy with diuretics and 6 received tacrolimus.<br />

At baseline 14% of patients had eGFR15 and these proportions did not change significantly<br />

during therapy and follow up. eGFR did not change<br />

significantly during therapy and was lower in RR than non-RR<br />

patients (74 vs. 93min/ml, p=0.02). Cystatin C and NGAL<br />

levels were higher in RR patients and increased during the first<br />

4 weeks of therapy in all patients (TW0 vs. TW4- cystatin C:<br />

1.46 vs. 1.55mg/l, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 759A<br />

1115<br />

Resistance-associated variants in NS5A region of hepatitis<br />

C virus are susceptible to interferon-based therapy<br />

Jun Itakura, Masayuki Kurosaki, Natsuko Nakakuki, Hitomi<br />

Takada, Nobuharu Tamaki, Yutaka Yasui, Shoko Suzuki, Kaoru<br />

Tsuchiya, Hiroyuki Nakanishi, Namiki Izumi; Division of Gastroenterology<br />

and Hepatology, Musashino Red Cross Hospital, Tokyo,<br />

Japan<br />

BACKGROUND & AIMS: Treatments with direct acting antivirals<br />

(DAAs) has become the standard of chronic hepatitis C treatment<br />

in Japan. But the presence of resistance-associated variants<br />

(RAVs) of hepatitis C virus (HCV) attenuates the efficacy of<br />

DAAs. The aim of this study was to characterize the susceptibility<br />

of RAVs to interferon-based therapy. METHODS: Twenty nine<br />

genotype 1b patients with detectable Y93H variant in NS5A<br />

region at baseline were enrolled and treated by a combination<br />

of simeprevir (NS3/4A inhibitor), pegylated interferon and<br />

ribavirin. The longitudinal changes of RAV were determined<br />

by direct sequencing during therapy and at breakthrough or<br />

relapse, and the longitudinal changes in the proportion of<br />

Y93H RAV were determined by deep sequencing. RESULTS:<br />

By direct sequencing, Y93H RAV became undetectable or its<br />

proportion decreased in 57% of patients with both Y93H and<br />

Y93wild type at baseline at an early time point during therapy<br />

when HCV RNA was still detectable. By deep sequencing, the<br />

proportion of Y93H RAV was decreased from 52.7% (5.8% –<br />

97.4%) at baseline to 29.7% (0.16% - 98.3%) within 7 days of<br />

initiation of treatment (p=0.023). The proportion of Y93H RAV<br />

against Y93 wild type was reduced in 21 of 29 cases (72.4%)<br />

and a marked reduction of more than 10% was observed in<br />

14 cases (48.7%). In 4 cases with breakthrough/relapse, the<br />

proportion of Y93H RAV decreased during therapy and at<br />

breakthrough/relapse but recovered to baseline 3 months after<br />

discontinuation of treatment. HCV RNA reduction was significantly<br />

greater for Y93H RAV (-3.65±1.3 logIU/mL/day) than<br />

the Y93wild type (-3.35±1.0 logIU/mL/day) (p<br />

Shuhei Hige, Itaru Ozeki, Masakatsu Yamaguchi, Mutuumi<br />

Kimura, Tomohiro Arakawa, Tomoaki Nakajima, Yasuaki Kuwata,<br />

Takahiro Sato, Takumi Ohmura, Joji Toyota, Yoshiyasu Karino;<br />

Hepatology, Sapporo-Kosei General Hospital, Sapporo, Japan<br />

Background/Aims: One of the most serious problems in the<br />

treatment of hepatitis C is the aging of hepatitis C patients<br />

who had been left behind without being treated with interferon-based<br />

therapy. Combination therapy with daclatasvir (DCV)<br />

and asunaprevir (ASV) is the only option for Japanese patients<br />

with genotype 1b chronic hepatitis C at the moment. However,<br />

the efficacy and safety of DCV/ASV combination therapy has<br />

not been fully elucidated especially for those who are in their<br />

late 70’s and 80’s. The aim of this study was to evaluate the<br />

efficacy and safety of DCV/ASV for senior patients including<br />

pharmacological analysis. Patients and Methods: Two<br />

hundred and ninety-three genotype 1 patients were treated<br />

with fixed daily dose of DCV (60mg) and ASV (200mg) for<br />

24 weeks. The patients were divided into 3 groups by age:<br />

Group I (


760A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

mononuclear cells were collected at baseline, treatment week<br />

4 (TW4), end-of treatment (EoT) and follow-up week 12 in all<br />

patients and flowcytometry was performed after multicolour<br />

staining for NK cells subsets (CD3, CD16, CD56, NKp30 and<br />

NKG2D). The frequency of NK cells subsets was compared<br />

between responders (R) (n=5) vs. relapsers (REL) (n=6) at different<br />

therapy time-points. Plasma IP10, IL10 and IL12 levels were<br />

measured by ELISA at same time-points. Results are presented<br />

as medians. Results: Baseline HCV RNA levels were similar<br />

between R and REL. Overall NK cells frequency (CD3-CD56+)<br />

was higher in R than REL at baseline (5.79% vs. 2.38%,<br />

p=0.02). The frequency of NK cells increased during therapy,<br />

but proportions remained higher in R than REL (7.5% vs. 4.2%).<br />

CD56bright cells subset was higher in R than REL at baseline<br />

(14.5% vs. 7.5%, p=0.01), but increased only in REL at EoT<br />

to 14.8% and dropped to pre-treatment levels (7.6%) at FU.<br />

The frequency of NKG2D was similar between R and REL at all<br />

time-points. NKp30+ cells increased during therapy only in REL<br />

than R (11.2% vs. 1.1%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 761A<br />

ties and anemia did not affect SRV-12. A sub-group of patients<br />

over age 75, analyzed separately, showed SVR-12 of 100%,<br />

10% reported fatigue and 5.2% had diarrhea, with no hospitalizations<br />

during the treatment. Conclusion: Sofosbuvir based<br />

regimen is safe and effective in patients over the age of 65<br />

years including those with cirrhosis.<br />

Genotype (GT)<br />

Treatment naïve (TN)<br />

Treatment experienced (TE)<br />

Disclosures:<br />

The following authors have nothing to disclose: Beshoy T. Yanny, Amandeep K.<br />

Sahota<br />

1120<br />

Sofosbuvir and Ledipasvir/Sofosbuvir for the Treatment<br />

of Patients with Chronic Genotype 6 Hepatitis C Virus<br />

Infection: Integrated Analysis of Phase 2 and Phase 3<br />

Studies<br />

Edward J. Gane 1 , Wan-Long Chuang 2 , Tarek I. Hassanein 3 ,<br />

Kris V. Kowdley 4 , Eric Lawitz 5 , Bing Gao 6 , Hongmei Mo 6 , Robert<br />

H. Hyland 6 , Jenny C. Yang 6 , Shampa De-Oertel 6 , Diana M.<br />

Brainard 6 , Steven J. Knox 6 , John G. McHutchison 6 , Ching-Lung<br />

Lai 7 , Henry Lik-Yuen Chan 8 ; 1 New Zealand Liver Transplant Unit,<br />

Auckland, New Zealand; 2 Kaohsiung Medical University, Kaohsiung,<br />

Taiwan; 3 Southern California Liver Centers, Coronado, CA;<br />

4 Swedish Medical Center, Seattle, WA; 5 The Texas Liver Institute,<br />

San Antonio, TX; 6 Gilead Sciences, Inc, Foster City, CA; 7 Queen<br />

Mary Hospital, Hong Kong, Hong Kong; 8 The Chinese University<br />

of Hong Kong, Hong Kong, Hong Kong<br />

Introduction: Chronic HCV infection is endemic in South East<br />

Asia with HCV genotype 6 (GT6) accounting for 18-49% of<br />

those infected. In the US and Canada, nearly one-third of immigrants<br />

from Southeast Asia with HCV are infected with GT6.<br />

Few <strong>studies</strong> have examined the efficacy and safety of direct<br />

acting antiviral (DAA) regimens in GT6 infected patients. The<br />

aim of this integrated analysis was to characterize the efficacy<br />

and safety of sofosbuvir (SOF)-based regimens in patients with<br />

chronic GT6 HCV infection. Methods: GT6 infected subjects<br />

were identified in 5 <strong>studies</strong> (ATOMIC, NEUTRINO, GS-US-<br />

334-0115, ELECTRON2, GS-US-337-0131) and are included<br />

in this analysis. Treatment-naïve or treatment-experienced<br />

patients received SOF+RBV±Peg-IFNα or ledipasvir (LDV)/SOF<br />

for 12-24 weeks. The primary efficacy endpoint in all <strong>studies</strong><br />

was SVR12. Results: A total of 52 subjects with GT6 HCV infection<br />

were identified. The majority were treatment-naïve (94%),<br />

Asian (81%), male (58%), and had IL28B CC alleles (81%).<br />

The mean age was 50 years (range 26-76) and 10% had cirrhosis.<br />

GT6 subtypes included 6a, 6a/b, 6c-1, 6e, 6g, 6j, 6l,<br />

6m, 6o, 6p, 6q, and 6r. The table below presents SVR12 by<br />

regimen. One subject in ELECTRON2 withdrew consent after<br />

receiving 8 weeks of LDV/SOF and relapsed with the emergent<br />

NS5B RAV S282T. All remaining 51 patients achieved<br />

SVR12, including 100% (3/3) experienced and 100% (5/5)<br />

cirrhotics. Conclusions: SOF+RBV±Peg-IFNα and LDV/SOF<br />

regimens are well-tolerated and highly effective in patients with<br />

chronic GT6 HCV infection including those who are treatment<br />

experienced and have compensated cirrhosis. These regimens<br />

provide multiple therapeutic options for consideration when<br />

evaluating optimal therapy for individual patients with chronic<br />

GT6 HCV infection.<br />

SVR12 Rates in GT6 HCV Patients<br />

1ATOMIC (P7977-0724), NEUTRINO (GS-US-334-0110); 2GS-<br />

US-334-0115; 3ELECTRON2 (GS-US-337-0122), GS-US-337-<br />

0131<br />

Disclosures:<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Tarek I. Hassanein - Advisory Committees or Review Panels: AbbVie, Bristol-Myers<br />

Squibb; Grant/Research Support: AbbVie Pharmaceuticals, Obalon, Bristol-Myers<br />

Squibb, Eiasi Pharmaceuticals, Gilead Sciences, Janssen R&D, Idenix<br />

Pharmaceuticals, Ikaria Therapeutics, Merck Sharp & Dohme, NGM BioPharmaceuticals,<br />

Ocera Therapeutics, Salix Pharmaceuticals, Sundise, TaiGen Biotechnology,<br />

Takeda Pharmaceuticals, Vital Therapies, Tobria; Speaking and<br />

Teaching: Baxter, Bristol-Myers Squibb, Gilead, Salix, AbbVie<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Bing Gao - Employment: Gilead; Stock Shareholder: Gilead<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Robert H. Hyland - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Jenny C. Yang - Employment: Gilead Sciences, Inc<br />

Shampa De-Oertel - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Steven J. Knox - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Ching-Lung Lai - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead Sciences Inc; Consulting: Bristol-Myers Squibb, Gilead Sciences, Inc;<br />

Speaking and Teaching: Bristol-Myers Squibb, Gilead Sciences, Inc<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

1121<br />

Real Life Experience with Sofosbuvir and Ledipasvir<br />

Fixed Dose Regimen in the Multiethnic Cohort of Patients<br />

with Chronic Hepatitis C<br />

Marina Roytman 2,1 , Alister L. Tang 1 , Christina Wu 1 , Joy I.<br />

Piotrowski 1 , Leena K. Hong 2 , Ruby Trujillo 2 , Leslie Huddleston 2 ,<br />

Isabel Hung Wan 2 , Peter Poerzgen 2 , Sumodh Kalathil 2 , Naoky<br />

CS C. Tsai 2,1 ; 1 Medicine, University of Hawaii at Manoa, John A.<br />

Burns School of Medicine, Honolulu, HI; 2 Queen’s Liver Center,<br />

Honolulu, HI<br />

Purpose/Background: ION-1-3 trials have demonstrated high<br />

SVRs and favorable side effect profiles in the treatment of GT1<br />

HCV patients. However, reports of regimen experiences in<br />

the real life clinical setting are limited. This study investigates<br />

the safety and efficacy of Ledipasvir (LDV)- Sofosbuvir (SOF)<br />

regimen in the multiethnic patient cohort, including patients<br />

with decompensated cirrhosis, in a real life clinical setting.<br />

Methods: Retrospective review of 154 patients with genotype<br />

1 (GT1) chronic hepatitis C (HCV) infection treated with a


762A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

fixed dose Ledipasvir (LDV) and Sofosbuvir (SOF) (90/400mg)<br />

regimen November 2014- May 2015 at Hawaii’s single tertiary<br />

referral center. We collected data on demographics,<br />

adverse events, laboratory values, and SVR (sustained virological<br />

response). Statistical analysis was performed with R<br />

version 3.1.0. Results: Baseline characteristics: 42% cirrhotic<br />

(23.8% of those Child-Pugh Class B/C), 47.7% Caucasian,<br />

30.1% Asian, 9.2% Pacific Islander, 65.3% male, mean age<br />

60 ±10.5, mean BMI 27.1 ±5.8, 11% diabetic, 60.1% GT1a.<br />

Interim analysis data are presented. Viral load was negative<br />

in 100% of patients at the end of treatment (EOT). Main side<br />

effects: headache 14.4%, weight gain 7.8%, nausea 5.9%,<br />

and diarrhea 3.9%. Headache was more frequent in cirrhotics<br />

versus non-cirrhotics (19 vs 11%). In patients with baseline<br />

abnormal ALT (>30 U/L), 84/99 showed normal ALT levels<br />

(15 IU/cc) at 4 weeks post treatment.<br />

Conclusions: 12-week treatment was highly effective and well<br />

tolerated in a multiethnic cohort of patients including patients<br />

with advanced liver disease. Interim analysis indicates SVR<br />

rates comparable to ION trials despite higher proportion (42%)<br />

of cirrhotics in our cohort. Early improvement in albumin levels<br />

in cirrhotic patients may indicate improvement in synthetic function.<br />

Complete data on SVR4 and SVR12 will be available on<br />

all patients by October 2015.<br />

reduced toxicities, the new Direct Acting Antivirals (DAA) could<br />

be attractive therapeutic options for HCV treatment after KT.<br />

The approach of safe HCV treatment post KT, could allow the<br />

use of organs from HCV +ve donors, thus reducing organ discard<br />

rates the wait-list time for KT. Methods: 15 HCV +ve KT<br />

recipients were prospectively enrolled in the study at the time<br />

of initiation of DAA. We report the data from 8 recipients who<br />

had completed HCV treatment till now. Results: From the 8<br />

adult (Mean age 58.5 years; Male: Female=6:2) KT recipients,<br />

4 received HCV +ve grafts. 2 completed Sofosbuvir 400<br />

mg and Simeprevir 150 mg daily combination and 6 completed<br />

Sofosbuvir/Ledipasvir 150mg/90 mg daily 12 weeks<br />

regimen. All achieved End Treatment Response. Data is being<br />

collected for Sustained Virologic Responses (SVR). There were<br />

no episodes of graft rejection and none required modification<br />

in immunosuppression. There were no adverse events requiring<br />

cessation of therapy. All KT recipients had stable renal and<br />

liver function during and after the completion of therapy. Conclusion:<br />

In our cohort of KT recipients, DAAs showed excellent<br />

safety profile and good virologic response.<br />

Study population characteristics and treatment responses of each<br />

subject.<br />

(KT) Kidney Transplant<br />

(IFN) Interferon<br />

(R) Ribavirin<br />

(PI) Protease Inhibitor<br />

(T) Tacrolimus<br />

(M) Mycophenolate<br />

(P) Prednisone<br />

Disclosures:<br />

Marina Roytman - Advisory Committees or Review Panels: Gilead; Speaking and<br />

Teaching: Gilead<br />

Leena K. Hong - Advisory Committees or Review Panels: Gilead Sciences; Speaking<br />

and Teaching: Gilead Sciences, Abbvie<br />

Naoky CS C. Tsai - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, AbbVie; Consulting: BMS, Gilead; Grant/Research Support: BMS,<br />

Gilead, AbbVie; Speaking and Teaching: Bristol-Myers Squibb, Gilead, Salix,<br />

Bayer, AbbVie<br />

The following authors have nothing to disclose: Alister L. Tang, Christina Wu, Joy<br />

I. Piotrowski, Ruby Trujillo, Leslie Huddleston, Isabel Hung Wan, Peter Poerzgen,<br />

Sumodh Kalathil<br />

(Sof) Sofosbuvir<br />

(Sim) Simeprevir<br />

(Led) Ledipasvir<br />

(EOT) End of Treatment<br />

(SVR) Sustained Virologic Response<br />

1122<br />

Treatment of Chronic Hepatitis C Infection in Kidney<br />

Transplant Recipients with Direct Acting Antiviral Medications<br />

– Initial Experience<br />

Sushrut Trakroo 1 , Sirish Sanaka 1 , Hawah Musa 1 , Mohammed<br />

Eyad Yaseen Alsabbagh 2 , Mythili Ghanta 1 , Swati Rao 1 , Lee Peng 2 ,<br />

Antonio Di Carlo 1 , Abdullah M. Al-Osaimi 2 , Kamran Qureshi 2 ;<br />

1 Temple University Hospital, Philadelphia, PA; 2 Gastroenterology<br />

and Hepatology, Temple University School of Medicine, Philadelphia,<br />

PA<br />

Background: Chronic Hepatitis C infection (HCV) is the leading<br />

cause of liver disease after Kidney Transplant (KT). HCV positive<br />

(+ve) KT recipients have worse overall and graft survival<br />

compared with HCV negative KT recipients. Treatment of HCV<br />

before KT with Interferon (IFN) based therapies showed poor<br />

response rates and tolerability. IFN use is relatively contraindicated<br />

after KT due to increased risk of allograft rejection<br />

and organ failure. Because of the shown greater efficacy and<br />

Disclosures:<br />

Abdullah M. Al-Osaimi - Advisory Committees or Review Panels: Abbvie, Gilead,<br />

Intercept Pharmaceuticals; Grant/Research Support: Chronic Liver Disease<br />

Foundation (CLDF), Beckman Coulter Inc., Conatus Pharmaceuticals, BioPharma<br />

Alliance, Bayer HealthCare Pharmaceuticals, Inc., Vital Therapies Inc., Roche<br />

Molecular Systems, Inc., Abbvie


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 763A<br />

The following authors have nothing to disclose: Sushrut Trakroo, Sirish Sanaka,<br />

Hawah Musa, Mohammed Eyad Yaseen Alsabbagh, Mythili Ghanta, Swati Rao,<br />

Lee Peng, Antonio Di Carlo, Kamran Qureshi<br />

1123<br />

Retreatment with sofosbuvir and simeprevir of patients<br />

who previously failed on an HCV NS5A inhibitor–containing<br />

regimen<br />

Christophe Hezode 1,2 , Stephane Chevaliez 3,2 , Giovanna<br />

Scoazec 1 , Alexandre Soulier 3,2 , Anne Varaut 1 , Magali Bouvier-Alias<br />

3,2 , Isaac Ruiz 3,2 , Ariane Mallat 1,2 , Cyrille Feray 1,2 ,<br />

Jean-Michel Pawlotsky 3,2 ; 1 Hepatologie, Hospital Henri Mondor,<br />

Creteil, France; 2 INSERM U955, Creteil, France; 3 Virology, Henri<br />

Mondor Hospital, Creteil, France<br />

Background: Failure to achieve SVR with IFN-free regimens<br />

based on HCV DAAs is associated with the selection of resistant<br />

HCV variants. In contrast to variants resistant to PIs that<br />

are progressively overgrown by wild-type virus after treatment<br />

failure, NS5A RAVs persist for years as the dominant species<br />

in the majority of patients. In spite of the lack of data, recent<br />

guidelines recommended that patients exposed to NS5A inhibitors<br />

be retreated with a combination of SOF and SMV to avoid<br />

cross-resistance with NS5A inhibitors. Materials & Methods: 16<br />

patients (mean age: 54 yrs) infected with GT 1 (1a: n=11; 1b:<br />

n=3) or 4 (n=2) with compensated disease and advanced liver<br />

fibrosis (fibroscan: 9.6-70 kPa; cirrhosis n=9) who failed to<br />

achieve SVR after receiving DCV/PegIFN/RBV (n=13) DCV/<br />

ASV/PegIFN/RBV (n=3), were included. They were retreated<br />

for 12 weeks with SOF/SMV, without RBV. HCV RNA levels<br />

were measured with Abbott RealTime Assay (LLOQ/LOD: 12<br />

IU/mL). HCV resistance was assessed at retreatment baseline<br />

by means of population sequencing home made methods targeting<br />

the NS3 and NS5A coding regions. Results: The median<br />

baseline HCV RNA level at retreatment was 1.38 x 10 6 IU/mL<br />

(>800,000 IU/mL: n=14). No patient discontinued treatment<br />

due to an adverse event or for virologic failure. Among the 16<br />

patients, 10 had a rapid virologic response (HCV RNA


764A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Ju Dong Yang, Bashar A. Aqel,<br />

Kymberly D. Watt, Hugo E. Vargas, Andrew P. Keaveny, Michael D. Leise<br />

1125<br />

Combination therapy with daclatasvir and asunaprevir<br />

in cirrhotic patients with hepatitis C virus genotype 1b:<br />

On-treatment efficacy and adverse effects<br />

Akihiro Tamori, Masaru Enomoto, Hoang Hai, Yuga Teranishi,<br />

Hiroyuki Motoyama, Ritsuzo Kozuka, Etsushi Kawamura, Atsushi<br />

Hagihara, Sawako K. Uchida, Hiroyasu Morikawa, Yoshiki<br />

Murakami, Norifumi Kawada; Hepatology, Osaka City University<br />

Graduate School of Medicine, Osaka, Japan<br />

Background and Aim: Combination therapy with daclatasvir<br />

(DCV; NS5A inhibitor) and asunaprevir (ASV; second-generation<br />

HCV-NS3/NS4A protease inhibitor) was approved for<br />

patients with HCV genotype 1 in Japan since September 2014.<br />

Now, elderly patients and those with advanced hepatic fibrosis<br />

including chronic liver cancer were administered IFN-free therapy.<br />

Our objective was to assess the efficacy and tolerability<br />

of DCV/ASV combination therapy in patients with hepatic cirrhosis.<br />

Methods: In total, 140 consecutive patients with HCV<br />

1b initiating DCV/ASV therapy were enrolled. The cohort comprised<br />

49 patients with compensated cirrhosis and 91 patients<br />

without cirrhosis (62 males and 78 females; median age, 71<br />

years; 6 patients were >80 years old). NS5A resistance-associated<br />

variants (RAV) were examined using direct sequencing.<br />

The patients were treated with 60 mg of DCV once daily and<br />

100 mg of ASV twice per day for 24 weeks. Clinical, biological,<br />

and virological data, including adverse effects, were<br />

recorded at baseline and during follow-up. Results: Only 10<br />

(7%) patients had L31M or Y93H RAVs. There was no statistically<br />

significant difference in age, sex, IL28B genotypes, HCV<br />

viral load at baseline, ALT level, creatinine level, or NS5A<br />

RAVs between patients with and without cirrhosis. On the other<br />

hand, those with cirrhosis showed significantly lower levels of<br />

platelets, white blood cells, and hemoglobin and higher levels<br />

of alpha fetoprotein. The rapid viral response rate (HCV-RNA <<br />

25 IU/ml at week 4) was the same between patients with and<br />

without cirrhosis (80% and 81%, respectively). Now, DCV/<br />

ASV therapy was completed in 67 patients. One of 27 patients<br />

with cirrhosis, and two of 40 patients without cirrhosis who<br />

did not have NS5A RAVs at baseline developed viral breakthrough.<br />

The rate of SVR4 was 94% (16/17) in patients with<br />

cirrhosis and 90% (19/21) in patients without cirrhosis. Grade<br />

3/4 complications frequently occurred in 22% of patients with<br />

cirrhosis, of whom one had an elevated ALT level, two progressed<br />

to decompensated cirrhosis without ALT elevation,<br />

one developed HCC, one developed interstitial pneumonia,<br />

one had severe bronchitis, one had arterial fibrillation, three<br />

had gastrointestinal bleeding, and two developed edema. Of<br />

the patients without cirrhosis, ALT elevation was observed in<br />

two patients, and HCC developed in one patient. Conclusion:<br />

DCV/ASV therapy achieved a high anti-HCV effect in patients<br />

both with and without cirrhosis. However, careful management<br />

is necessary in patients with cirrhosis.<br />

Disclosures:<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

The following authors have nothing to disclose: Akihiro Tamori, Masaru Enomoto,<br />

Hoang Hai, Yuga Teranishi, Hiroyuki Motoyama, Ritsuzo Kozuka, Etsushi<br />

Kawamura, Atsushi Hagihara, Sawako K. Uchida, Hiroyasu Morikawa, Yoshiki<br />

Murakami<br />

1126<br />

Safety and Efficacy of All-Oral Regimens for Hepatitis C<br />

in Cirrhotic Patients with Thalassemia<br />

Emmanouil Sinakos 1 , Dimitrios Kountouras 2,3 , Ioannis Koskinas 4 ,<br />

Stylianos Karatapanis 5 , Alexandra Kourakli 6 , Efthimia Vlachaki 7 ,<br />

Ioannis Goulis 1 , Antonios Kattamis 8 , Konstantinos Maragkos 9 ,<br />

Fotini Petropoulou 10 , Barbara Toli 3 , Evangelos Akriviadis 1 , George<br />

V. Papatheodoridis 2 ; 1 4th Medical Department, Hippokratio Hospital,<br />

Aristotle University of Thessaloniki, Thessaloniki, Greece;<br />

2 Department of Gastroenterology, Laiko Hospital, University of Athens,<br />

Athens, Greece; 3 Mitera Hospital, Athens, Athens, Greece;<br />

4 2nd Medical Department, Hippokratio Hospital, University of Athens,<br />

Athens, Greece; 5 Department of Internal Medicine, Hospital<br />

of Rhodes, Rhodes, Greece; 6 Thalassemia Unit, Hospital of Patra,<br />

Patra, Greece; 7 Thalassemia Unit, Aristotle University of Thessaloniki,<br />

Thessaloniki, Greece; 8 Agia Sofia Children Hospital, University<br />

of Athens, Athens, Greece; 9 Thalassemia Unit, Hippokratio Hospital,<br />

Athens, Athens, Greece; 10 Thalassemia Unit, General Hospital,<br />

Athens, Athens, Greece<br />

Background-Aims: Thalassemia syndromes have been traditionally<br />

associated with high hepatitis C prevalence rates due<br />

to the regular blood transfusion requirements. Treatment with<br />

interferon and ribavirin usually leads to anemia, poor tolerance<br />

and subsequently low clearance rates, thus often failing to halt<br />

progression to cirrhosis. Although current guidelines suggest<br />

that interferon-free and ribavirin-free regimens should be used<br />

in these patients, no data have been yet presented. The aim<br />

of this study was to assess the safety and efficacy of all-oral<br />

regimens against hepatitis C virus in cirrhotic patients with thalassemia.<br />

Patients-Methods: Data on patients with thalassemia<br />

treated for hepatitis C in everyday clinical practice were analyzed.<br />

All patients were exposed to regular blood transfusions<br />

and chelation therapy. The predominant chelation therapy was<br />

combination of deferoxamine and deferiprone. The stage of<br />

liver disease was assessed histologically or with transient elastography.<br />

Three different regimens were used: sofosbuvir+ribavirin:<br />

n=3 (10%) for 24 weeks, sofosbuvir+simeprevir: n=13<br />

(43%) for 12 weeks and sofosbuvir+daclatasvir: n=14 (47%)<br />

for 12 weeks. Monitoring for side effects was performed every<br />

month during treatment and as needed after treatment. Results:<br />

30 patients with mean age of 45±5 years were included. All<br />

patients had compensated cirrhosis, except for one who had<br />

F3 stage. 25 patients (83%) were non-responders to previous<br />

courses (up to 5), either of (peg-)interferon and ribavirin (n=<br />

17), or interferon monotherapy (n=8). The genotype distribution<br />

was as follows: G1a: n=5 (17%), G1b: n=12 (40%), G2:<br />

n=1 (3%), G3: n=6 (20%), G4: n=6 (20%). Baseline laboratory<br />

values were: Hb=10.2±0.9 g/dL, serum ferritin=593±364<br />

ng/mL, AST=98±51 IU/mL, ALT=104±47 IU/mL, serum albumin=3.7±0.5<br />

g/dL, total bilirubin=2.2±1.5 mg/dL, HCV<br />

RNA=1.5±1.8x10 6 IU/ml. The mean liver magnetic resonance<br />

imaging Τ2* was 19.6±11.7 msec (data on 14 patients). At<br />

the time of this report 10 patients have completed treatment.<br />

No early discontinuation, increase in blood transfusion requirements<br />

or cardiac arrhythmia was reported. 7/8 (87%) patients<br />

evaluated have reached SVR12. Two patients with SVR12 had<br />

received sofosbuvir+ribavirin (one with G1b and one with G2)<br />

and five had received sofosbuvir+simeprevir (two with G1a<br />

and three with G1b). The only failure was observed in a patient<br />

with G1b who received sofosbuvir+ribavirin for 24 weeks.<br />

Conclusions: Treatment of hepatitis C with all oral regimens in<br />

cirrhotic patients with thalassemia is safe. Preliminary efficacy<br />

results are promising and seem not to differ from published SVR<br />

rates in other populations.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 765A<br />

Disclosures:<br />

Emmanouil Sinakos - Advisory Committees or Review Panels: AbbVie; Speaking<br />

and Teaching: AbbVie, Bristol-Myers Squibb, Gilead, Janssen<br />

Ioannis Goulis - Consulting: Gilead Sciences, Abbvie, Janssen-Cilag, Janssen-Cilag,<br />

BMS; Grant/Research Support: BMS; Speaking and Teaching: BMS, Gilead<br />

Sciences, Janssen-Cilag<br />

George V. Papatheodoridis - Advisory Committees or Review Panels: Merck<br />

Sharp & Dohme, Novartis, Abbvie, Boerhinger Ingelheim, Bristol-Meyer Squibb,<br />

Gilead, Roche, Janssen, GlaxoSmith Kleine; Grant/Research Support: Roche,<br />

Gilead, Bristol-Meyer Squibb, Abbvie, Janssen; Speaking and Teaching: Merck<br />

Sharp & Dohme, Bristol-Meyer Squibb, Gilead, Roche, Janssen, Abbvie<br />

The following authors have nothing to disclose: Dimitrios Kountouras, Ioannis<br />

Koskinas, Stylianos Karatapanis, Alexandra Kourakli, Efthimia Vlachaki, Antonios<br />

Kattamis, Konstantinos Maragkos, Fotini Petropoulou, Barbara Toli, Evangelos<br />

Akriviadis<br />

1127<br />

NS3 Q80K Polymorphism in Viral Isolates from Liver<br />

Tissues and Serum Samples of Naïve HCV Genotype 1a<br />

Patients.<br />

Deborah D’Aliberti 2 , Irene Cacciola 2 , Salvatore Benfatto 1 , Federica<br />

Mannino 1 , Roberto Filomia 2 , Concetta Beninati 1 , Giovanni<br />

Raimondo 2 , Teresa Pollicino 1 ; 1 Pediatric, Gynecological, Microbiological<br />

and Biomedical Sciences, University Hospital of Messina,<br />

Messina, Italy; 2 Clinical and Experimental Medicine, University<br />

Hospital of Messina, Messina, Italy<br />

Background and Aim – NS3 Q80K is a polymorphism commonly<br />

detected in genotype (G) 1a HCV strains. It is associated<br />

with a reduced antiviral activity of HCV proteinase inhibitors<br />

(PIs) in vitro and its emergence is generally considered a cause<br />

of response failure to Simeprevir as well as to other PIs in<br />

HCV-G1a patients. Several <strong>studies</strong> have evaluated the prevalence<br />

of Q80K polymorphism in HCV isolates from treatment<br />

naïve patients of different geographic areas. However, the<br />

data available concern only circulating viral isolates, whereas<br />

there is no information about the variability of HCV strains in<br />

the liver where the viral replication occurs and the HCV quasispecies<br />

diversity reaches the highest level of complexity. Aim of<br />

our study was to investigate the presence of Q80K substitution<br />

in HCV-G1a isolates from liver tissues and serum samples of<br />

patients from south of Italy, naïve to any antiviral treatment.<br />

Patients and Methods – HCV NS3 protease sequences of HCV<br />

isolates from paired serum and liver biopsy samples of 11 naïve<br />

patients with HCV-G1a chronic infection, and from serum samples<br />

of additional 20 naïve HCV-G1a infected patients were<br />

analysed by population sequencing and ultra-deep pyrosequencing<br />

(UDPS) (mutant detection sensitivity 1%; >3000<br />

sequences/patient). Results – The Q80K substitution was found<br />

in 8/11 liver tissues (72.7%; in 7 present at levels between 1%<br />

and 10%, and in 1 at level ≈97% of the virus quasispecies) and<br />

in 5/11 corresponding serum samples (45.4%; in 4 present at<br />

level ≈1%, and in 1 at level ≈97% of the virus quasispecies) by<br />

UDPS. Of the additional 20 sera examined, 13 (65%) showed<br />

the Q80K substitution by UDPS (in 11 present at level ≈1%,<br />

in 1 present at level ≈20%, and in 1 present at level ≈99% of<br />

the virus quasispecies). On the whole, the Q80K substitution<br />

was detected in 18/31 sera (58%) analysed. By population<br />

sequencing, the Q80K substitution was detected only in samples<br />

from 2 of the 31 patients (6.4%) analyzed. In particular,<br />

the Q80K substitution was found in the paired liver and serum<br />

samples from 1 patient and in the serum sample from another<br />

patient with no liver specimen available (in all these 3 samples<br />

the Q80K was present in 97-99% of the entire viral populations<br />

by UDPS). Conclusions – HCV-G1a strains carrying<br />

Q80K polymorphism are commonly present both in liver and<br />

in serum of infected patients. However, they are present as<br />

minor populations in almost all the cases, suggesting that the<br />

Q80K substitution does not confer fitness advantage at least to<br />

the HCV-G1a viral strains circulating in our geographic area.<br />

Disclosures:<br />

Giovanni Raimondo - Speaking and Teaching: BMS, Gilead, Roche, Merck,<br />

Janssen, Bayer, MSD<br />

Teresa Pollicino - Speaking and Teaching: Gilead, Roche, Janssen, BMS, MSD,<br />

Bayer, Abbvie<br />

The following authors have nothing to disclose: Deborah D’Aliberti, Irene Cacciola,<br />

Salvatore Benfatto, Federica Mannino, Roberto Filomia, Concetta Beninati<br />

1128<br />

Safety and Efficacy of Treatment with Daily Sofosbuvir<br />

400 mg + Ribavirin 200 mg for 24 Weeks in Genotype<br />

1 or 3 HCV-Infected Patients with Severe Renal Impairment<br />

Paul Martin 2 , Edward J. Gane 3 , Grisell Ortiz-Lasanta 4 , Lin Liu 1 ,<br />

Karim Sajwani 1 , Brian Kirby 1 , Jill M. Denning 1 , Luisa M. Stamm 1 ,<br />

Diana M. Brainard 1 , John G. McHutchison 1 , Eric Lawitz 5 , Stuart<br />

C. Gordon 6 , Richard A. Robson 7 ; 1 Gilead Sciences, Raleigh, NC;<br />

2 University of Miami, FL, USA, Miami, FL; 3 University of Auckland,<br />

Auckland, New Zealand; 4 Fundcion de Investigacion, San Juan,<br />

PR; 5 Texas Liver Institute, San Antonio, TX; 6 Henry Ford Health<br />

System, Detroit, MI; 7 Christchurch Clinical Studies Trust, Ltd, Christchurch,<br />

New Zealand<br />

Background: HCV-infected patients with severe renal impairment<br />

have had limited treatment options with suboptimal<br />

response rates and poor tolerability of IFN-based regimens.<br />

SOF 200 mg+RBV studied previously was safe and relatively<br />

well tolerated in patients with severe renal impairment, but SVR<br />

rates were low. The current study evaluated the safety, pharmacokinetics<br />

(PK) and efficacy of SOF 400mg+RBV for 24 weeks<br />

in genotype 1 or 3 HCV-infected patients with severe renal<br />

impairment. Methods: Treatment-naïve or experienced patients<br />

± cirrhosis and a creatinine clearance (CLcr) less than 30mL/<br />

min (by Cockcroft-Gault equation), not on dialysis, received<br />

open-label treatment with SOF 400mg+RBV 200mg once daily<br />

for 24 wks. We examined on-treatment virologic response, PK<br />

and safety including echocardiograms read by a central reader<br />

pre-treatment and at Week 12 of therapy. Results: 10 patients (6<br />

GT1a, 2 GT1b, 2 GT3a) were enrolled and have been treated<br />

for 12-24 wks (median 16 wks): 8 male, 5 white, 4 black, 1<br />

Pacific Islander, mean age 58; 4 with cirrhosis, 7 treatment-experienced,<br />

5 IL28B genotype non-CC, mean baseline CLcr<br />

26.2 mL/min, and mean baseline hemoglobin (Hb)11.3 g/dL.<br />

All patients had a rapid virologic decline similar to those with<br />

normal renal function. Exposure of GS-331007, the primary<br />

circulating SOF metabolite which is renally eliminated, was<br />

~6-fold higher and SOF was ~1.4 fold higher than observed<br />

in subjects in the Phase 2/3 population. Adverse events (AEs)<br />

reported in more than one patient were anemia (n=3) and<br />

dizziness (n=2). Two patients had treatment-emergent SAEs,<br />

1 patient discontinued treatment just before week 12 due to<br />

an SAE of acute respiratory failure, and 1 patient experienced<br />

hematemesis. No patients had RBV dose-reduction, interruption<br />

or discontinuation with the exception of the one patient who<br />

discontinued study treatment due to an SAE. Three patients<br />

on epoetin at baseline continued throughout treatment. No<br />

patients started epoetin or had a blood transfusion during the<br />

treatment period. Changes in renal function, hemoglobin, and<br />

cardiac parameters are shown in Table 1. Conclusions: SOF<br />

400mg daily + RBV 200mg in GT1 or 3 HCV-infected patients<br />

with severe renal impairment was well-tolerated and resulted in<br />

rapid virologic suppression through Wk 12 of treatment. Final<br />

safety, PK and efficacy (SVR12) will be presented.


766A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Table 1. Mean changes in clinical parameters through Week 12<br />

on treatment<br />

Disclosures:<br />

Paul Martin - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Merck, Gilead, Janssen, Abbvie<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Lin Liu - Employment: Gilead Sciences, Inc.<br />

Karim Sajwani - Employment: Gilead Sciences, Inc.<br />

Brian Kirby - Employment: Gilead Sciences<br />

Jill M. Denning - Employment: Gilead Sciences, Inc.<br />

Luisa M. Stamm - Employment: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

The following authors have nothing to disclose: Grisell Ortiz-Lasanta, Richard<br />

A. Robson<br />

1129<br />

A Meta-Analysis of the Association Between Pre-Treatment<br />

Variables and SVR Amongst Genotype 1 Patients<br />

Treated with ABT-450/r–Ombitasvir and Dasabuvir<br />

(VIEKIRA PAK) plus Ribavirin for 12 Weeks<br />

Gregory J. Botwin 2 , Timothy R. Morgan 1,3 ; 1 Gastroenterology (11),<br />

VA Long Beach Healthcare System, Long Beach, CA; 2 Research<br />

Service, VA Long Beach Healthcare System, Long Beach, CA;<br />

3 Division of Gastroenterology, University of California - Irvine,<br />

Irvine, CA<br />

Background: VIEKIRA PAK plus Ribavirin (RBV) is a potent, all<br />

oral, treatment regimen indicated for the treatment of chronic<br />

hepatitis C (HCV) genotype (GT)-1. Due to the overall high<br />

rates of sustained virologic response (SVR) achieved in Phase<br />

II and Phase III clinical trials, little is known about the association<br />

between pre-treatment variables and SVR. Purpose: To<br />

perform a meta-analysis to examine the association between<br />

pre-treatment variables and SVR in GT-1 patients. Methods:<br />

We reviewed all registration trials for VIEKIRA PAK. Trials<br />

that treated GT-1a, or GT-1a and GT-1b, monoinfected, non-cirrhotic,<br />

HCV patients for 12 weeks with VIEKIRA PAK plus<br />

RBV were included in our initial analysis. Trials that enrolled<br />

exclusively GT-1b patients were excluded due to the overall<br />

high SVR rates. Relative risk (RR) was calculated for the following<br />

pre-treatment variables: age, body-mass index (BMI), ethnicity,<br />

IL28B, race, sex, and viral load (VL). A binary random<br />

effects model was used in the meta-analysis. Results: Three<br />

<strong>studies</strong> (SAPPHIRE-I and II, and PEARL-IV) and 870 patients met<br />

the criteria. Of the evaluated patients 595 were genotype 1a,<br />

and 297 were treatment experienced. Compared with patients<br />

with a BMI < 30, a BMI ≥ 30 was associated with a reduced<br />

RR of SVR, 0.948 (95% confidence-interval [CI], 0.903-0.995,<br />

p=0.029). The overall SVR in each group was 92% (139/151;<br />

BMI ≥ 30) and 97% (699/719; BMI < 30). Patients with a VL<br />

≥ 800,000IU/mL (SVR 96%, 686/716) also had a reduced<br />

RR of SVR, 0.973 (CI 0.948-0.999, p=0.041) compared to<br />

patients with a VL below this level (SVR 99%, 152/154). Neither<br />

age < 55 years, Hispanic ethnicity, IL28B CC genotype,<br />

African American race, nor male sex were found to be significantly<br />

positively or negatively associated with SVR. When<br />

380 GT-1 cirrhotic patients, treated for 12 or 24 weeks (TUR-<br />

QUOISE-II) were included in the analysis, only BMI remained<br />

significantly associated with SVR (RR 0.956, CI 0.921-0.993,<br />

p=0.021). Conclusion: Genotype 1 non-cirrhotic patients with<br />

an elevated BMI or VL are associated with a reduced chance<br />

of achieving SVR when treated with VIEKIRA PAK plus RBV<br />

for 12 weeks. Research evaluating the mechanism for these<br />

differences may be warranted.<br />

Pre-Treatment Variables Association with SVR<br />

Disclosures:<br />

Timothy R. Morgan - Grant/Research Support: Merck, Abbvie, Genentech, Gilead,<br />

Bristol Myers Squibb<br />

The following authors have nothing to disclose: Gregory J. Botwin<br />

1130<br />

On-treatment HCV RNA Decline in Pre-and Post-Liver<br />

Transplant Patients with Different Degrees of Fibrosis<br />

and Cirrhosis: a combined analysis of the SOLAR trials<br />

Tania M. Welzel 1 , K. Rajender Reddy 2 , Michael P. Manns 3 , Steven<br />

L. Flamm 4 , David J. Mutimer 5 , Edward J. Gane 6 , Robert H.<br />

Hyland 7 , Ming Lin 7 , Phillip S. Pang 7 , John G. McHutchison 7 , Didier<br />

Samuel 8 , Michael R. Charlton 9 , Xavier Forns 10 , Gregory T. Everson<br />

11 , Stefan Zeuzem 1 , Nezam H. Afdhal 12 ; 1 Johann Wolfgang<br />

Goethe University Medical Center, Frankfurt am Main, Germany;<br />

2 University of Pennsylvania School of Medicine, Philadelphia, PA;<br />

3 Hannover Medical School, Hannover, Germany; 4 Northwestern<br />

University Feinberg School of Medicine, Chicago, IL; 5 Queen Elizabeth<br />

Hospital and University of Birmingham, Birmingham, United<br />

Kingdom; 6 University of Auckland, Auckland, New Zealand; 7 Gilead<br />

Sciences, Inc, Foster City, CA; 8 Université Paris-Sud, Villejuif,<br />

France; 9 Intermountain Medical Center, Murray, UT; 10 Liver<br />

Unit, IDIBAPS and CIBEREHD, Barcelona, Spain; 11 University of<br />

Colorado Denver, Aurora, CO; 12 Beth Israel Deaconess Medical<br />

Center, Boston, MA<br />

Background and Aims: In the SOLAR-1 and SOLAR-2 <strong>studies</strong>,<br />

ledipasvir/sofosbuvir (LDV/SOF)+ribavirin (RBV) for 12 or 24<br />

weeks resulted in high SVR rates in genotype 1 or 4 HCV-infected<br />

patients with decompensated cirrhosis or who were liver<br />

transplant recipients. In this large combined post hoc analysis<br />

of on-treatment HCV RNA levels in SOLAR-1 and SOLAR-2 study<br />

participants we investigate whether on-treatment response varied<br />

by patient population and/or was predictive of treatment<br />

outcome. Methods: Data from the identically designed SOLAR1<br />

(NCT01938430) and SOLAR 2 (NCT02010255) <strong>studies</strong> were<br />

combined. Six groups of genotype 1 or 4 HCV-infected patients<br />

were randomized to receive 12 or 24 weeks of LDV/SOF+RBV<br />

treatment: patients without transplant and either 1) Child-Pugh-<br />

Turcotte (CPT) B cirrhosis, or 2) CPT C cirrhosis; or patients<br />

who have undergone transplantation and who were either 3)<br />

without cirrhosis (F0 to F3), 4) CPT A cirrhosis, 5) CPT B cirrhosis,<br />

or 6) CPT C cirrhosis. Patients with FCH were excluded


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 767A<br />

from this analysis. For analysis of early viral kinetics, the 12<br />

and 24 week treatment durations were combined. Serum HCV<br />

RNA was quantified using Roche CAP/CTM v2.0 with a lower<br />

limit of quantitation (LLOQ) of 15 IU/mL. Results: Patients with<br />

advanced liver disease had SVR12 rates of 86-89%. Patients<br />

who were post liver transplant had SVR rates that ranged from<br />

96%-98% in those with F0-F3 fibrosis to 60-75% of patients<br />

with severe hepatic impairment. Rapid HCV RNA declines<br />

were observed in all treatment groups. The majority of subjects<br />

achieved HCV RNA


768A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Benjamin Maasoumy - Advisory Committees or Review Panels: Abbott Molecular,<br />

Janssen-Cilaq; Grant/Research Support: Abbott Molecular; Speaking and Teaching:<br />

MSD, Roche Diagnostics, Roche Pharma, Janssen-Cilaq, Fujirebio, BMS<br />

Fiona Marra - Advisory Committees or Review Panels: Gilead, AbbVie, Merck;<br />

Consulting: Gilead; Speaking and Teaching: Gilead, Abbvie, Merck, Janssen<br />

Kerstin Port - Advisory Committees or Review Panels: Janssen; Speaking and<br />

Teaching: Roche, Gilead, MSD, Janssen<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

The following authors have nothing to disclose: Christoph Hoener zu Siederdissen,<br />

Katja Deterding, David J. Back<br />

1132<br />

Interferon Signaling and the Innate Immune Response:<br />

Still Relevant in the Age of Interferon Free Treatment<br />

Kate E. Childs, Ivana Carey, William M. Swanson, Suman Verma,<br />

Kosh Agarwal; Institute of Liver Studies, King’s College Hospital,<br />

London, United Kingdom<br />

Chronic innate immune activation predicts non-response to<br />

interferon based treatment for Hepatitic C Virus (HCV). Predictors<br />

of response to DAAs are not well characterized. Methods:<br />

46 patients (26 F, 20 M, median age 58 (51, 63) with HCV<br />

decompensated liver cirrhosis, Childs Pugh B were treated with<br />

12 weeks of an IFN-free regimen (77%SOF/ledipasvir+RBV<br />

and 23%SOF+DCV+RBV). EDTA plasma was taken at baseline<br />

(TW0), treatment week 4 (TW4), TW12, SVR 4 and SVR<br />

12. ELISAs for IP-10, CD26 which cleaves IP-10 to a shorter<br />

antagonistic form, IFN A, G L1, 2 and 3 were performed.<br />

Results: 37 patients responded (RES) to therapy, achieving SVR<br />

12. 9 patients experienced HCV relapse (REL) post TW12. At<br />

baseline, median IP-10 was higher in REL, CD26 trend higher<br />

(p=0.053) Baseline IFN G, L1, L2 and L3 did not differ. The<br />

kinetics of IP-10 and CD26 differed significantly between RES<br />

and REL,(fig) IP-10 was higher in REL and fell at TW4, IP-10<br />

increased in RES at TW4. Univariate logistic analysis was<br />

carried out for HCV genotype, biochemical parameters, HCV<br />

RNA, baseline IP10 and CD26, baseline IFN G and L and<br />

changes in IP-10 and CD26 on therapy. In the final model only<br />

baseline IP-10 and change in IP-10 at TW4 were significant<br />

(p=0.043 and p=0.045 respectively) This model was able to<br />

predict the treatment outcome in 95% of cases. Discussion- In<br />

this group of patients with decompensated cirrhosis, IP-10, a<br />

surrogate marker of hepatic ISG expression and its changes<br />

early in treatment predicted response to DAA treatment. REL<br />

had high baseline IP-10 which then decreased on treatment,<br />

whereas RES had lower baseline levels and an increase on<br />

treatment, which may mirror findings from the peg-interferon<br />

era where high baseline ISG expression could not upregulate<br />

in response to exogenous interferon. More work is warranted<br />

to investigate predictors of treatment outcome. These data suggest<br />

that the innate interferon driven response to HCV remains<br />

relevant as a predictor of response to interferon free regimens<br />

for HCV.<br />

Disclosures:<br />

Ivana Carey - Grant/Research Support: Gilead, Roche; Speaking and Teaching:<br />

BMS<br />

Kosh Agarwal - Advisory Committees or Review Panels: Gilead, BMS, Novartis,<br />

Janssen, AbbVie, Gilead; Consulting: MSD, Janssen; Grant/Research Support:<br />

Roche, Gilead, BMS, BMS; Speaking and Teaching: Astellas<br />

The following authors have nothing to disclose: Kate E. Childs, William M.<br />

Swanson, Suman Verma<br />

1133<br />

Pharmacokinetic Analyses of Ledipasvir/Sofosbuvir and<br />

HIV Antiviral Regimens in Subjects with HCV/HIV Co-infection.<br />

Polina German, Liyun Ni, Luisa M. Stamm, Jenny C. Yang, Maryanne<br />

Lenoci, John Ling, Anita Mathias; GIlead Sciences, Inc, Foster<br />

City, CA, CA<br />

Background Administration of ledipasvir (LDV)/sofosbuvir<br />

(SOF) fixed-dose combination tablet once-daily for 12 weeks<br />

achieved overall SVR12 rates of 96% (322/335) in treatment-naïve<br />

and treatment-experienced HCV/HIV co-infected<br />

subjects (ION4). Pharmacokinetic (PK) data were collected<br />

to examine the relationship between exposure and extrinsic/<br />

intrinsic variables or treatment outcome and to compare PK in<br />

co-infection and HCV mono-infection. Methods Treatment-naïve<br />

(N=150) and treatment-experienced (N=185) subjects on a<br />

stable ARV regimen of efavirenz (EFV 600 mg QD), rilpivirine<br />

(RPV; 25 mg QD) or raltegravir (RAL; 400 mg BID or 800<br />

mg QD) with a backbone of emtricitabine/tenofovir DF (FTC/<br />

TDF; 200 mg/300 mg) received LDV/SOF 90mg/400mg for<br />

12 weeks. The PK of LDV, SOF, GS-331007 (SOF predominant<br />

circulating metabolite) and tenofovir (TFV) were evaluated<br />

in all subjects with measurable plasma concentrations using<br />

previously established population PK models. EFV, RPV, RAL<br />

and FTC PK were evaluated using noncompartmental methods<br />

in subjects with intensive PK sampling data. LDV, SOF and<br />

GS-331007 exposures in HCV/HIV co-infected subjects were<br />

compared across ARV regimens, race and treatment outcome,<br />

and to subjects with HCV monoinfection. The PK of ARVs was<br />

descriptively compared to historical data from HIV monoinfection<br />

<strong>studies</strong>. Results Exposures of LDV, SOF and GS-331007<br />

were comparable across ARV regimens, race and treatment<br />

outcome. There were no clinically relevant differences in the<br />

PK of LDV/SOF in HCV/HIV co-infected subjects compared to<br />

HCV-monoinfected subjects. EFV, RPV, RAL and FTC PK were<br />

consistent with historical data. In agreement with the findings<br />

from the Phase 1 evaluations, which revealed higher TFV<br />

exposure with LDV/SOF, TFV PK was modestly higher than<br />

observed historically with NNRTI or InSTI-based regimens plus<br />

FTC/TDF. Conclusion PK results in conjunction with safety and<br />

efficacy data support the use of LDV/SOF 90 mg/400 mg for


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 769A<br />

12 weeks in HCV/HIV co-infection. Higher TFV exposures are<br />

observed in the presence of LDV/SOF; therefore, monitoring<br />

for TFV-associated AEs is warranted during co-administration<br />

of LDV/SOF and TDF-based regimens.<br />

Disclosures:<br />

Polina German - Employment: GIlead Sciences, Inc; Stock Shareholder: GIlead<br />

Sciences, Inc<br />

Luisa M. Stamm - Employment: Gilead Sciences<br />

Jenny C. Yang - Employment: Gilead Sciences, Inc<br />

Anita Mathias - Employment: Gilead Sciences Inc.,<br />

The following authors have nothing to disclose: Liyun Ni, Maryanne Lenoci, John<br />

Ling<br />

1134<br />

HCV genotype 3: Meta-analysis with current options<br />

Javier Ampuero 1 , K. Rajender Reddy 2 , Manuel Romero-Gomez 1 ;<br />

1 Unit for the Clinical Management of Digestive Diseases, Valme<br />

University Hospital, Sevilla, Spain; 2 Division of Gastroenterology<br />

and Hepatology, Hospital of the University of Pennsylvania,<br />

Philadephia, PA<br />

Background & Aim: Landscape of HCV therapy is changing<br />

rapidly with new direct acting antivirals (DAAs), with sustained<br />

virological response (SVR) rates around 90%. However, genotype<br />

3 represents the new challenge, as these patients show<br />

lower SVR rates, particularly cirrhotic and treatment-experienced<br />

patients. Many treatment options have been used but the<br />

definite therapy remains unclear. Our aim was to address therapeutic<br />

efficacy of the various treatment regimens through a<br />

meta-analysis in HCV genotype 3: a) sofosbuvir (SOF) + peginterferon<br />

(IFN) + ribavirin (RBV) 12 weeks vs. SOF+RBV 24w<br />

(n=645); b) SOF+RBV 12w/16w vs. SOF+RBV 24w (n=503);<br />

c) SOF + daclatasvir (DCV) vs. SOF+DCV+RBV (n=162); d)<br />

SOF + ledipasvir (LDV) vs. SOF+LDV+RBV (n=169). Methods:<br />

We reviewed and identified <strong>studies</strong> in Pubmed and EMBASE<br />

until June 2015. We included certain number of published<br />

papers and presented abstracts (last AASLD and EASL meetings)<br />

in genotype 3 patients randomized to at least, 2 arms<br />

of HCV therapy. The random effect models of Der Simonian<br />

and Laird method were used for heterogeneous <strong>studies</strong> using<br />

the Meta-Disc software 1.4 (Madrid, Spain). Results: Patients<br />

taking SOF+INF+RBV 12w (92.9%; 196/211) showed higher<br />

SVR rates than those treated with SOF+RBV 24w (79.4%;<br />

201/253) (OR 2.67 (95%CI 1.43-4.98); I 2 =0%) (Figure).<br />

Twenty-four weeks of combination SOF+RBV had higher SVR<br />

rates (84.4%; 391/463) than those treated for 12/16w<br />

(70.7%; 157/222) (OR 3.45 (95%CI 1.02-11.11); I 2 =70%).<br />

In SOF+DCV-based therapies, the addition of RBV (90.3%;<br />

28/31) did not enhance SVR rates significantly (81.7%;<br />

107/131) (OR 1.16 (95%CI 0.44-3.08); I 2 =0%). In contrast,<br />

the addition of RBV ((81%; 111/137) vs. (62.5%; 20/32)) was<br />

able to increase the SVR rates in patients treated with SOF+LDV<br />

(OR 3.30 (95%CI 1.35-8.33); I 2 =0%). Conclusion: Triple therapy<br />

of SOF+IFN+RBV has demonstrated better SVR rates than<br />

IFN-free regimens in genotype 3 patients. To optimize SVR,<br />

interferon free regimen of SOF+RBV needs to be administrated<br />

over longer period (24 weeks) rather than shorter duration of<br />

12w-16 weeks. With SOF+LDV combination, RBV is essential<br />

to increase SVR rates probably because LDV has a very low<br />

activity against genotype 3. By contrast, the addition of RBV in<br />

DCV-based treatment does not increase SVR.<br />

Disclosures:<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

Manuel Romero-Gomez - Advisory Committees or Review Panels: Roche Farma,SA.,<br />

MSD, S.A., Janssen, S.A., Abbott, S.A.; Grant/Research Support: Ferrer,<br />

S.A.<br />

The following authors have nothing to disclose: Javier Ampuero<br />

1135<br />

Utilization and Effectiveness of Sofosbuvir Based HCV<br />

Regimens in a Community Based Setting<br />

T. Craig Cheetham 1 , Sara Y. Tartof 1 , Zoe Bider-Canfield 1 , Vincent<br />

Louie 3 , Derenik Gharibian 3 , Amandeep K. Sahota 2 ; 1 Research &<br />

Evaluations Department, Kaiser Permanente Southern California,<br />

Pasadena, CA; 2 Gastroenterology/Hepatology, Southern California<br />

Permenente Medical Group, Los Angeles, CA; 3 Pharmacy<br />

Department, Kaiser Foundation Hospitals, Los Angeles, CA<br />

Background: High rates of sustained virologic response have<br />

been reported in the clinical trials for sofosbuvir (Sof) based<br />

regimens. Less information is available regarding the effectiveness<br />

(sustained virologic response 12 weeks, SVR12) of Sofbased<br />

regimens outside of clinical trials. Objective: Determine<br />

the rates of SVR12, for Sof-based regimes, in a community<br />

setting. Methods: Kaiser Permanente Southern California is an<br />

integrated health-care delivery system providing medical services<br />

to 3.8 million members. Adults started on a Sof-based<br />

regimen beginning 12/1/2013 and completing treatment by<br />

1/31/2105 were eligible. Six months continuous membership<br />

prior to starting therapy was required. Co-administered<br />

drugs included pegylated-interferon (PEG), simeprevir (Sim)<br />

and ribavirin (Riba). Baseline co-morbid conditions, viral load,<br />

genotype (GT), and prior Hepatitis C (HCV) treatments were<br />

collected from the electronic medical record. Viral load tests<br />

were collected through 4/30/2015 allowing 12 weeks follow-up<br />

for all patients. Descriptive and multivariable analyses<br />

were conducted to evaluate the utilization and outcomes of<br />

Sof-based regimens. SVR12 was the primary endpoint. Results:<br />

A total of 355 patients were identified. Twelve patients did not<br />

complete therapy and were excluded (2 died, 6 experienced<br />

adverse events, and 4 were stopped for non-adherence). For<br />

the cohort the average age was 58.5 years and 64.7% were<br />

male. Cirrhosis was present at baseline in 39.7% of patients<br />

and 43.7% were treatment experienced. Genotypes 1, 2 & 3<br />

accounted for 97.1% of the HCV cases (100 - GT1, 194 - GT2,<br />

39 - GT3). The most common regimens were Sof-Rib (n=254),<br />

Sim-Sof (n=12) and Sof-PEG-Rib (n=75). The overall SVR12<br />

rate was 82.2%. The table has a breakdown of SVR12 by<br />

genotype and major risk factors. In general, higher rates of<br />

SVR12 were seen in treatment naïve and non-cirrhotic patients.<br />

Rates of SVR12 were greater than 90% in GT2 patients (almost<br />

all received Sof-Riba). In the multivariable logistic regression<br />

models, cirrhosis was associated with a higher risk of failure.<br />

Conclusion: In a community based setting, Sof base regimens<br />

were associated with high SVR12 rates in GT2 patients. Significantly<br />

lower rates of SVR12 were seen in GT1 and GT3, which<br />

may be due to the high number of treatment experienced and<br />

cirrhotic patients.


770A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

SVR12 Rates by Major Risk Factors<br />

Demographics(%)<br />

Disclosures:<br />

T. Craig Cheetham - Grant/Research Support: Gilead, BMS<br />

The following authors have nothing to disclose: Sara Y. Tartof, Zoe Bider-Canfield,<br />

Vincent Louie, Derenik Gharibian, Amandeep K. Sahota<br />

1136<br />

Evaluation of an Integrated HIV/HCV Care Model in an<br />

Urban, Low-Income, HIV/HCV Co-Infection Clinic<br />

Angela Kapalko 1 , Linden Lalley-Chareczko 1 , Meghan F. Fibbi 2,3 ,<br />

Karam Mounzer 1,4 ; 1 Jonathan Lax Treatment Center, Philadelphia<br />

FIGHT, Philadelphia, PA; 2 Department of Family and Community<br />

Medicine, University of Pennsylvania, Philadelphia, PA; 3 Philadelphia<br />

College of Osteopathic Medicine, Philadelphia, PA; 4 Perelman<br />

School of Medicine, University of Pennsylvania, Philadelphia,<br />

PA<br />

Background The Jonathan Lax Treatment Center is an urban,<br />

low-income, Ryan White and FQHC funded, outpatient medical<br />

center providing services via an integrated care model to<br />

approximately1700 patients with HIV, of whom 459 are HIV/<br />

Hepatitis C(HCV) co-infected. The HIV/HCV co-infection treatment<br />

protocol at the Lax Center incorporates on-site disease<br />

staging, pharmacy refills/pill-packing services with flexible<br />

dispensation schedules, adherence monitoring, case management<br />

and mental health services.The purpose of the present<br />

observational study is to evaluate the comprehensive HIV/HCV<br />

co-infection treatment protocol in a patient-centered care setting<br />

at the Lax Center. Methodology A comprehensive chart review<br />

of all Lax Center patients carrying a diagnosis of chronic<br />

HCV between the calendar years of 2011 and 2014(n=459)<br />

was performed. Patient demographics, HCV genotyping and<br />

disease staging, prior HCV treatment experience, HCV sustained<br />

virologic response(SVR) and information on comorbid<br />

conditions present during HCV treatment were recorded. All<br />

co-infected patients were included in the overall demographic<br />

analysis; however, patients treated on a clinical trial were<br />

excluded from the treatment outcome analyses. Results 459<br />

co-infected patient records active between Jan 1, 2011 and<br />

Dec 31 2014 were reviewed. 88 individual patients(19.1%,<br />

n=459) received treatment during this time frame, 10 of whom<br />

were retreated, making the total number of treatments delivered<br />

98. 26 treatments were excluded due to participation in<br />

clinical trials, leaving 72 eligible treatments for final analyses.<br />

The demographics of those treated were similar to the overall<br />

Lax HIV/HCV population (See Table 1). 31.9% of all treatment<br />

efforts consisted of Telaprevir based regimens with an<br />

SVR12 rate 56.5%(n=23). IFN free regimens (SOF/RBV, SOF/<br />

SMV, SOF/LDV)(n=43) yielded an SVR12(n=39) of 87.1%<br />

with 4 SVR12 results pending(2 confirmed SVR4). No patients<br />

were lost-to-follow-up while on treatment or during post-treatment<br />

monitoring. Conclusions Operating under a patient-centered,<br />

integrated care model, the Lax Center was able to treat<br />

19.1%(n=88) of active HIV/HCV patients, retain 100% in<br />

care, and achieved an overall SVR rate of 75.0%(with 4 SVR<br />

12 pending). These data support the utility of an integrated<br />

care model for optimization of treatment outcome among HIV/<br />

HCV co-infected patients in an urban, low-income community<br />

setting.<br />

Disclosures:<br />

Angela Kapalko - Advisory Committees or Review Panels: Gilead Sciences, Abbvie;<br />

Speaking and Teaching: Abbvie<br />

Karam Mounzer - Advisory Committees or Review Panels: Gilead Sciences, BMS,<br />

J& J, Merck; Grant/Research Support: Boehrlingher Ingelheim, Gilead Sciences,<br />

Glaxo Smith Kline ViiV, Janssen and Janssen, Merck; Speaking and Teaching:<br />

Gilead Sciences, BMS, J& J, Merck<br />

The following authors have nothing to disclose: Linden Lalley-Chareczko, Meghan<br />

F. Fibbi<br />

1137<br />

What does Fibroscan® measure shortly after IFN-free<br />

DAA therapy?<br />

Karin Kozbial, Albert Stättermayer, Clarissa Freissmuth, Rafael<br />

Stern, Sandra Beinhardt, Christian Rechling, Petra E. Steindl-<br />

Munda, Michael Trauner, Peter Ferenci, Harald Hofer; Internal<br />

Medicine 3, Medical University of Vienna, Vienna, Austria<br />

Background:Sustained virological response (SVR) by any antiviral<br />

therapy may improve liver fibrosis in the majority of patients<br />

during long-term follow up. Interferon free combinations of<br />

direct acting antivirals (DAAs) achieve a rapid and profound<br />

inhibition of HCV replication during treatment and offer the<br />

possibility to study short-term changes in fibrosis.Therefore<br />

this study investigated liver stiffness in patients with advanced<br />

fibrosis (F3-F4) shortly after DAA treatment. Methods:Liver stiffness<br />

was measured before and at the end or up to 12 weeks<br />

after the end of DAA-therapy (Sofosbuvir[SOF] combined with<br />

daclatasvir[DCV, n=52], ribavirin [n=10], ledipasvir [n=14],<br />

Simeprevir [SMV, n=24], or the combination of DCV and SMV<br />

[n=9])by Fibroscan®(FS; Echosens 502, Paris, France) in 109<br />

patients (f/m=40/69; mean age: 55.2±10.5yrs; genotype<br />

(GT) 1: 87, GT3: 16, GT4: 6; F3/F4=28/81; BMI 27.1±4.4).<br />

Treatment duration was 12 (n=48) up to 24 weeks (n=61).<br />

Only FS results with IQR ≤ 30% were included. Liver fibrosis categories<br />

were defined as advanced fibrosis (F3:9.6-12.4kPa),<br />

and cirrhosis (F4: ≥12.5kPa). Results:Overall, liver stiffness<br />

decreased from 23.2kPa (95% CI 20.4-26.0) to 20.1kPa (95%<br />

CI 17.0-23.3;p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 771A<br />

Harald Hofer - Advisory Committees or Review Panels: Gilead, Abbvie; Speaking<br />

and Teaching: Janssen, BMS, Gilead, Abbvie<br />

The following authors have nothing to disclose: Karin Kozbial, Albert Stättermayer,<br />

Clarissa Freissmuth, Rafael Stern, Sandra Beinhardt, Christian Rechling,<br />

Petra E. Steindl-Munda<br />

1138<br />

Association between severe rash and HLA polymorphism<br />

in telaprevir-based triple therapy for chronic hepatitis<br />

C in a multi-centric study in Japan<br />

Yoko Yamagiwa 1 , Naohiko Masaki 1 , Nao Nishida 1 , Minae<br />

Kawashima 2 , Seik-Soon Khor 2 , Hiroko Miyadera 1 , Masaya Sugiyama<br />

1 , Masaaki Korenaga 1 , Kazumoto Murata 1 , Tatsuya Kanto 1 ,<br />

Yasuhito Tanaka 3 , Kazuhiko Koike 4 , Katsushi Tokunaga 2 , Masashi<br />

Mizokami 1 ; 1 The Research Center for Hepatitis and Immunology,<br />

National Center for Global Health and Medicine, Ichikawa, Japan;<br />

2 Department of Human Genetics, Graduate School of Medicine<br />

The University of Tokyo, Tokyo, Japan; 3 Department of Virology<br />

and Liver Unit, Nagoya City University Graduate School of Medical<br />

Sciences, Nagoya, Japan; 4 Department of Gastroenterology,<br />

The University of Tokyo, Tokyo, Japan<br />

Background and Aim Although telaprevir (TVR) enabled a<br />

remarkable improvement in efficacy of interferon treatment<br />

for chronic hepatitis C, it had serious adverse events such as<br />

dermatological adverse events which required discontinuance<br />

of treatment in serious grade of rash which was potentially<br />

life-threatening. However, mechanism of severe rash in TVRbased<br />

therapy has not been clarified yet. Thus, we aimed to<br />

identify genetic risk factors associated with serious grade of<br />

rash. Methods and Results Genome-wide association study<br />

(GWAS) was performed in 384 patients received TVR combined<br />

with ribavirin and peg-interferon using Affymetrix<br />

AXIOM ASI arrays. We compared allele frequency of each single<br />

nucleotide polymorphism (SNP) between 182 patients who<br />

did not experience skin eruption and 50 patients with grade 3<br />

of rash defined by rash over 50% of body surface area, rash<br />

with additional changes such as vesicles, or with systematic<br />

reaction such as fever. However, no significant association<br />

was identified in the GWAS stage. Next, SNP imputation and<br />

association testing was performed using PLINK in 141 SNPs<br />

with p < 1.00E-05 in the GWAS stage. As a result, 22 out<br />

of 25 SNPs with p < 1.00E-06 in the SNP imputation, were<br />

located in the human leukocyte antigen (HLA) region on chromosome<br />

6. Finally, statistical imputation analysis of classical<br />

HLA alleles was performed in the SNP data using HIBAG R<br />

package optimized for Japanese population (Pharmacogenomics<br />

J. 2015). HLA-DRB1*09:01 showed a significant association<br />

with an odds ratio (OR) of 3.376 (p = 9.12E-06) and<br />

HLA-DQB1*03:03, which is known to be in strong linkage<br />

disequilibrium with DRB1*09:01 in the Japanese population,<br />

also showed a significant association with an OR of 2.907 (p<br />

= 7.15E-05). Discussion and Conclusions More frequent severe<br />

rash in phase III trial in Japan (9%) compared with that in the<br />

USA and Europe (5%), may related with higher frequency of<br />

HLA-DRB1*09:01 in Japanese population than in Caucasian<br />

(15% vs 1%). Genetic variants located in the HLA region in<br />

Japanese should be candidate genetic factors in susceptibility<br />

to severe rash in TVR-based triple therapy.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Yoko Yamagiwa, Naohiko<br />

Masaki, Nao Nishida, Minae Kawashima, Seik-Soon Khor, Hiroko Miyadera,<br />

Masaya Sugiyama, Masaaki Korenaga, Kazumoto Murata, Tatsuya Kanto,<br />

Kazuhiko Koike, Katsushi Tokunaga, Masashi Mizokami<br />

1139<br />

Patient Reasons for Initiating or Delaying Hepatitis C<br />

Virus (HCV) Treatment and Physician Reasons for Selecting<br />

an HCV Treatment: A Prospective Observational<br />

Study<br />

Shelagh M. Szabo 2 , David R. Walker 1 , Timothy R. Juday 1 ,<br />

Suzanne Lane 2 , Sammy Saab 3 ; 1 HEOR, Abbvie, North Chicago,<br />

IL; 2 Icon Health Economics, Vancouver, BC, Canada; 3 David Geffen<br />

School of Medicine, UCLA, Los Angeles, CA<br />

BACKGROUND This is a prospective, multicenter observational<br />

study of humanistic and economic outcomes among patients<br />

initiating direct acting antiviral treatments for hepatitis C virus<br />

(HCV). The objective of this analysis is to describe patient reasons<br />

for treatment delay or initiation and physician reasons<br />

for selection of a specific treatment regimen. METHODS All<br />

consecutive HCV patients attending clinics at ten sites in the<br />

United States were screened to identify the cohort initiating<br />

treatment for HCV. From that cohort, data were collected using<br />

patient-completed surveys at up to five time points over the treatment<br />

course, supplemented by a medical chart review at treatment<br />

initiation and completion. Type of data collected included<br />

demographic, clinical, economic, and humanistic, however, for<br />

this analysis, only screening data were used. Data on patients’<br />

reasons for initiating or delaying treatment, and physicians’<br />

reasons for recommending treatment were collected; multiple<br />

responses were allowed per patient. The first patient enrolled<br />

January 2013 and the last patient enrolled September 2014.<br />

RESULTS There were 143 patients, from eight clinics, enrolled<br />

in the study out of the 787 that were screened across ten clinics.<br />

Among enrolled patients, 137 were from community clinics<br />

and 6 were from academic/hospitals. The mean age was 56<br />

years, 52% male, 60% white, 20% Hispanic and 15% black.<br />

75% graduated from high school and 33% were employed.<br />

100 patients were on a sofosbuvir and/or simeprevir-based<br />

regimen, and 43 patients were on other treatments. 50% were<br />

on an interferon-free regimen. Patient-reported reasons for initiating<br />

treatment were desire to be cured (74.1%), physician’s<br />

recommendation (70.6%), to prevent future liver damage<br />

(62.2%), peace of mind (49%), to manage HCV symptoms<br />

(31.5%), worry about transmission (26.6%), to avoid social<br />

stigma (10.5%), desire to have children (7.7%), family pressure<br />

(6.3%), and other (2.8%). Patient-reported reasons for initially<br />

delaying treatment were physician recommendation (23.8%),<br />

waiting for the availability of new treatments (21.0%), no or<br />

mild symptoms (17.5%), cost (9.1%), health conditions other<br />

than HCV (5.6%) and other (39.2%). The primary physician-reported<br />

reason for selection of a specific treatment regimen was<br />

likelihood of treatment success (83.9%), followed by tolerability<br />

(7.7%), reimbursement/access (2.8%), co-pay (0%), and other<br />

(4.9%). CONCLUSIONS The potential for achieving cure was<br />

the most important reason for a physician choosing a specific<br />

HCV treatment regimen. Cost however, did not play an important<br />

role for treatment selection or delay for physicians and<br />

patients, respectively.<br />

Disclosures:<br />

Shelagh M. Szabo - Employment: Oxford Outcomes<br />

David R. Walker - Employment: Abbvie; Stock Shareholder: Abbvie, Baxter<br />

Healthcare<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Sammy Saab - Advisory Committees or Review Panels: BMS, Gilead, Merck,<br />

Janssen; Grant/Research Support: Gilead; Speaking and Teaching: BMS, Gilead,<br />

Merck, Janssen, Salix, Onyx, Bayer, Janssen; Stock Shareholder: Achillion,<br />

Johnson and Johnson, BMS, Gilead<br />

The following authors have nothing to disclose: Suzanne Lane


772A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1140<br />

Co-morbidities and co-medications of patients with<br />

chronic hepatitis C (CHC) under specialist care in the UK<br />

- Challenges for scaling up HCV treatment?<br />

Benjamin E. Hudson 3,8 , William Irving 4,5 , Stephen T. Barclay 1 ,<br />

Fiona Marra 2,6 , Alex J. Walker 7 ; 1 Walton Liver Clinic, Glasgow,<br />

United Kingdom; 2 NHS Greater Glasgow and Clyde, Glasgow,<br />

United Kingdom; 3 HCV Research UK, Nottingham, United Kingdom;<br />

4 NIHR Nottingham Digestive Diseases Biomedical Research<br />

Unit, Nottingham, United Kingdom; 5 School of Life Sciences, University<br />

of Nottingham, Nottingham, United Kingdom; 6 University of<br />

Liverpool, Liverpool, United Kingdom; 7 University of Nottingham,<br />

School of Life Sciences, Nottingham, United Kingdom; 8 Cambridge<br />

University Hospitals NHS trust, Cambridge, United Kingdom<br />

Introduction Most patients with CHC can benefit from direct<br />

acting antiviral (DAA). Alongside specific CHC disease indicators,<br />

physical/psychiatric comorbidities, lifestyle factors<br />

and related co-medications with drug-drug interaction (DDI)<br />

potential will inform appropriate CHC treatment selection. We<br />

describe demographics, comorbidities and common co-medications<br />

of CHC patients currently under specialist care in the<br />

UK. Methods Retrospective analysis of routinely collected data<br />

of CHC patients from 59 UK specialist centres enrolled with<br />

the National HCV Research UK Biobank between 03/2012-<br />

10/2014. Results 6,278 patients were analysed. Median age<br />

52 years (IQR 43-59), 70.4% male, 84.7% white. Genotype<br />

1: 50%, 2: 4.0%, 3: 33.7%, 4: 3.6%, 5/6: 0.3%. Cirrhotic:<br />

23.6% (age >60’s 36.6%). Acquisition mode: injection drug<br />

use (IDU) 59.2%, blood products 11.3%, non-UK born 9.4%,<br />

other 20.1%. Subgroup analysis for IDU acquisition and age<br />

was performed. Social history (Hx): previous/current heavy<br />

alcohol use: overall 38.3% (IDU 50.1%), current cannabis use:<br />

24.6% (IDU 35.2%). Comorbidities: Hx of depression: 45.4%<br />

(IDU 57.6%), Hx of attempted suicide or inpatient treatment<br />

for depression: 19.4% (IDU 27.3%). HIV co-infection 5.0%<br />

(IDU: 4.5%). Prevalence of diabetes, cancer and renal dialysis<br />

correlated with age (Diabetes: overall 11.3%, >60s 22.9%,<br />

Cancer: 8.1% vs 18.1%, Dialysis: 1.3% vs 2.2%). Most common<br />

co-medications with DDI potential were psychotropic with<br />

38.6% (IDU 53.3%) on either antidepressant (22.9%), opioid<br />

replacement (21.2%) or hypnotic (10.4%) (includes polypharmacy).<br />

Prescriptions of antidiabetics, statins and immunosuppressants<br />

rose sharply with age (60’s 31.7%).<br />

Conclusions We found high levels of co-morbidity, ongoing<br />

substance abuse and co-medication with DDI potential in<br />

CHC populations under specialist care in the UK. Age-related<br />

increasing levels of advanced liver disease, comorbidities and<br />

polypharmacy may further increase the challenges in managing<br />

this patient group. CHC treaters need to be aware of DDI<br />

potential when choosing appropriate CHC treatments.<br />

*antidepressants, opioid replacement, hypnotics;** Antiretrovirals;**not<br />

specified<br />

Disclosures:<br />

William Irving - Advisory Committees or Review Panels: Novartis, MSD, Janssen<br />

Cilag, Bristol Myers Squibb; Grant/Research Support: GSK, Pfizer, Janssen<br />

Cilag, Gilead Sciences; Speaking and Teaching: Janssen Cilag, Roche<br />

Stephen T. Barclay - Advisory Committees or Review Panels: Gilead, Janssen,<br />

MSD, Abbvie, BMS; Speaking and Teaching: Gilead, Janssen<br />

Fiona Marra - Advisory Committees or Review Panels: Gilead, AbbVie, Merck;<br />

Consulting: Gilead; Speaking and Teaching: Gilead, Abbvie, Merck, Janssen<br />

The following authors have nothing to disclose: Benjamin E. Hudson, Alex J.<br />

Walker<br />

1141<br />

Patterns of Care Since the Approval of New Therapies<br />

for Hepatitis C Virus (HCV) Infection: A 2013-2014 Real-<br />

World Analysis of US Community Specialty Practices<br />

Paul J. Gaglio 2 , Tamar Sapir 1 , Jeffrey D. Carter 1 , Laurence<br />

Greene 1 , Erica Rusie 1 , Kathleen Moreo 1 ; 1 PRIME Education, Inc.,<br />

Tamarac, FL; 2 Hepatology, Montefiore Medical Center, Bronx, NY<br />

Background: From 2013 to 2014, approval of all-oral therapies<br />

was associated with dramatic changes in managing HCV.<br />

To assess “real-world” use of HCV therapy and management<br />

trends, we conducted a retrospective study of 50 US community<br />

practices. Methods: The cohort comprised 17 hepatologists,<br />

17 gastroenterologists, and 16 infectious disease specialists<br />

across the US. Charts of adult HCV patients were reviewed for<br />

2013 (n = 300) and 2014 (n = 500). Charts were abstracted<br />

for patient and disease characteristics, care processes, adherence<br />

to HCV quality measures, and prescribing patterns. T-tests<br />

or chi-square tests were used to analyze differences in means<br />

for 2013 and 2014. Results: From 2013 to 2014, patient<br />

demographics, risk factors, disease characteristics, and comorbidities<br />

were as follows: average age (51 vs 54 yrs p=.07),<br />

time since diagnosis (8 vs 5 yrs, p=.04), GT1 (75% vs 76%,<br />

p=.75), GT2 (9% vs 10%, p=.77), GT3 (13% vs 10%, p=.15),<br />

substance abuse (41% vs 33%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 773A<br />

1142<br />

Treatment Outcomes of Veterans with Decompensated<br />

Cirrhosis Receiving All Oral Direct Acting Antiviral Hepatitis<br />

C Therapy<br />

Abigail Rabatin 1 , Mohamed Hashem 1 , Mary L. Townsend 1 , William<br />

E. Bryan 1 , Cynthia A. Moylan 1,2 , Steve S. Choi 1,2 , Susanna<br />

Naggie 1,2 ; 1 Durham VA Medical Center, Raleigh, NC; 2 Duke University<br />

Medical Center, Durham, NC<br />

Purpose Patients infected with hepatitis C virus (HCV) and decompensated<br />

cirrhosis are a therapeutic challenge. The purpose of<br />

this study was to evaluate the effectiveness and safety of oral<br />

sofosbuvir-based regimens in Veterans with decompensated cirrhosis.<br />

Design This is a multicenter, retrospective, cohort study<br />

utilizing the national Veterans Affairs (VA) electronic medical<br />

record. Veterans with decompensated cirrhosis who received<br />

a sofosbuvir-based HCV regimen at a VA between 12/06/13<br />

and 12/31/14 were included. Veterans who received interferon-containing<br />

regimens were excluded. Possible decompensation<br />

was identified using VA outpatient prescription data for<br />

either lactulose, rifaximin or sorafenib. Manual chart review<br />

was conducted confirming decompensating event (bleeding<br />

esophageal varices, hepatic encephalopathy, ascites, spontaneous<br />

bacterial peritonitis or hepatocellular carcinoma). The<br />

primary objective was sustained virologic response defined as<br />

an undetectable HCV RNA 12 weeks post-treatment (SVR12).<br />

This study was approved by the Durham VA IRB. Results A total<br />

of 7,622 Veterans received sofosbuvir during the study period<br />

and 517 met criteria based on prescription data for possible<br />

decompensation. A convenience sample of 250 was assessed<br />

by manual chart review, of which 150 met inclusion criteria.<br />

84% (N=126) completed treatment and 16% (N=24) discontinued<br />

treatment early. The majority of Veterans were male<br />

(97%), Caucasian (70%), with genotype 1-infection (84%) and<br />

a mean age of 61 years. Median platelet count was 78 and<br />

median albumin 3.1. There were 5 patients co-infected with<br />

HBV and 5 with HIV. SVR12 is provided in Table 1. Relapse<br />

was the only reason for virologic failure in Veterans completing<br />

therapy. A majority of Veterans (69%) experienced at least one<br />

adverse event (AE), most commonly fatigue (26%) and nausea<br />

(11%). Severe AEs were uncommon (8%, N=12) and included<br />

3 deaths that were not treatment related. Four patients died<br />

after end of treatment. Conclusion All oral direct acting antiviral<br />

regimens containing sofosbuvir are safe in Veterans with<br />

decompensated cirrhosis; however, relapse rates were high<br />

in this cohort. Patients with decompensated cirrhosis remain a<br />

special population with unmet medical need and a high mortality<br />

regardless of therapy.<br />

Table 1<br />

1143<br />

Cost-effectiveness of ombitasvir/paritaprevir/ritonavir,<br />

dasabuvir +/- ribavirin (3D±R) in patients with genotype<br />

(GT) 1 chronic hepatitis C virus (HCV) and human<br />

immunodeficiency virus (HIV) coinfection compared with<br />

other standards of care in the US<br />

Suchin Virabhak 2 , Scott J. Johnson 2 , Hélène Parisé 2 , Timothy R.<br />

Juday 1 , Alice Wang 1 , Yuri Sanchez 1 , Sammy Saab 3 ; 1 HEOR,<br />

AbbVie, Mettawa, ID; 2 Medicus Economics Inc., Boston, MA;<br />

3 Pfleger Liver Institute, UCLA, Los Angeles, CA<br />

BACKGROUND 3D±R has demonstrated safety and efficacy<br />

in patients with GT1 HCV and HIV coinfection. Its cost-effectiveness<br />

has not been evaluated vs no treatment (NT) or other<br />

standards of care in the US, including sofosbuvir plus ledipasvir<br />

(SOF+LDV), sofosbuvir plus pegylated-interferon and ribavirin<br />

(SOF+PR), and SOF+R. METHODS A cost-effectiveness Markov<br />

model was developed based on a natural HCV history with 13<br />

health states: 8 disease progression states (F0-F4, decompensated<br />

cirrhosis, hepatocellular carcinoma, and liver transplant),<br />

3 sustained virologic response states, a spontaneous clearance<br />

state, and a mortality state. Transition rates were obtained<br />

from published literature. Adverse event rates, treatment-related<br />

disutilities, treatment durations and efficacy rates were based<br />

on the following clinical trials: TURQUOISE I (3D±R), ERADI-<br />

CATE (SOF+LDV), PHOTON-1 and PHOTON-2 (SOF+R), and<br />

P7977-1910 (SOF+PR). Baseline patient characteristics were<br />

based on TURQUOISE I. Direct medical costs, including HIV<br />

antiretroviral therapy, were extracted from a systematic literature<br />

review and drug costs were based on the December 2014<br />

Red Book. The model had a lifetime horizon with an annual<br />

discount rate of 3%. Outcomes were measured in quality-adjusted<br />

life-years (QALYs), lifetime costs, and incremental cost<br />

effectiveness ratios (ICERs) in the overall coinfected population<br />

and the treatment-naïve non-cirrhotic subpopulation, where<br />

data were available. Probabilistic sensitivity analyses (PSA)<br />

were conducted by varying all parameters simultaneously.<br />

RESULTS In the overall coinfected population, 3D±R was “dominant”<br />

over SOF+R (i.e., was less costly and more effective)<br />

and was cost-effective at a threshold of $50,000 per QALY<br />

gained compared to NT and SOF+PR. In naïve non-cirrhotic<br />

coinfected patients, 3D±R was cost-effective compared to NT<br />

at a $50,000 threshold per QALY gained. Compared with<br />

SOF+LDV, 3D±R offers 0.1 less QALYs and $11,895 lower<br />

lifetime costs. 3D±R was the preferred regimen in the majority<br />

of PSA simulations when QALYs were valued at $50,000 or<br />

$100,000 each. CONCLUSIONS 3D±R is a cost-effective treatment<br />

option for GT1 HCV and HIV coinfected patients, both in<br />

the overall and naïve non-cirrhotic populations, compared to<br />

NT and other standards of care in the US.<br />

Disclosures:<br />

Susanna Naggie - Advisory Committees or Review Panels: BMS, Gilead, Abb-<br />

Vie, Merck; Grant/Research Support: Gilead, AbbVie, BMS, Jenssen, Merck,<br />

Achillion<br />

The following authors have nothing to disclose: Abigail Rabatin, Mohamed<br />

Hashem, Mary L. Townsend, William E. Bryan, Cynthia A. Moylan, Steve S. Choi<br />

NA: No clinical trial data available<br />

Disclosures:<br />

Suchin Virabhak - Consulting: AbbVie<br />

Scott J. Johnson - Consulting: AbbVie; Employment: Medicus Economics<br />

Hélène Parisé - Consulting: MedicusEconomics LLC, AbbVie ; Employment: Statlog<br />

Consulting Inc<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Yuri Sanchez - Employment: AbbVie; Stock Shareholder: AbbVie


774A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Sammy Saab - Advisory Committees or Review Panels: BMS, Gilead, Merck,<br />

Janssen; Grant/Research Support: Gilead; Speaking and Teaching: BMS, Gilead,<br />

Merck, Janssen, Salix, Onyx, Bayer, Janssen; Stock Shareholder: Achillion,<br />

Johnson and Johnson, BMS, Gilead<br />

The following authors have nothing to disclose: Alice Wang<br />

1144<br />

Lifetime risks of liver morbidity and mortality in patients<br />

with recurrent genotype 1 hepatitis C virus infection<br />

post-liver transplantation treated with ombitasvir/paritaprevir/ritonavir,<br />

dasabuvir and ribavirin (3D+R) vs<br />

former standards of care<br />

Sammy Saab 2 , Caroline Huber 3 , Alice Wang 1 , Yuri Sanchez 1 ,<br />

Timothy R. Juday 1 ; 1 HEOR, AbbVie, Mettawa, ID; 2 Pfleger Liver<br />

Institute, UCLA, Los Angeles, CA; 3 Precision Health Economics, Los<br />

Angeles, CA<br />

PURPOSE Liver transplant recipients with hepatitis C virus<br />

(HCV) infection historically have had limited treatment options.<br />

Without effective treatment, HCV can cause progressive liver<br />

disease in a short period of time. 3D+R is a recently approved<br />

interferon-free regimen with demonstrated safety and efficacy<br />

in the post-transplant genotype (GT) 1 HCV population and its<br />

long-term impact on liver outcomes is not yet fully understood.<br />

We assess the lifetime risks of liver morbidity and mortality<br />

for post-liver transplant patients with GT1 HCV recurrence<br />

receiving no treatment (NT) or treatment regimens with FDA<br />

approval or Phase III trial results. METHODS A US-based, twophase<br />

Markov health state model estimated outcomes over a<br />

lifetime horizon for a cohort of post-liver transplant patients<br />

with recurrent GT1 HCV and METAVIR fibrosis stage F0-F2.<br />

In the first phase, patients were distributed into disease states,<br />

including sustained virologic response, based on the efficacy of<br />

three treatment options: NT, pegylated interferon and ribavirin<br />

for 48 weeks (PR48), or 3D+R for 24 weeks. The second phase<br />

followed patients’ progression through compensated cirrhosis<br />

(CC), decompensated cirrhosis (DCC) and death. Transition<br />

rates were extracted from published literature. Population characteristics<br />

and treatment efficacy rates for PR48 and 3D+R<br />

were based on Phase III clinical trials in post-liver transplant<br />

patients. Lifetime risks of CC, DCC and 10-year all-cause mortality<br />

were analyzed. RESULTS The lifetime risks of liver morbidity<br />

and all-cause mortality were substantially lower for GT1<br />

HCV patients with liver transplant treated with 3D+R than those<br />

receiving NT or PR48. For a representative patient, the lifetime<br />

risk of developing CC was 63.3% with NT, 51.5% with PR48,<br />

and 1.9% with 3D+R. The lifetime risk of developing DCC<br />

was 55.6% with NT, 45.3% with PR48 and 1.7% with 3D+R.<br />

All-cause mortality after 10 years was 29.5% with NT, versus<br />

27.4% with PR48 and 18.6% with 3D+R treatment. CONCLU-<br />

SIONS The likelihood of liver transplant recipients infected<br />

with GT1 HCV experiencing liver morbidity (i.e., CC or DCC)<br />

or all-cause mortality 10 years post-treatment was substantially<br />

reduced with 3D+R compared to patients treated with PR48 or<br />

not treated.<br />

Disclosures:<br />

Sammy Saab - Advisory Committees or Review Panels: BMS, Gilead, Merck,<br />

Janssen; Grant/Research Support: Gilead; Speaking and Teaching: BMS, Gilead,<br />

Merck, Janssen, Salix, Onyx, Bayer, Janssen; Stock Shareholder: Achillion,<br />

Johnson and Johnson, BMS, Gilead<br />

Caroline Huber - Employment: Precision Health Economics<br />

Yuri Sanchez - Employment: AbbVie; Stock Shareholder: AbbVie<br />

Timothy R. Juday - Employment: AbbVie; Stock Shareholder: AbbVie<br />

The following authors have nothing to disclose: Alice Wang<br />

1145<br />

Quantifying and Predicting Early Clinical Gains with<br />

Direct Acting Antiviral Therapy in Well Compensated<br />

HCV Cirrhosis: Is There an Unrecognized “Point of<br />

No MELD-based Return” Despite Sustained Virologic<br />

Response?<br />

Meera Ramanathan 1 , David W. Backstedt 2 , Mark Pedersen 1 ,<br />

Myunghan Choi 4 , Anil B. Seetharam 3 ; 1 Internal Medicine, Banner<br />

University Medical Center, Phoenix, AZ; 2 Gastroenterology,<br />

Banner University Medical Center, Phoenix, AZ; 3 Banner Transplant<br />

and Advanced Liver Disease Center, University of Arizona<br />

College of Medicine-Phoenix, Phoenix, AZ; 4 College of Nursing<br />

and Healthcare Innovation, Arizona State University, Phoenix, AZ<br />

Rationale: Direct acting antiviral (DAA) regimens for Hepatitis<br />

C (HCV) achieve high sustained virologic response (SVR12)<br />

rates. Abrogation of fibrosis is a well described salutary effect<br />

of HCV eradication in non-cirrhotic populations; however beneficial<br />

effects in those with cirrhosis require further analysis.<br />

Aims: Evaluate SVR12 rate in a cohort of cirrhotic patients and<br />

quantify change in laboratory parameters with DAA therapy.<br />

Secondary: assess factors that predict magnitude or direction<br />

of change in model for end stage liver disease (MELD) score<br />

in cirrhotics achieving SVR12. Methods: Prospective cohort<br />

study of consecutive mono-infected cirrhotic patients initiated<br />

on DAA regimen: 1) pegylated interferon alfa 2a (IFN) +<br />

Sofosbuvir (SOF) + Ribavirin (RBV); 2) SOF + RBV; 3) SOF<br />

+ Simeprevir (SMV); 4) Ledipasvir (LDV) + SOF. Baseline<br />

demographics tabulated: gender, ethnicity, genotype, DAA<br />

regimen, prior treatment, and prevalence of pre-existing: diabetes,<br />

hypertension, and hyperlipidemia. Laboratory parameters<br />

recorded at start, end, and assessment of SVR12 including:<br />

MELD components/score and platelet count. Changes in lab<br />

values compared between groups that did or did not achieve<br />

SVR12 using student’s T test. Logistic regression incorporated<br />

baseline demographics and calculated adjusted odds ratios<br />

for a decrease in MELD with treatment (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 775A<br />

1146<br />

Efficacy of Ledipasvir plus Sofosobuvir with or without<br />

ribavirin in hepatitis C genotype 1 patients who failed<br />

previous treatment with Simeprevir plus Sofosbuvir<br />

Gabriel R. Gonzales 3 , Stevan A. Gonzalez 1 , Hector E. Nazario 2 ,<br />

Manjushree Gautam 1 , Maisha Barnes 2 , Parvez S. Mantry 2 , Jeffrey<br />

S. Weinstein 2 , Milka Ndungu 2 , Olga Teachenor 2 , Apurva A.<br />

Modi 1 ; 1 Liver Consultants of Texas, Fort Worth, TX; 2 Liver Institute<br />

at Methodist-Dallas, Dallas, TX; 3 Texas College of Osteopathic<br />

Medicine, University of North Texas HSC, Fort Worth, TX<br />

Background: Combination therapy of Simeprevir (SIM), NS3/4<br />

protease inhibitor, with Sofosbuvir (SOF), NS5B polymerase<br />

inhibitor is an FDA approved treatment option for chronic hepatitis<br />

C genotype 1 patients with an overall SVR12 rate of<br />

85-95%. Single tablet fixed dose combination of Ledipasvir<br />

(LDV), NS5A inhibitor, with SOF is also FDA approved for<br />

treatment of hepatitis C genotype 1 with SVR 12 rates of ≥<br />

95%. However, there is no data on the efficacy of re-treatment<br />

with LDV+SOF in patients who failed initial treatment<br />

with SIM+SOF. Aim: To evaluate the efficacy of re-treatment<br />

with LDV+SOF with or without ribavirin (RBV) in genotype 1<br />

patients who have previously failed treatment with SIM+SOF.<br />

Methods: Data from a combined treatment cohort of two large<br />

hepatology referral centers, which included patients who were<br />

previously treated with SIM+SOF with or without RBV for 12<br />

weeks but failed to achieve SVR12 & then undergone re-treatment<br />

with LDV+SOF with or without RBV was analyzed (n=28).<br />

LDV+SOF with or without ribavirin was administered for 12-24<br />

weeks based on the discretion of the treating hepatologist.<br />

Results: Of the 28 patients in this cohort, 21 (75%) were male,<br />

74% were Caucasians & 24 (86%) were genotype 1A. 22<br />

(79%) had cirrhosis of which 54% had decompensated Child<br />

class B/C cirrhosis. Three patients had recurrent hepatitis C<br />

post-liver transplant. 21/28 (75%) achieved undetectable viral<br />

load after 4 weeks of treatment. Of the 3 patients who have<br />

currently completed treatment, all achieved undetectable viral<br />

load at the end of treatment (EOT). Treatment was well tolerated<br />

overall with over 38% reporting no adverse events. In<br />

those reporting adverse events, the most common were fatigue,<br />

headache, insomnia, nausea, & diarrhea. No patients with<br />

decompensated child class B/C cirrhosis had further decompensation<br />

requiring hospitalization or death during treatment.<br />

SVR12 results to follow. Conclusions: Single tablet fixed dose<br />

combination of Ledipasvir plus Sofosbuvir with or without ribavirin<br />

is well tolerated in patients who previously failed treatment<br />

with Simeprevir plus Sofosbuvir including those with decompensated<br />

cirrhosis & recurrent hepatitis C post-liver transplant.<br />

Overall, 78% have achieved negative viral load after 4 weeks<br />

of treatment & all (n=3) who have reached EOT have undetectable<br />

viral load.<br />

The following authors have nothing to disclose: Gabriel R. Gonzales, Maisha<br />

Barnes, Milka Ndungu, Olga Teachenor<br />

1147<br />

Safety And Efficacy Of Two IFN- Free Daas Regimens<br />

In Genotype 4 Chronic HCV Patients: First Real Clinical<br />

Practice Data In Gulf From Qatar HCV Registry<br />

Moutaz F. Derbala 1 , Aliaa Amer 2 , Saad R. Alkaabi 1 , Yasser M.<br />

Kamel 1 , Khaleel H. Sultan 1 , Elham Elsayad 3 , Omer E. Hajelssedig 1 ,<br />

Nazeeh Z. Dweik 1 , Mohammed T. Butt 1 , Mohammed E. Elbadri 1 ;<br />

1 GASTROENTEROLOGY & HEPATOLOGY, Hamad Hospitaland<br />

Weill Cornell Medical College in Qatar,Consultant Gastroenterology<br />

& Hepatology HMC, Doha, Qatar; 2 Laboratory Medicine and<br />

Histopathology, Hematology Section, Doha, Qatar., Hamad Medical<br />

Corporation, Doha, Qatar; 3 Clinical Pharmacology, Hamad<br />

Medical Corporation, Doha, Qatar<br />

Background and Aims: IFN-free regimens of direct acting antivirals<br />

(DAAs): sofosbuvir (SOF), simeprevir (SMV), and daclatasvir<br />

(DCV) with or without ribavirin (RBV) have been used in<br />

Qatar. In real world; little data are currently available on use<br />

of DAAs in genotype 4 HCV patients including patients with<br />

advanced liver diseases and comorbidities in gulf. Our aim is<br />

to assess efficacy and safety of 12 weeks regimen of SOF/SMV<br />

or SOF/DCV versus 24 weeks regimen of SOF/RBV in genotype<br />

4 HCV patients. Methods: 85 chronic HCV genotype 4<br />

patients received either 12 weeks of SOF/SMV or SOF/DCV,<br />

or 24 weeks of SOF plus ribavirin (RBV) therapy. 30(35.3%)<br />

and 10(11.8%) patients were treated with SOF/SMV and<br />

SOF/DCV for 12 weeks respectively, while 45(52.9%) patients<br />

were treated with SOF/RBV for 24 weeks. Demographics, history<br />

of liver disease, laboratory tests, adverse events and virological<br />

data were collected at baseline, throughout treatment<br />

and post-treatment follow-up. Results: 34.3% of patients had<br />

cirrhosis and 58.8% failed on previous PR treatment. 12.9%<br />

of patients underwent liver transplantation. (Table) To date<br />

84/85 patients continued treatment on both regimen, while<br />

1 patient on SOF/DCV discontinued voluntarily. The SVR4 is<br />

96%, with 1 virological breakthrough due to non- compliance<br />

and 2 relapses in experienced patients were reported in the<br />

12 weeks regimen group of SOF/SMV or DCV. No serious<br />

adverse events were reported in any treatment groups, while<br />

26% of the SOF/RBV group reported grade 3 or 4 anemia<br />

which are managed by RBV dose reduction ; with 2 patients<br />

received erythropoietine and 1 patient received blood transfusion.<br />

Conclusion: In Genotype 4 chronic HCV patients, both<br />

12 weeks regimen of SOF in combination with SMV or DCV<br />

and 24 weeks regimen of SOF/RBV achieved high SVR in real<br />

world setting. However, the shorter duration of 12 weeks of<br />

SOF/SMV or DCV and the favorable tolerability and safety<br />

profile make this regimen more recommended in Genotype 4.<br />

Disclosures:<br />

Stevan A. Gonzalez - Consulting: AbbVie; Speaking and Teaching: Gilead,<br />

Salix, AbbVie<br />

Hector E. Nazario - Speaking and Teaching: Merck, Salix<br />

Manjushree Gautam - Speaking and Teaching: Gilead<br />

Parvez S. Mantry - Consulting: Salix, Gilead, Janssen, Abbvie, BMS; Grant/<br />

Research Support: Salix, Merck, Gilead, Boehringer-Ingelheim, Mass Biologics,<br />

Vital Therapies, Santaris, mass biologics, Bristol-Myers Squibb, Abbive, Bayer-Onyx,<br />

Shinogi, Tacere, Intercept; Speaking and Teaching: Gilead, Janssen,<br />

Salix<br />

Jeffrey S. Weinstein - Speaking and Teaching: Merck<br />

Apurva A. Modi - Speaking and Teaching: Salix, Merck, Gilead, Abbvie, Janssen<br />

Disclosures:<br />

The following authors have nothing to disclose: Moutaz F. Derbala, Aliaa Amer,<br />

Saad R. Alkaabi, Yasser M. Kamel, Khaleel H. Sultan, Elham Elsayad, Omer<br />

E. Hajelssedig, Nazeeh Z. Dweik, Mohammed T. Butt, Mohammed E. Elbadri


776A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1148<br />

Preliminary Experience Of Direct Acting Antiviral Therapy<br />

In Hepatitis C Infected Kidney Transplant Recipients,<br />

Who Received Grafts From Hepatitis C Positive Or Negative<br />

Donors<br />

Kalyan R. Bhamidimarri, David Roth, Giselle Guerra, Cynthia Levy,<br />

Paul Martin; University of Miami-Miller School of Medicine, Miami,<br />

FL<br />

Background: Highly effective direct acting antiviral therapy<br />

(DAA) is now feasible post kidney transplantation (KT) and has<br />

also facilitated a new strategy of abbreviating time in waitlisted<br />

patients who could accept Hepatitis C (HCV) positive<br />

grafts. However there are no reports of treatment experience<br />

in this group to date. Aim: We describe our preliminary experience<br />

of HCV treatment with DAA’s in post KT patients, a<br />

majority of whom received a graft from HCV+ donor. Methods:<br />

Fourteen patients with chronic HCV, genotype 1, 71% males,<br />

57% African Americans, mean age of 54 years underwent<br />

deceased donor kidney transplantation (71% HCV+ graft).<br />

Baseline features include 86% with genotype 1a, 50% with<br />

fibrosis level greater than F2, mean HCV viral load of 3.4<br />

Million IU, mean hemoglobin of 12gm/ dL and mean GFR of<br />

60mL/ minute. Open label treatment with Sofosbuvir / Ledipasvir<br />

was initiated in 13 patients and one patient received<br />

Sofosbuvir plus Simeprevir. Only 65% of the patients had ribavirin<br />

approved in their regimen and the dosage was titrated<br />

according to their tolerance. Treatment was initiated within<br />

a mean period of 5months since the KT (range 1-10months)<br />

and immunosuppression was closely monitored. Results: None<br />

of the patients developed fibrosing cholestatic hepatitis post<br />

KT, despite receiving a graft from HCV+ donor. All patients<br />

are at various stages of completion of their respective treatment<br />

duration. Five patients reached end of treatment (EOT), 2<br />

patients reached post-treatment week 4 and 1patient reached<br />

post treatment week 12. According to per protocol analysis,<br />

the EOT, SVR4 and SVR12 rates are 100%. Anemia was frequent<br />

in the patients taking ribavirin which warranted frequent<br />

dose titrations. Conclusions: Our preliminary experience shows<br />

an evolving successful strategy of transplanting kidneys from<br />

HCV+ donor in HCV infected recipients and the feasibility of<br />

HCV treatment with newer DAA’s. As per protocol analysis,<br />

all the patients had excellent virologic response and full data<br />

awaited by AASLD Liver Meeting.<br />

HCV Treatment in KT recipients<br />

Disclosures:<br />

Kalyan R. Bhamidimarri - Advisory Committees or Review Panels: Gilead, Abb-<br />

Vie, Janssen; Grant/Research Support: Bristol Squibb Myers, Biotest, Synageva,<br />

Salix, Vital Therapies, Ocera Inc<br />

David Roth - Advisory Committees or Review Panels: Merck, Sharp and Dome;<br />

Consulting: Merck, Sharp and Dome<br />

Paul Martin - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Merck, Gilead, Janssen, Abbvie<br />

The following authors have nothing to disclose: Giselle Guerra, Cynthia Levy<br />

1149<br />

IFNL4 ss469415590 Polymorphism Contributes to Treatment<br />

Decisions in Patients with Chronic Hepatitis C virus<br />

Genotype 1b but not 2a Infection<br />

Ruihong Wu 1 , Xiumei Chi 1 , Haibo Sun 1 , Juan Lv 1 , Xiaomei<br />

Wang 1 , Xiuzhu Gao 1 , Ge Yu 1 , Fei Kong 1 , Peng Zhang 2 , Mo-li<br />

Wang 3 , Dongsheng Wang 4 , Tao Jiang 4 , Hongqin Xu 1 , Rui Hua 1 ,<br />

Gerald Y. Minuk 5 , Jing Jiang 6 , Bing Sun 7 , Jin Zhong 7 , Yu Pan 1 ,<br />

Junqi Niu 1 ; 1 Department of Hepatology, First hospital of JiLin University,<br />

ChangChun, China; 2 Department of Infectious Diseases,<br />

First hospital of Jilin University, Changchun, China; 3 Department of<br />

Infectious Diseases, Fourth Hospital of Jilin University, Changchun,<br />

China; 4 Hospital of HepatologyBiliary of Jilin Province, Changchun,<br />

China; 5 Section of Hepatology, University of Manitoba, Winnipeg,<br />

MB, Canada; 6 Department of Clinical Epidemiology, First<br />

Hospital of Jilin University, Changchun, China; 7 Institute Pasteur<br />

of Shanghai, Shanghai Institutes for Biological Sciences, Chinese<br />

Academy of Sciences, Shanghai, China<br />

Background & Aim: Recently, the dinucleotide variant<br />

ss469415590(TT/ΔG) in the novel gene IFNL4 was identified<br />

as a stronger predictor of hepatitis c virus (HCV) clearance in<br />

individuals of African ancestry compared with rs12979860.<br />

The aim of this study was to determine whether this variant<br />

contributes to treatment decisions in a Chinese population,<br />

predominantly infected with HCV genotype (GT) 1b and 2a.<br />

Methods: Four hundred and forty seven chronic hepatitis c<br />

(CHC) patients (including 328 interferon alpha-2b and ribavirin<br />

treated), 129 individuals who had undergone spontaneous<br />

HCV clearance (SHC) and 169 healthy controls were<br />

retrospectively investigated. ss469415590 genotyping was<br />

performed using a Mass spectra method (SEQUENOM).<br />

Results: A higher proportion of SHC individuals carried TT/TT<br />

genotype compared to CHC patients (95.3% vs. 88.8%, P =<br />

0.027). ss469415590 variant was independently associated<br />

with sustained virological response (SVR, OR = 3.247, 95%<br />

CI = 1.038-10.159, P = 0.043) and on-treatment virological<br />

responses including rapid virological response, complete early<br />

virological response, early virological response and end of<br />

treatment virological response with a minimal OR of 3.73.<br />

Patients with high viral load (≥4×10 5 IU/mL) and the ΔG allele<br />

had a higher risk of failing to achieve SVR compared to TT/<br />

TT genotype (92.9% vs. 64%, P = 0.034). In GT-2a patients,<br />

no significant association of ss154949590 variant with virological<br />

response was identified. In addition, we found that<br />

ss154949590 was in complete linkage disequilibrium with<br />

rs12979860. Conclusions: IFNL4 ss154949590 TT/TT genotype<br />

favours spontaneous clearance of HCV, and treatment<br />

induced clearance in genotype 1b but not 2a infected patients.<br />

IFNL4 ss469415590 (or rs12979860) genotyping should<br />

be considered for patients with genotype 1b and high viral<br />

load when making a choice between standard dual therapy<br />

and IFN-free direct-acting antiviral regimen. If with the ΔG for<br />

ss469415590 (or T for rs12979860), IFN-free DAA regimens<br />

may be better options.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ruihong Wu, Xiumei Chi, Haibo<br />

Sun, Juan Lv, Xiaomei Wang, Xiuzhu Gao, Ge Yu, Fei Kong, Peng Zhang, Mo-li<br />

Wang, Dongsheng Wang, Tao Jiang, Hongqin Xu, Rui Hua, Gerald Y. Minuk,<br />

Jing Jiang, Bing Sun, Jin Zhong, Yu Pan, Junqi Niu


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 777A<br />

1150<br />

Combination therapies with daclatasvir and asunaprevir<br />

on NS3-D168 mutated hepatitis C virus in human<br />

hepatocyte chimeric mice<br />

Hiromi Kan 1 , Michio Imamura 1 , Nobuhiko Hiraga 1 , Eisuke<br />

Miyaki 1 , Takuro Uchida 1 , Masataka Tsuge 1 , Hiromi Abe 1 , Nelson<br />

Hayes 1 , Hiroshi Aikata 1 , Chise Tateno 2 , Kazuaki Chayama 1 ;<br />

1 Department of Gastroenterology and Metabolism, Applied Life<br />

Sciences, Institute of Biomedical & Health Sciences, Hiroshima<br />

University, Hiroshima, Japan; 2 PhoenixBio Co., Ltd., Higashihiroshima,<br />

Japan<br />

Background and Aim: The frequency of emergent drug-resistant<br />

strains of hepatitis C virus (HCV) in patients who failed to<br />

respond to simeprevir plus peginterferon (PEG-IFN) and ribavirin<br />

(RBV) combination therapy decreased after cessation of<br />

the treatment. However, it is not clear whether or not the HCV<br />

NS3-D168 mutation affects the outcome of an IFN-free combination<br />

therapy including NS5A inhibitor, daclatasvir and NS3<br />

protease inhibitor, asunaprevir. In this study, we investigated<br />

the relationship between the effect of daclatasvir plus asunaprevir<br />

treatment and the frequencies of drug-resistant NS3-D168<br />

variants using HCV-infected mice. Methods: Injection with wildtype<br />

and NS3 D168 mutated genotype 1b HCV mixed serum<br />

to human hepatocyte chimeric uPA-SCID mice enabled us to<br />

develop mice with various frequencies of NS3-D168 mutation.<br />

These mice were treated with 40 mg/kg of asunaprevir alone or<br />

in combination with 10 mg/kg daclatasvir for four weeks (provided<br />

by Bristol-Meyers Squibb Research and Development).<br />

Frequencies of NS3-D168 mutation at baseline were analyzed<br />

by ultra-deep sequencing. Some mice infected with NS3-D168<br />

substitutions were administered with either intramuscular injection<br />

of 30 μg/kg of PEG-IFN-alfa twice a week or 200 mg/kg<br />

of telaprevir (provided from Mitsubishi Tanabe Pharma) orally<br />

twice per day for four weeks. Results: Mice with high NS3-<br />

D168 variant frequencies showed a low susceptibility to asunaprevir<br />

mono-therapy and failed to respond to daclatasvir plus<br />

asunaprevir therapy. In contrast, mice with a low frequency<br />

(less than approximately 14%) of NS3-D168 variants showed<br />

a similar susceptibility to asunaprevir mono-therapy with wildtype<br />

HCV-infected mice and achieved viral eradication with<br />

four weeks of daclatasvir plus asunaprevir therapy. Although<br />

telaprevir or PEG-IFN treatment resulted in a reduction of serum<br />

HCV RNA levels, no significant decrease in the frequency of<br />

NS3-D168 substitutions was achieved. Conclusion: Daclatasvir<br />

and asunaprevir treatment could eliminate NS3-D168-mutated<br />

HCV if the frequency at baseline was low. Using either telaprevir<br />

or IFN might cause no significant reductions in NS3-<br />

D168 mutation. It is necessary to confirm that the frequency of<br />

NS3-D168 variants has decreased sufficiently before adopting<br />

daclatasvir plus asunaprevir therapy in patients with simeprevir<br />

plus PEG-IFN/RBV treatment failure.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Asuka, Bayer, Daiichi Sankyo, Dainippon Sumitomo, Eisai, Janssen,<br />

Kowa, Mitsubishi Tanabe, MSD, Eli Lily, Nippon Kayaku, Nippon Shinyaku,<br />

Otsuka, Roche, Takeda, Toray, Torii, Tsumura, Zeria; Speaking and Teaching:<br />

Eisai, Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Asuka, Bayer, BMS,<br />

Chugai, Daiichi Sankyo, Dainippon Sumitomo, J&J, Jimro, Miyarisan, MSD,<br />

Nihon Kayaku, Olympus<br />

The following authors have nothing to disclose: Hiromi Kan, Michio Imamura,<br />

Nobuhiko Hiraga, Eisuke Miyaki, Takuro Uchida, Masataka Tsuge, Hiromi Abe,<br />

Nelson Hayes, Hiroshi Aikata, Chise Tateno<br />

1151<br />

HepCure: An Innovative web-based toolkit to train a<br />

new cohort of hepatitis C providers and increase patient<br />

engagement<br />

Ponni V. Perumalswami 1 , Korin Parrella 2 , Jason Rogers 3 , Rahul<br />

Patel 4 , Brooke Wyatt 1 , Kian Bichoupan 1 , Andrea D. Branch 1 , Brian<br />

Kim 1 , Anna Patel 1 , Joseph Lawler 1 , Alicia Stivala 2 , Alyson Harty 1 ,<br />

Donald Gardenier 2 , Michel Ng 1 , Catherine Amory 2 , Aaron M.<br />

Vanderhoff 1 , Douglas Dieterich 1 , Ashish Atreja 3 , Jeffrey J. Weiss 2 ;<br />

1 Liver Diseases, Icahn School of Medicine at Mount Sinai, New<br />

York, NY; 2 Medicine, Icahn School of Medicine at Mount Sinai,<br />

New York, NY; 3 Medicine- Division of Gastroenterology, Icahn<br />

School of Medicine at Mount Sinai, New York, NY; 4 Parsus Solutions,<br />

LLC, Scottsdale, AZ<br />

Background and Aims: There is a great need to expand the<br />

number of health care professionals treating hepatitis C and<br />

to optimize patient engagement. The aim of this study is to<br />

develop and operationalize HepCure (http://hepcure.org),<br />

a web-based application (app) with three components: An<br />

open access toolkit (a dashboard) that enhances providers’<br />

ability to deliver guideline-based HCV care; a patient app that<br />

provides education, medication reminders, and a platform<br />

for tracking adherence and symptoms; and a tele-education<br />

platform for medical providers. Methods: The HepCure app<br />

is a collaboration among an academic, tertiary care medical<br />

center, the New York State (NYS) Department of Health,<br />

and NYS community health centers. A working group of multidisciplinary<br />

HCV experts (Hepatology, Internal Medicine,<br />

Psychology, Social Work, and Public Health) was created to<br />

develop content for the HepCure dashboard. Input from this<br />

group was used to develop wireframes. Focus groups were<br />

held with providers and patients to refine wireframes. National<br />

guidelines (AASLD/IDSA) were used to identify key patient<br />

characteristics to determine treatment recommendations to<br />

build decision support algorithms. The app has the capability<br />

to de-identify patient data to directly discuss cases selected<br />

by providers at weekly tele-education sessions with experts.<br />

Results: The HepCure platform captures United States Centers<br />

for Medicaid and Medicare and NYS HCV quality indicators<br />

for population health management, provides decision support<br />

algorithms for providers that can be updated in real time with<br />

newly approved HCV treatment regimens and incorporates<br />

a tele-education platform with access to experts. The Hep-<br />

Cure Provider Dashboard (http://providers.hepcure.org) was<br />

launched in November 2014. Weekly tele-education sessions<br />

have been conducted since February 2015 (n=17) and each<br />

session has been archived with open access. These sessions<br />

have been attended by an average of 13 ± 7 attendees a<br />

week and archived sessions have been viewed an average of<br />

8 ± 3 times. A patient mobile app which can be linked to the<br />

provider dashboard was launched in June 2015 to increase<br />

patient engagement. Conclusions: Providers are interested in<br />

web-based tools to help them master HCV treatment and stay<br />

current with ongoing changes in treatment paradigms. Patients<br />

with HCV desire web-based tools to help them learn more<br />

about and navigate HCV treatment. The HepCure platform is<br />

an innovative model that can be adapted globally in line with<br />

local HCV treatment guidelines to train a new generation of<br />

HCV providers, collect real world data, and foster comparative<br />

effectiveness research.<br />

Disclosures:<br />

Kian Bichoupan - Consulting: Janssen Pharmaceuticals, Gilead Sciences<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

Alyson Harty - Advisory Committees or Review Panels: Gilead; Consulting: Gilead,<br />

Jannsen, Acaria Pharmacy, Abbvie


778A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Michel Ng - Advisory Committees or Review Panels: abbvie; Speaking and<br />

Teaching: abbvie<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Jeffrey J. Weiss - Consulting: AbbVie Inc.<br />

The following authors have nothing to disclose: Ponni V. Perumalswami, Korin<br />

Parrella, Jason Rogers, Rahul Patel, Brooke Wyatt, Brian Kim, Anna Patel, Joseph<br />

Lawler, Alicia Stivala, Donald Gardenier, Catherine Amory, Aaron M. Vanderhoff,<br />

Ashish Atreja<br />

1152<br />

Patient Expectations of the Impact on Quality of Life of<br />

Hepatitis C Treatment<br />

Shelagh M. Szabo 2 , David R. Walker 1 , Timothy R. Juday 1 ,<br />

Suzanne Lane 2 , Adrian Levy 2 , Sammy Saab 3 ; 1 HEOR, Abbvie,<br />

North Chicago, IL; 2 Icon Health Economics, Vancouver, BC, Canada;<br />

3 David Geffen School of Medicine, UCLA, Los Angeles, CA<br />

BACKGROUND The Investigative Multicenter Prospective Observational<br />

Study (IMPROOV) is a study of humanistic and economic<br />

outcomes among patients initiating direct acting antiviral<br />

treatments for hepatitis C virus (HCV). The objective of this analysis<br />

is to describe patient-reported expectations of the impact<br />

on quality of life of newly-initiated HCV treatment. METHODS<br />

All consecutive HCV patients attending clinics at ten sites in<br />

the United States were screened to identify the cohort initiating<br />

treatment for HCV. Data were collected using patient-completed<br />

surveys at up to five time points over the treatment<br />

course, supplemented by a medical chart review at treatment<br />

initiation and completion. Data collected included demographic,<br />

clinical, economic, and humanistic. Data on patients’<br />

expectations of the impact on quality of life of newly-initiated<br />

treatment were collected; multiple responses were allowed per<br />

patient. The specific question asked at baseline was: “Do you<br />

expect the following to occur, when you start HCV treatment”;<br />

and a series of potential patient experiences were listed to<br />

which patients responded with the expected impact. Response<br />

options included: none, a little, moderate or severe. The sum<br />

of patients expecting moderate/severe outcomes was calculated.<br />

First patient enrolled 1/2013 and last patient enrolled<br />

9/2014. RESULTS 143 patients, from 8 clinics, enrolled out<br />

of 787 that were screened across 10 clinics of which 137<br />

were from community clinics and 6 from academic/hospitals.<br />

Mean age was 56 years, 52% male, 60% white, 20% hispanic<br />

and 15% black. 75% graduated from high school and<br />

33% were employed. 100 patients were on a sofosbuvir and/<br />

or simeprevir-based regimen, and 43 patients were on other<br />

treatments. 74 were on an interferon-free (IFF) regimen. The<br />

proportion of patients expecting moderate/severe impacts of<br />

newly initiated treatment for those on IFF and interferon-containing<br />

treatment, respectively, were: treatment-related side-effects<br />

(31.5%, 50.7%;p=0.02), missed work (7.2%,17.9%;p=0.06<br />

), decreased productivity at work (9.9%, 25.0%;p=0.03), limited<br />

social activities (25.0%,40.6%;p=0.05), increased reliance<br />

on family/friends (18.8%,36.2%;p=0.02), impact on<br />

your quality of life (21.4%,49.3%;p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 779A<br />

Concluding, the combined use of IFNL4 polymorphisms and<br />

ALT reduction at 4 week of treatment is able to optimize candidates’<br />

selection for antiviral therapy with PEG-IFN and RBV in<br />

patients with early liver fibrosis, discriminating those that could<br />

still benefit from dual therapy from the ones that will need the<br />

new regimens.<br />

Disclosures:<br />

Luchino Chessa - Board Membership: Abbvie; Speaking and Teaching: BMS,<br />

Jannsen<br />

The following authors have nothing to disclose: Francesco Figorilli, Simona<br />

Onali, Stefania Catone, Claudio Argentini, Stefania Casu, Cinzia Balestrieri,<br />

Maria Conti, Giancarlo Serra, Michele Casale, Maria Cristina Pasetto, Laura<br />

Matta, Lucia Barca, Rosetta Scioscia, Irene Canini, Maria Giovanna Quaranta,<br />

Domenico Genovese, Stefano Vella<br />

1154<br />

Real Life Treatment Outcomes with 8 week Course<br />

Treatment for Hepatitis C without Cirrhosis Confirmed<br />

by Transient Elastography<br />

Vanessa Marshall 2,3 , Kelsey Rife 1 , Amy Hirsch 1,3 , Marina G. Silveira<br />

2,3 , Anita Compan 2 , Anita L. Moreland 2 , Yngve Falck-Ytter 2,3 ;<br />

1 Pharmacy, Louis Stokes Cleveland VA Medical Center, Cleveland,<br />

OH; 2 Gastroenterology Section, Louis Stokes Cleveland VA Medical<br />

Center, Cleveland, OH; 3 School of Medicine, Case Western<br />

Reserve University, Cleveland, OH<br />

Background: Short course treatments for Hepatitis C (HCV)<br />

are less expensive for payers and easier for patients. Data is<br />

needed to confirm that real world sustained virologic response<br />

(SVR) rates for short courses of treatment in patients deemed<br />

non-cirrhotic based on transient elastography (TE, FibroScan)<br />

results are comparable to SVR rates seen in registrational trials.<br />

It is also not known if cutoffs from non-invasive markers of<br />

liver fibrosis assessment (Fibrosis-4 score (FIB-4) or aspartate<br />

transaminase to platelet ratio index (APRI) score) are sufficient<br />

to select for short course treatment. Objective: This study was<br />

designed to examine (1) if a TE score < 12.5 kPa is sufficient<br />

criteria to rule out cirrhosis in candidates eligible for 8 weeks<br />

of ledipasvir/sofosbuvir; and (2) evaluate if the clinical cutoffs<br />

for FIB-4 (< 3.25) and APRI score (< 1) are sufficient criteria<br />

to shorten the course of treatment. Methods: Veteran genotype<br />

1 (GT1), treatment naïve and previous interferon/ribavirin<br />

relapsers with pre-treatment TE scores of < 12.5 kPa and HCV<br />

viral loads < 6 million IU/mL were treated with 8 weeks of<br />

ledipasvir/sofosbuvir. Pre-treatment FIB-4 and APRI scores were<br />

calculated. Outcomes were assessed at 4 (SVR4), 12 (SVR12),<br />

and 24 (SVR24) weeks post-treatment. Results: To date, 52 Veterans<br />

were initiated on an 8 week course of ledipasvir/sofosbuvir.<br />

94% were male with a mean age of 61 (range 32-75). 6<br />

patients were previous treatment relapsers with interferon/ribavirin.<br />

31 patients (60%) had GT1a, 18 (35%) had GT1b and<br />

3 (5%) had GT1 (no subtype) or GT 1a/1b. 41 patients (78%)<br />

had baseline TE scores < 9.5 kPa. To date, 17/18 (94%)<br />

patients achieved SVR4 and 1 patient relapsed. 5/17 (29%)<br />

patients who achieved SVR4 had an APRI ≥ 1 and/or FIB-4 ≥<br />

3.25. The patient who relapsed after 8 weeks of ledipasvir/<br />

sofosbuvir was treatment naïve, GT1a, had a TE score of 11.2<br />

kPa, and APRI > 1 but FIB-4 < 3.25. Viral loads were undetectable<br />

at week 4 and end of treatment and the patient reported<br />

no missed doses of ledipasvir/sofosbuvir. Final SVR12 results<br />

will be presented for all 52 patients. Conclusions: A treatment<br />

strategy using TE to qualify patients for an 8 week course of<br />

ledipasvir/sofosbuvir resulted in similar SVR rates to those seen<br />

in registrational trials. These results suggest a TE score < 12.5<br />

kPa is sufficient criteria to shorten the course of HCV treatment<br />

to 8 weeks with > 90% SVR rate. Due to lower diagnostic accuracy<br />

of APRI and FIB-4 scores, their application at suggested<br />

cutoffs may result in an additional 4 weeks of ledipasvir/sofosbuvir,<br />

resulting in unnecessary resource expenditure.<br />

Disclosures:<br />

The following authors have nothing to disclose: Vanessa Marshall, Kelsey Rife,<br />

Amy Hirsch, Marina G. Silveira, Anita Compan, Anita L. Moreland, Yngve<br />

Falck-Ytter<br />

1155<br />

Effectiveness of Treatment of Chronic Hepatitis C in<br />

Patients Being Treated for Hepatocellular Cancer: A Single<br />

Center Experience<br />

Shruti Mony 2 , Shirin Nafisi 1 , Justin A. Reynolds 1 , Ann Moore 1 ,<br />

Ashley Hepner 1 , Karen Arendt 1 , Yvette Cummings 1 , Robert Gish 1 ,<br />

Richard A. Manch 1 , Anita Kohli 1 ; 1 Hepatology, St. Joseph’s Hospital<br />

and Medical Center, Phoenix, AZ; 2 Internal Medicine, St.<br />

Joseph’s Hospital and Medical Center, Phoenix, AZ<br />

Treatment of chronic hepatitis C (CHC) with directly acting<br />

antivirals has improved SVR rates for patients with advanced<br />

liver disease. Patients with hepatocellular cancer (HCC) remain<br />

a group with a poor prognosis. Effectiveness of CHC treatment<br />

in this group, often receiving other anti-cancer agents or therapies,<br />

is unknown. METHODS: Patients treated for CHC during<br />

or after HCC treatment at a community based academic medical<br />

center, between 1/12- 5/15 were retrospectively identified<br />

and included. Demographics, CHC and HCC treatment,<br />

virologic data, treatment outcomes, HCC stage, and survival<br />

were collected. RESULTS: 35 patients with HCC treated for<br />

CHC concomitantly or after HCC therapy were identified. 62%<br />

were men, 88% Caucasian, 17% Hispanic and 54% treatment<br />

experienced. 22 patients were treated for HCC and CHC<br />

concomitantly. In patients with GT1, 12 received ledipasvir<br />

(LDV)/sofosbuvir (SOF) for 24 weeks: 8 (single HCC) received<br />

Y90/microwave ablation/sorafenib, Y90, TACE, RFA/TACE/<br />

Y90, microwave ablation, TACE/quadsphere or await liver<br />

transplant; 3 (multifocal HCC) received TACE/sorafenib,<br />

TACE or TACE/RFA/radiation for prostate adenocarcinoma;<br />

1 (metastatic HCC)received Y90. 1 patient (metastatic HCC)<br />

received LDV/SOF for 12 weeks with embolization/ sorafenib.<br />

4 patients received SOF/simeprevir (SMV) for 12 weeks with<br />

TACE, TACE/ Y90/transplant (single HCC), cyberknife/<br />

sorafenib or TACE/sorafenib (metastatic HCC). 1 patient (GT2<br />

& metastatic HCC) received SOF/ribavirin (RBV) for 20 weeks<br />

& Y90/TACE/sorafenib. 4 patients (GT3) received SOF/RBV<br />

for 24 weeks with Y90/TACE/transplant or microablation<br />

(single HCC) or TACE/sorafenib or TACE(multifocal HCC). In<br />

patients who reached SVR 12, 50% (n=3) achieved SVR 12,<br />

33% (n=2) relapsed & 17% (n=1) died. SVR12 is pending<br />

for 16 patients. 13 patients were treated for CHC after HCC<br />

therapy. 3 (GT1,single HCC) received Y90/microablation/<br />

RFA or Y90 followed by LDV/SOF for 12 weeks. 6 (GT1,<br />

single HCC) received TACE, RFA, microablation or TACE/Y90<br />

followed by SOF/SIM for 12 weeks. 2 (GT2) received TACE<br />

for single & multiple HCC followed by SOF/RBV for 12 weeks.<br />

1 patient (GT3) received Y90 for a single HCC followed by<br />

SOF/RBV for 24 weeks. In patients who reached SVR 12,<br />

70% (n=7) achieved SVR 12 & 30% (n=3) relapsed. SVR12 is<br />

pending for 3 patients. Full results of SVR12 will be presented.<br />

CONCLUSION: Early results suggest that the treatment of CHC<br />

in patients with HCC, particularly those undergoing concomitant<br />

HCC therapy may be lower than other advanced liver<br />

disease patients. Immune modulating agents and liver targeted<br />

therapies may reduce the effectiveness of HCV interferon free<br />

therapies.


780A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Robert Gish - Advisory Committees or Review Panels: Gilead, AbbVie, Arrowhead;<br />

Consulting: Eiger, Isis, Genentech; Speaking and Teaching: Gilead, Abb-<br />

Vie; Stock Shareholder: Arrowhead<br />

Richard A. Manch - Speaking and Teaching: Gilead, AbbVie, Bayer, Salix, BMS<br />

The following authors have nothing to disclose: Shruti Mony, Shirin Nafisi, Justin<br />

A. Reynolds, Ann Moore, Ashley Hepner, Karen Arendt, Yvette Cummings, Anita<br />

Kohli<br />

1156<br />

Sofosbuvir-based treatments for patients with hepatitis<br />

C virus (HCV) mono-infection and human immunodeficiency<br />

virus (HIV)-HCV co-infection with genotype 1 and<br />

4 in clinical practice – Results from the GErman hepatitis<br />

C COhort (GECCO)<br />

Stefan Mauss 1 , Knud Schewe 2 , Jürgen K. Rockstroh 3 , Dietrich<br />

Hueppe 4 , Axel Baumgarten 5 , Guenther Schmutz 1 , Karl-Georg<br />

Simon 6 , Thomas Lutz 8 , Heiner W. Busch 7 , Patrick Ingiliz 5 , Stefan<br />

Christensen 7 ; 1 Center for HIV and Hepatogastroenterology,<br />

Duesseldorf, Germany; 2 ICH, Hamburg, Germany; 3 Department I<br />

University Hospital Bonn, Bonn, Germany; 4 Center for Gastroenerology,<br />

Herne, Germany; 5 MIB, Berlin, Germany; 6 Practice for<br />

Gastroenterology, Leverkusen, Germany; 7 CIM, Muenster, Germany;<br />

8 Infektiologikum, Frankfurt, Germany<br />

Introduction: The first direct acting antivirals (DAA) were<br />

approved in Europe in 2014 based on limited study data.<br />

In particular, HIV-HCV co-infected patients and pre-treated<br />

patients had not been systematically studied. Here, we present<br />

real-life data on efficacy and safety of sofosbuvir (SOF)-based<br />

therapies from Germany. Methods: In this multicenter cohort,<br />

patients who were started on sofosbuvir-based treatments were<br />

documented consecutively. For the current analysis, patients<br />

included had HCV genotype 1 and 4 treated with: SOF/<br />

PegIFN/RBV (ribavirin) 12 weeks, SOF/DCV (daclatasvir)<br />

12/24 weeks, SOF/SMV (simeprevir) 12 weeks and SOF/LDV<br />

(ledipasvir) 8/12 weeks. Patients within the GECCO cohort<br />

are part of the prospective German hepatitis C registry. Results:<br />

Overall, 836 patients on SOF-containing regimens have been<br />

enrolled so far. Of those, 629 (75%) are HCV-monoinfected<br />

and 207 (25%) HIV-HCV co-infected. 536 patients (60%) were<br />

male, the median age was 52 years. 421 patients (52%) had<br />

been pretreated, mainly with PegIFN and RBV. Liver cirrhosis<br />

was reported in 217 patients (27%), in the majority of cases<br />

(71%) defined by a transient elastography value (Fibroscan®)<br />

>12.5kPa. 17/217 (8%) cirrhotic patients had Child-Pugh<br />

stage B or C. HCV genotype 1 was present in 584 patients<br />

(69%), and genotype 4 in 48 patients (6%). Most HIV-HCV<br />

co-infected patients (205/207) were on antiretroviral therapy<br />

and had HIV-RNA 2) were less likely to achieve<br />

SVR12 compared to non-cirrhotic patients (APRI


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 781A<br />

sion] The combination therapy of DCV and ASV for patients<br />

with HCV genotype 1 had highly effectiveness. The efficacy<br />

and the safety of this therapy for old patients were equal to the<br />

therapy for young patients. NS5B polymerase inhibitor sofosbuvir<br />

is not available for patients with severe renal dysfunction.<br />

We may positively select the combination therapy DCV and<br />

ASV for elderly patients, because elderly patients tended to<br />

have decreased renal function.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kenta Motomura, Masayoshi<br />

Yada, Taiji Mutsuki, Masayuki Miyazaki, Takeshi Senju, Motoyuki Kohjima,<br />

Makoto Nakamuta, Akihide Masumoto<br />

1158<br />

Sofosbuvir-based antiviral therapy in HCV patients with<br />

severe renal failure<br />

Jérôme Dumortier 1 , François Bailly 2 , Georges-Philippe Pageaux 3 ,<br />

Anaïs Vallet-Pichard 4 , Sylvie Radenne 2 , François Habersetzer 5 ,<br />

Marie-Claude Gagnieu 1 , Jean Didier Grange 6 , Anne Minello 7 ,<br />

Nassim Kamar 8 , Laurent Alric 9 , Vincent Leroy 10 ; 1 Hôpital Edouard<br />

Herriot, Lyon, France; 2 Croix-Rousse Hospital, Lyon, France; 3 Saint<br />

Eloi University Hospital, Montpellier, France; 4 Cochin Hospital,<br />

Paris, France; 5 University Hospital of Strasbourg, Strasbourg,<br />

France; 6 Tenon Hospital, Paris, France; 7 Dijon University Hospital,<br />

Dijon, France; 8 Rangueil University Hospital, Toulouse, France;<br />

9 Purpan University Hospital, Toulouse, France; 10 CHU A Michallon,<br />

Grenoble, France<br />

Background: Chronic hepatitis C virus (HCV) infection is the<br />

most common chronic liver disease in patients with end stage<br />

renal disease (ESRD). During the last year, second generation<br />

direct acting antivirals have been a revolution for the treatment<br />

of hepatitis C, and sofosbuvir (SOF) is the cornerstone<br />

of modern treatment. Since SOF is eliminated through kidney,<br />

the aim of this multicentre retrospective study was to assess its<br />

antiviral efficacy and tolerability in HCV infected patients with<br />

severe renal failure (including haemodialysed). Patients: Fifty<br />

patients (36 male, mean age 60.5±7.5 years) with chronic<br />

HCV-infection (7 GT1a, 21 GT1b; 6 GT2; 5 GT3; 9 GT4; 2<br />

GT5; cirrhosis: 27/54%) with severe renal failure, i.e. GFR<br />

< 35 mL/min, including 35 haemodialysed, were included<br />

in this study. There were 17 patients with history of kidney<br />

transplantation, 11 patients with history of liver transplantation,<br />

and 27 patients were on waiting list for kidney transplantation.<br />

Fourteen patients were naïve of antiviral treatment. Antiviral<br />

treatment consisted in SOF/ribavirin (RBV) (n=7), SOF/RBV/<br />

PEG-IFN (n=2), SOF/daclatasvir (DCV) ±RBV (n= 30), or SOF/<br />

simeprevir (SMV) ±RBV (n= 11). Treatment duration was 12 or<br />

24 weeks according to regimen. Reduced dose of SOF (400<br />

mg 3 times a week or 400 mg every other day) was given<br />

in all haemodialysed patients. Results. On-treatment response<br />

(rate of undetectable RNA) was as follows: week 4 : 36/44<br />

(82%), week 8 : 38/40 (95%), week 12: 50/50 (100%),<br />

week 24 : 10/10 (100%). Sustained virological response rate<br />

was 27/29 (93%) at 4 weeks (SVR4) and 24/26 (92%) at 12<br />

weeks (SVR12). The 2 patients with virological relapse were<br />

both G3, with severe fibrosis, who were treated for 12 weeks<br />

without RBV. Dose of RBV (n=13) ranged from 200 mg 3 times<br />

a week to 600 mg/day. The mean baseline hemoglobin level<br />

was 11.6±1.4 g/dL and 19 patients were on Epoetin before<br />

starting antiviral therapy. At the end of treatment, the mean<br />

hemoglobin level was 11.7±1.8 g/dL and 19 patients were on<br />

Epoetin. Hemoglobin level was not different in patients receiving<br />

RBV. No patient experienced severe anemia (hemoglobin<br />

level < 8 g/dL) during treatment. In non-haemodialysed<br />

patients, GFR was not significantly modified during treatment<br />

(29.6±6.2 mL/min at base line vs. 27.9±6.5 mL/min at the<br />

end of treatment). No patient had treatment discontinuation<br />

and antiviral therapy was well tolerated. Conclusion. Our<br />

results strongly suggest that SOF-based antiviral therapy, with<br />

reduced dose of SOF, is safe and effective for the treatment of<br />

HCV patients with ESRD, including haemodialysed, similarly<br />

to HCV patients without ESRD. Final SVR12 will be presented.<br />

Disclosures:<br />

Jérôme Dumortier - Board Membership: Novartis, Astellas, Roche; Consulting:<br />

Novartis; Grant/Research Support: Novartis, Astellas, Roche, MSD, GSK<br />

François Bailly - Board Membership: ABBVIE, MSD, BMS, GILEAD; Speaking and<br />

Teaching: JANSSEN<br />

Georges-Philippe Pageaux - Advisory Committees or Review Panels: Roche,<br />

Roche, Roche, Roche; Board Membership: Astellas, Astellas, Astellas, Astellas<br />

Anaïs Vallet-Pichard - Independent Contractor: Schering Plough, Gilead, BMS,<br />

Roche<br />

François Habersetzer - Advisory Committees or Review Panels: Cytheris; Board<br />

Membership: Gilead, BMS; Speaking and Teaching: Gilead, Janssen, Roche,<br />

BMS<br />

Laurent Alric - Board Membership: Schering Plough, Schering Plough, Schering<br />

Plough, Schering Plough; Consulting: MSD; Speaking and Teaching: Roches,<br />

BMS, Gilead, Roches, BMS, Gilead, Roches, BMS, Gilead, Roches, BMS, Gilead,<br />

MSD, Abbvie<br />

Vincent Leroy - Board Membership: Abbvie, BMS, Gilead; Consulting: Janssen,<br />

MSD; Speaking and Teaching: Abbvie, BMS, Gilead, Janssen, MSD<br />

The following authors have nothing to disclose: Sylvie Radenne, Marie-Claude<br />

Gagnieu, Jean Didier Grange, Anne Minello, Nassim Kamar<br />

1159<br />

Therapeutic effect of dual oral therapy with Daclatasvir<br />

and Asunaprevir<br />

Motoyuki Kohjima 1,2 , Kenta Motomura 1 , Toshimasa Koyanagi 1 ,<br />

Seiya Tada 1 , Takeaki Satoh 1 , Naoki Yamashita 1 , Masaki Yokota 1 ,<br />

Rie Sugimoto 1 , Syoji Nagase 1 , Syusuke Morizono 1 , Nobito Higuchi<br />

1 , Tsuyoshi Yoshimoto 1 , Shigeki Tashiro 1 , Yuki Tanaka 1 , Kunitaka<br />

Fukuizumi 1 , Kazuhiro Kotoh 1 , Makoto Nakamuta 1 ; 1 Fukuoka<br />

Kanzo Treatment research (FKT) group, Fukuoka, Japan; 2 Gastroenterology,<br />

Clinical Research Center, Kyushu Medical Center,<br />

National Hospital Organization, Fukuoka, Japan<br />

Background: The dual oral therapy with the first-in-class NS5A<br />

replication complex inhibitor Daclatasvir (DCV) and the<br />

potent NS3 protease inhibitor Asunaprevir (ASV) has recently<br />

approved for the treatment of chronic hepatitis C genotype 1<br />

patients. Previous trials for the treatment was shown promising<br />

results for improving SVR to more than 80% in patients with<br />

HCV genotype 1b. We conducted a prospective, multicenter<br />

study to investigate the effectiveness of the dual oral therapy<br />

with DVC and ASV. Methods: The patients with HCV genotype<br />

1b have been treated with DCV and ASV since Sep. 2014<br />

(n=523). The enrolled population was generally older (median<br />

69 years old) that was consistent with HCV epidemiology in<br />

Japan and predominantly female (63%). HCV genotype and<br />

IL-28B polymorphisms were determined by PCR amplification<br />

and sequencing. HCV resistant associated polymorphisms<br />

were analyzed by direct sequence of the HCV NS3 and NS5A<br />

domains. Results: HCV-RNA declined rapidly after the initiation<br />

of the DCV + ASV treatment, and HCV-RNA was lower than<br />

detection sensitivity limit in 69% of the patients after 2 weeks<br />

and 82% of patients achieved RVR. The viral response of HCV<br />

was not affected by age, sex, liver fibrosis, previous treatment<br />

response, or IL-28B SNPs. The positive rate for HCV-RNA at<br />

4 weeks of treatment was significantly higher in patients with<br />

resistance-associated polymorphism in NS3 or NS5A and RVR<br />

rate was significantly higher in patients with add-on lipid modulators,<br />

pitavastain and EPA. Although 19 patients experienced<br />

viral breakthrough, 8 patients had no resistance-associated<br />

polymorphism in NS3 or NS5A among the patients with viral


782A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

breakthrough. Only 2 cases experienced viral relapse and<br />

93% of patients achieved SVR 4w. The response rate was<br />

tend to be higher in patients treated with pitavastain and EPA<br />

and significantly lower for patients with resistance-associated<br />

polymorphism (80%, p= 0.002) or treated cases with simeprevir<br />

(20%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 783A<br />

ropoietin. Among pts receiving TPV+pegIFN/RBV, 12 (10%)<br />

were transfused and 1 received erythropoietin due to anemia.<br />

Conclusions: In MALACHITE-I and MALACHITE-II, 3D+/-RBV<br />

had more favorable AE profiles than TPV+pegIFN/RBV. With<br />

3D+/-RBV, anemia was less frequent and less severe, and<br />

never required transfusion or use of erythropoietin. RBV was<br />

better tolerated in the context of newer DAAs than of TPV and<br />

pegIFN.<br />

Adverse events and hemoglobin decreases in MALACHITE-I and<br />

MALACHITE-II.<br />

*P1755μg.h/L after the first dose is predictive of sustained virological<br />

response (SVR) in HCV-patients treated with peginterferon/RBV*.<br />

The aim of this study was to test the clinical benefit<br />

of RBV early dose adjustment based on this threshold (Bayesian<br />

estimation) in under-exposed naïve genotype 1-infected<br />

patients. Methods: multicenter, randomized, controlled trial<br />

with two parallel groups; SC-group: standard of care in 2010-<br />

2011, i.e. peginterferon-α2a 180μg/week and weight-based<br />

RBV 1000-1200mg/day during 48 weeks; AD-group: personalized<br />

increase of RBV dose if D0AUC 0-4H<br />


784A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Veronique Loustaud-Ratti - Board Membership: Gilead, Roche, Schering Plough<br />

MSD; Speaking and Teaching: Roche, Schering Plough MSD, Janssen, Bristol<br />

meyers squibb, gilead, Abbvie<br />

Pierre Marquet - Advisory Committees or Review Panels: Astellas France, Sandoz<br />

France, BMS France; Consulting: Chiesi<br />

The following authors have nothing to disclose: Marianne Maynard-Muet, Sylvie<br />

Thevenon, Pierre Pradat, Annick Rousseau, Marie-Claude Gagnieu, Sophie<br />

Alain, Paul Deny, Françoise Lunel-Fabiani, Nicolas Picard, Irène Zublena, Christian<br />

Trepo<br />

1163<br />

Treatment of Hepatitis C Genotype 4 patients with Simeprevir<br />

and Sofosbuvir: Preliminary Results from a Phase<br />

IIa, Partially Randomised, Open-label Trial conducted in<br />

Egypt (OSIRIS)<br />

Maissa El Raziky 2 , Mohamed Gamil 2 , Radi Hammad 3 , Mohammed<br />

Saad Hashem 4 , Marwa Khairy 2 , Aisha Elsharkawy 2 , Asmaa<br />

Gomaa 4 , Sofia Keim 5 , Gino Van Dooren 1 , Robert Ryan 6 , Ralph<br />

DeMasi 6 , Isabelle Londjon-Domanec 7 , Mohamad Hassany 3 ,<br />

Wahid H. Doss 3 , Imam Waked 4 ; 1 Janssen Infectious Diseases<br />

BVBA, Beerse, Belgium; 2 Cairo University, Cairo, Egypt; 3 National<br />

Hepatology and Tropical Medicine Research Institute, Cairo,<br />

Egypt; 4 National Liver Institute, Menoufiya, Egypt; 5 Janssen-Cilag<br />

Portugal, Lisbon, Portugal; 6 Janssen Research & Development LLC,<br />

Titusville, NJ; 7 Janssen-Cilag, Paris, France<br />

Background and Aims Egypt has the highest prevalence of<br />

chronic hepatitis C virus (HCV) infection in the world, with<br />

>90% of patients infected with HCV G4. OSIRIS is a Phase IIa,<br />

open-label trial evaluating the efficacy and safety of 8 or 12<br />

weeks’ treatment with once-daily simeprevir (SMV) + sofosbuvir<br />

(SOF) for HCV genotype 4-infected patients in Egypt. Methods<br />

Sixty-three treatment-naïve or prior peg-interferon/ribavirin (PR)<br />

treatment-experienced patients with HCV G4 were treated with<br />

SMV 150 mg + SOF 400 mg daily, and were assigned to<br />

one of two groups; Group A (n=40): patients without cirrhosis<br />

were stratified by treatment history and fibrosis score, and<br />

randomised 1:1 to receive either 8 or 12 weeks of treatment;<br />

Group B (n=23): patients with cirrhosis, received 12 weeks<br />

of treatment. Absence of cirrhosis was documented by liver<br />

biopsy using METAVIR score. Cirrhosis was documented by<br />

either liver biopsy or FibroScan score >14 kPa; cirrhotic<br />

patients had to have platelet count >50 000/mm 3 and serum<br />

albumin >3 g/dL. The primary endpoint was the proportion<br />

of patients with HCV RNA 6 x10 6 IU/mL. One<br />

treatment-naïve patient in the 12-week arm with F1 fibrosis and<br />

BVL 1.2 x10 6 IU/mL relapsed. Initial safety data showed no<br />

discontinuations due to adverse events (AE), and no grade 3<br />

or 4 treatment-related adverse events have been reported. One<br />

serious AE (Grade 2 pleural effusion and pulmonary hypertension)<br />

was observed in a 69-year-old cirrhotic female patient<br />

after 6 weeks of treatment. This was unrelated to treatment and<br />

resolved in 2 weeks without treatment discontinuation. Conclusions<br />

SMV+SOF for 12 weeks in patients with HCV G4 is<br />

generally safe and well tolerated. High SVR4 rates (95–100%)<br />

were seen with 12 weeks’ treatment independently of prior PR<br />

response or cirrhosis. Final SVR12 data will be presented.<br />

Disclosures:<br />

Maissa El Raziky - Grant/Research Support: Janssen pharmaceuticals<br />

Mohamed Gamil - Speaking and Teaching: Bristol-Myers Squibb, Gilead Sciences,<br />

Janssen Pharmaceuticals, Merck Sharp & Dohme, Roche<br />

Radi Hammad - Grant/Research Support: Gilead Sciences,Inc, Janssen Pharmaceuticals,<br />

Inc.<br />

Sofia Keim - Employment: Janssen-Cilag Farmaceutica; Stock Shareholder: Johnson<br />

& Johnson<br />

Gino Van Dooren - Stock Shareholder: Johnson and Johnson<br />

Robert Ryan - Employment: Janssen Pharmaceuticals; Stock Shareholder: Janssen<br />

Pharmaceutical<br />

Ralph DeMasi - Employment: Janssen Pharmaceuticals<br />

Isabelle Londjon-Domanec - Employment: Janssen<br />

Mohamad Hassany - Grant/Research Support: Gilead Sc<br />

The following authors have nothing to disclose: Mohammed Saad Hashem,<br />

Marwa Khairy, Aisha Elsharkawy, Asmaa Gomaa, Wahid H. Doss, Imam<br />

Waked<br />

1164<br />

Limited Effectiveness of Sofosbuvir and Ribavirin for<br />

the Treatment of HCV GT-3 in Patients with Cirrhosis: A<br />

Real-World GT-3 HCV Treatment Experience<br />

Shirin Nafisi, Garrett Fante, Ann Moore, Ashley Hepner, Justin A.<br />

Reynolds, Karen Arendt, Yvette Cummings, Robert Gish, Richard<br />

A. Manch, Anita Kohli; Hepatology, St. Joseph’s Hospital and<br />

Medical Center, Scottsdale, AZ<br />

Recently approved direct-acting antivirals (DAA) regimens have<br />

improved the outcomes for many chronic hepatitis C (CHC)<br />

infected patients, particularly with genotypes 1, 2 and 4. Realworld<br />

non-clinical trial data on patients with CHC genotype<br />

3 (GT3) and new DAA regimens is lacking. AIM: To evaluate<br />

the effectiveness of DAA containing regimens for treatment<br />

of HCV GT3 in a real-world patient population and to compare<br />

to outcomes with previous use of pegylated interferon<br />

(PEG) and ribavirin (RBV). METHODS: We performed a retrospective<br />

analysis of patients with CHC GT3 treated between<br />

1/2012-6/2015 (n=73) at a community based clinic within<br />

an academic medical center in Phoenix, AZ or by ECHO telemedicine<br />

to rural clinic sites. Demographics, CHC treatment<br />

regimens, liver fibrosis, virologic data, and psychiatric history<br />

were collected throughout treatment and post-treatment from<br />

patient charts. RESULTS: Since 2012, 73 GT3 patients started<br />

treatment for CHC and are included in this analysis. 81% of<br />

patients were Caucasian, 14% Hispanic, and 1.7% Native<br />

American. 66% of patients were men, the mean age was 53<br />

yrs. (range: 21-72 yrs.) and mean HCV viral load 2,015,535<br />

IU/mL (range: 72-25,000,000 IU/mL). 27% of patients had<br />

cirrhosis by imaging, Fibrosure or liver biopsy. 55% of patients<br />

had a psychiatric illness. 35.1% of patients were previously<br />

treated with PEG containing therapy. 71.2% (n=52) were<br />

treated with the sofosbuvir (SOF)/RBV for 24 weeks, 1.3%<br />

(n=1) treated with SOF/RBV for 48 weeks, 5.4% (n=4) were<br />

treated with SOF/PEG/RBV for 12 weeks, 20.5% (n=15) were<br />

treated with PEG/RBV for 24 weeks, and 1.3% (n=1) were<br />

treated with ledipasvir (LDV)/SOF/RBV for 12 weeks. Overall,<br />

66% percent of patients treated with SOF/RBV for 24-48<br />

weeks and who have reached the SVR12 timepoint (n=30)<br />

have achieved SVR12 (pending for 23 patients). Comparing<br />

patients with or without cirrhosis who have reached SVR12


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 785A<br />

and were treated with SOF/RBV, 33% of these patients with<br />

cirrhosis achieved SVR12 as compared to 91% of patients<br />

without cirrhosis. 100% of patients treated with PEG/SOF/RBV<br />

who reached the SVR12 timepoint (n=2) have achieved SVR12<br />

(pending for 2 patients). 80% of patients treated with PEG/RBV<br />

achieved SVR12. Complete SVR12 results will be presented.<br />

CONCLUSION: Treatment of CHC GT3 using the PEG sparing<br />

regimen of SOF/RBV has poor results in cirrhotic patients in a<br />

real-world clinic setting. Consideration of alternative regimens,<br />

including SOF/PEG/RBV may be reasonable for these select<br />

patients or to wait for new treatments. Further development<br />

of PEG sparing DAA regimens for the treatment CHC GT3 is<br />

necessary for this difficult to treat population.<br />

Disclosures:<br />

Robert Gish - Advisory Committees or Review Panels: Gilead, AbbVie, Arrowhead;<br />

Consulting: Eiger, Isis, Genentech; Speaking and Teaching: Gilead, Abb-<br />

Vie; Stock Shareholder: Arrowhead<br />

Richard A. Manch - Speaking and Teaching: Gilead, AbbVie, Bayer, Salix, BMS<br />

The following authors have nothing to disclose: Shirin Nafisi, Garrett Fante, Ann<br />

Moore, Ashley Hepner, Justin A. Reynolds, Karen Arendt, Yvette Cummings,<br />

Anita Kohli<br />

1165<br />

Efficacy and Safety of Sofosbuvir-based Regimens in<br />

Asian-Americans With Chronic Hepatitis C Virus (HCV)<br />

Mono-Infection: A Multi-Center Study in the United<br />

States<br />

Calvin Q. Pan 1 , Elsa Ouyang 2 , Myron Tong 3 , Albert Min 4 , Ke-Qin<br />

Hu 5 , James Park 1 ; 1 Division of Gastroenterology and Hepatology,<br />

NYU Langone Medical Center, NYU School of Medicine, Flushing,<br />

NY; 2 Columbia University, New York, NY; 3 Digestive Diseases<br />

and Liver Center, David Geffen School of Medicine at UCLA, Los<br />

Angeles, CA; 4 Gasterenterology and Hepatology, Mount Sinai<br />

Beth Israel, Mount Sinai School of Medicine, New York, NY; 5 Division<br />

of Gastroenterology and Hepatology, UC Irvine Medical Center,<br />

Orange, CA<br />

Background Treatment with sofosbuvir (SOF)-based regimens<br />

for HCV infection has resulted in sustained virologic response<br />

(SVR) rates of approximately 90% in pivotal trials, in which<br />

Asian patients were underrepresented. This study aims to<br />

assess the efficacy and safety of SOF-based therapy in the<br />

Asian-American patients. Methods Asian patients with HCV<br />

genotype 1-6 mono-infection, who received SOF-based therapy<br />

for 12 or 24 weeks, were retrospectively enrolled from<br />

multiple centers throughout the United States. The primary endpoint<br />

was SVR 12. Secondary endpoints were the safety and<br />

tolerability of SOF-based treatment regimens. Results Among<br />

the 80 patients enrolled, 45 were treated with SOF+ribavirin<br />

(RBV), 15 with SOF+simprevir (SIM), 10 with SOF+RBV+peginterferon,<br />

8 with ledipasvir+SOF and 2 with SIM+SOF+RBV.<br />

Patient baseline values are shown in Table 1. By week 4, a<br />

rapid decrease in HCV RNA levels to


786A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

patients. Cox analyses showed that SVR was associated with<br />

reduced occurrence of cirrhosis-related complications (adjusted<br />

HR 0.19, 95%CI 0.10-0.35). The overall 5-year event rate was<br />

22.4% (95%CI 18.3-26.5) among those without SVR. With<br />

95% SVR the NNT to prevent 1 cirrhosis-related complication<br />

in 5 years was 27, 14, 9 or 7 at a 5-year complication rate of<br />

5, 10, 15 or 20%, respectively. With 85% SVR this was 30,<br />

15, 10 or 8, respectively. CONCLUSION The clinical efficacy<br />

of interferon-free therapy among patients with advanced but<br />

compensated liver disease is high, especially in case of an<br />

increased risk of cirrhosis-related complications. Small differences<br />

in the high SVR rates of these regimens have only limited<br />

effect on the clinical efficacy<br />

Disclosures:<br />

Adriaan J. van der Meer - Consulting: Gilead; Speaking and Teaching: MSD,<br />

Gilead<br />

Raoel Maan - Consulting: AbbVie<br />

Jordan J. Feld - Advisory Committees or Review Panels: Merck, Janssen, Gilead,<br />

AbbVie, Theravance, Bristol Meiers Squibb; Grant/Research Support: AbbVie,<br />

Boehringer Ingelheim, Janssen, Gilead, Merck<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

Jean-Francois Dufour - Advisory Committees or Review Panels: Bayer, BMS, Gilead,<br />

AbbVie, Novartis, Sillagen, Genfit<br />

Andres Duarte-Rojo - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Vital Therapies; Speaking and Teaching: Roche<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

Bart J. Veldt - Board Membership: GSK, Janssen Therapeutics<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Giovanna Fattovich, Frank Lammert,<br />

Wolf P. Hofmann, Donatella Ieluzzi, Bettina E. Hansen<br />

1167<br />

Early response and efficacy of dual oral therapy with<br />

asunaprevir and daclatasvir for elderly patients with<br />

hepatitis C<br />

Takashi Honda, Masatoshi Ishigami, Yoji Ishizu, Teiji Kuzuya,<br />

Kazuhiko Hayashi, Yoshiki Hirooka, Hidemi Goto; Gastroenterology<br />

and Hepatology, Nagoya University Graduate School of<br />

Medicine, Nagoya, Japan<br />

Background and aim: In Japan, patients with hepatitis C virus<br />

(HCV)-associated liver disease including hepatocellular carcinoma<br />

(HCC) are getting older and many patients die of such<br />

disease. The efficacy of dual oral therapy by using asunaprevir<br />

(NS3 protease inhibitor) and daclatasvir (NS5A replication<br />

complex inhibitor) for elderly patients with HCV genotype 1b<br />

has not been clarified. The aim of this study was to evaluate<br />

the early response and efficacy of dual oral therapy in such<br />

patients. Methods: Four hundred forty two consecutive chronic<br />

patients with HCV genotype 1b including compensated cirrhosis<br />

were treated with dual oral therapy with asunaprevir<br />

and daclatasvir in our hospital and affiliated hospital. These<br />

patients were divided into two groups according to age:<br />

elderly patients (aged ≥ 70 years, n = 240) and non-elderly<br />

patients (aged < 70 years, n = 202). Clinical characteristics,<br />

the rapid virological response (RVR) and the sustained virological<br />

response at post-treatment week 4 (SVR4) obtained<br />

by intention-to-treat analysis were compared between the two<br />

groups. Biochemical response and α-fetoprotein were compared<br />

at the start point and 4weeks later. Results: The ratio of<br />

female and compensated cirrhosis was significantly higher in<br />

elderly patients than in non-elderly patients (p = 0.006, 0.003,<br />

respectively). ALT, gGTP and hemoglobin level was significantly<br />

lower in elderly patients than that in non-elderly patients<br />

(P =0.0001, 0.010, 0.0001, respectively). However, other<br />

clinical characteristics of patients were not significantly different<br />

between two groups. The RVR and SVR4 rates didn’t differ<br />

significantly between elderly patients and non-elderly patients<br />

(RVR: 94.3 % vs 97.0 %, SVR4: 84.6 % vs 92.1 %). According<br />

to multivariate analysis, only RVR proved to be significantly<br />

associated with SVR4 while patient age did not affect SVR4.<br />

Multivariate analysis also indicated there is no significant factor<br />

associated with SVR4 in elderly patients. HCVRNA, ALT,<br />

γGTP and α-fetoprotein was rapidly and significantly reduced<br />

from start point to 4 weeks later in both elderly patients and<br />

non-elderly patients. Conclusions: Treatment of chronic hepatitis<br />

C including compensated cirrhosis with dual oral therapy was<br />

comparably effective between elderly patients and non-elderly<br />

patients. This dual oral therapy rapidly reduced ALT, γGTP and<br />

α-fetoprotein and may reduce incidence of HCC even in elderly<br />

patients.<br />

Disclosures:<br />

Hidemi Goto - Grant/Research Support: MSD, Roche, Bayer, Bristol-Myers, Eisai,<br />

Ajinomoto, Otsuka, Astra, Tanabe, Takeda<br />

The following authors have nothing to disclose: Takashi Honda, Masatoshi Ishigami,<br />

Yoji Ishizu, Teiji Kuzuya, Kazuhiko Hayashi, Yoshiki Hirooka


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 787A<br />

1168<br />

Resistance Analyses of Phase 3 Studies in Korean and<br />

Taiwanese Patients with Chronic Hepatitis C Receiving<br />

Sofosbuvir or Ledipasvir/Sofosbuvir Containing Regimens<br />

Kwang-Hyub Han 2 , Brian Doehle 1 , Hadas Dvory-Sobol 1 , Evguenia<br />

S. Svarovskaia 1 , Ramakrishna K. Chodavarapu 1 , Jenny C. Yang 1 ,<br />

Steven J. Knox 1 , Michael D. Miller 1 , Hongmei Mo 1 , Wan-Long<br />

Chuang 3 ; 1 Gilead Sciences, Foster City, CA; 2 Severance Hospital,<br />

Yonsei University Health System, Korea (the Republic of); 3 Kaohsiung<br />

Medical University Hospital, Kaohsiung City, Taiwan<br />

Background: The prevalence of chronic HCV infection varies<br />

globally, with approximately 1% and 4% of the population<br />

infected in Korea and Taiwan, respectively. Sofosbuvir (SOF)<br />

containing regimens can provide safe and highly effective all<br />

oral treatment regimens for patients infected with genotype<br />

(GT) 1 or 2 HCV. In this study we evaluated the impact of preexisting<br />

resistant-associated variants (RAVs) on treatment outcome<br />

and emergence of RAVs at relapse in patients retreated<br />

with SOF+RBV or ledipasvir (LDV)/SOF for 12 weeks in<br />

Korea and Taiwan. Methods: LDV/SOF (n=178) or SOF+RBV<br />

(n=216) was administered for 12 weeks to Korean or Taiwanese<br />

patients chronically infected with GT1 or 2 HCV. NS5A<br />

and/or NS5B deep sequencing analysis (cut-off of 1%) was<br />

performed for patients at baseline according to the assigned<br />

treatment regimen and at the time of virologic failure in patients<br />

failing to achieve SVR12. Data were compared to reference<br />

sequences and patient baseline sequence for patients failing<br />

to achieve SVR. Results: Virologic failure rates were low<br />

in Korean and Taiwanese patients; 1.1% (2/178) in LDV/<br />

SOF (GT1) recipients and 0.9% (2/216) in those receiving<br />

SOF+RBV (GT2). HCV GT1: Of the 178 patients with GT1<br />

HCV infection, 5 patients were incorrectly genotyped by LiPA<br />

and corrected to GT6 by sequence analysis; all 5 achieved<br />

SVR. Thirty-eight of 172 patients (22%) with available NS5A<br />

sequence data were observed to have baseline NS5A RAVs<br />

with 37/38 (97%) achieving SVR. The Y93H (n=30) and<br />

L31F/I/M/V (n=8) NS5A RAVs as single variants were the<br />

most commonly observed in both populations. Four of the 170<br />

patients with available sequence data (2%) were observed to<br />

have the NS5B RAV L159F at baseline and all achieved SVR.<br />

In the two GT1 patients that relapsed post treatment, one was<br />

observed to have NS5A RAVs at baseline and relapse, and the<br />

other had treatment emergent NS5A RAVs. No NS5B RAVs<br />

were observed in these two patients. HCV GT2: Six of 184<br />

patients (3%) with available baseline NS5B sequencing data<br />

were observed to have M289I/L variants, and all achieved<br />

SVR when treated with SOF+RBV. No other baseline NS5B<br />

RAVs were observed. Two GT2 patients experienced virologic<br />

failure, and no baseline or treatment emergent NS5B RAVs<br />

were observed. Conclusions: Low rates of virologic failure were<br />

observed in Korean and Taiwanese patients and the presence<br />

of NS5A and NS5B RAVs did not affect the treatment outcome<br />

to LDV/SOF and SOF+RBV in patients with GT1 and GT2<br />

infection, respectively. No S282T or other SOF treatment-associated<br />

variants have been detected in the patients that experienced<br />

virologic failure during treatment.<br />

Disclosures:<br />

Brian Doehle - Employment: Gilead Sciences<br />

Hadas Dvory-Sobol - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Evguenia S. Svarovskaia - Employment: Gilead Sciences Inc.; Stock Shareholder:<br />

Gilead Sciences Inc.<br />

Ramakrishna K. Chodavarapu - Employment: Gilead Sciences, Inc<br />

Jenny C. Yang - Employment: Gilead Sciences, Inc<br />

Steven J. Knox - Employment: Gilead Sciences<br />

Michael D. Miller - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Hongmei Mo - Employment: Gilead Science Inc<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

The following authors have nothing to disclose: Kwang-Hyub Han<br />

1169<br />

Sustained Virologic Response (SVR) of Liver Transplant<br />

Recipients (OLT) with Advanced Hepatitis C Recurrence<br />

(HCVR), All Genotype 1, after Combination Antiviral<br />

Therapy (Rx) with either Telaprevir(TVR) or Sofosbuvir/<br />

Simeprevir (SOF/SIM): A Single Center Experience<br />

Carlos Fasola, Parvez S. Mantry, Jeffrey S. Weinstein, Hector E.<br />

Nazario, Adil Habib, Maisha Barnes, Alejandro Mejia, Edward<br />

A. Dominguez, Richard Dickerman, Stephen Cheng; Liver Institute,<br />

Methodist Dallas Medical Center, Dallas, TX<br />

Aim: Liver allografts of viremic HCV-OLT have a higher risk of<br />

failure. Timely intervention is necessary. However, long-term<br />

outcome reports, especially using newer Rx, is lacking. Methods:<br />

in a 4-year period, we compared 2 groups of OLT-HCVR<br />

Rx: A) TVR (n=19) + Pegylated interferons (P-IFN) + Ribavirin<br />

(RBV) for 12 weeks, followed by P-IFN+RBV for 12-36 weeks<br />

or, B) SOF/SIM (n=28) Rx for 12-24 weeks. Rx criteria: OLT<br />

graft dysfunction and/or HCVR histology (stage 0-4, Batts&Ludwig).<br />

Immunosuppression based on tacrolimus or cyclosporine<br />

+ mycophenolic acid + low-dose prednisone (limited time).<br />

Demographics, laboratory tests, including HCV-RNA levels,<br />

graft and OLT survivals were analyzed. Minimal follow: 24<br />

weeks post end of Rx. Attention focused on rapid virologic<br />

response (RVR: aviremic in < 4 weeks), early virologic response<br />

(EVR: aviremic in < 8 weeks) and SVR (aviremic for > 24 weeks<br />

post Rx). Statistics: t-tests, Chi2-tests and Kaplan-Meier. Results:<br />

The majority of the patients (14/19 and 21/28, 74-75%) were<br />

non-responder or intolerant to previous Rx. Only a quarter of<br />

patients were naive to HCV Rx on each group. No demographic<br />

or laboratory differences were found, except for a<br />

similar incidence of anemia that required epoetin Rx. Liver tests<br />

remained normal in 26/28 (93%) of the SOF/SIM group. The<br />

pre-Rx advanced stage was significantly higher in the SOF/<br />

SIM group (table). No differences were observed regarding<br />

patient and allograft survivals. Three patients (16%) died due<br />

to allograft failure in group A. In group B, 2 (7%) died, one<br />

of them of trauma with stable LFT’s. Three patient relapsed in<br />

the TVR group. No differences, between groups, were found<br />

in the rate of HCV-RNA levels decrease log during the first 8<br />

weeks, however there was a significant difference in the SVR<br />

rates favoring the SOF/SIM combination (table). Conclusions:<br />

Newer anti-HCV therapies have significantly improved the SVR<br />

and outcome of OLT recipients with HCVR. The benefit of SOF/<br />

SIM combination therapy has shown, in our experience, to be<br />

an important addition to the effort of rescueing liver allograft<br />

function post OLT in HCVR recipients.<br />

Characteristics and Outcome of Liver Transplant Recipients with<br />

Hepatitis C Recurrence after Antiviral Therapy<br />

P value < 0.05: 1 vs. 1 and 2 vs. 2


788A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Parvez S. Mantry - Consulting: Salix, Gilead, Janssen, Abbvie, BMS; Grant/<br />

Research Support: Salix, Merck, Gilead, Boehringer-Ingelheim, Mass Biologics,<br />

Vital Therapies, Santaris, mass biologics, Bristol-Myers Squibb, Abbive, Bayer-Onyx,<br />

Shinogi, Tacere, Intercept; Speaking and Teaching: Gilead, Janssen,<br />

Salix<br />

Jeffrey S. Weinstein - Speaking and Teaching: Merck<br />

Hector E. Nazario - Advisory Committees or Review Panels: Gilead, Janssen;<br />

Speaking and Teaching: Gilead, Merck, Abbvie, Janssen<br />

Edward A. Dominguez - Advisory Committees or Review Panels: Gilead, Pfizer;<br />

Grant/Research Support: Cubist; Speaking and Teaching: Amgen, Astelleas<br />

The following authors have nothing to disclose: Carlos Fasola, Adil Habib, Maisha<br />

Barnes, Alejandro Mejia, Richard Dickerman, Stephen Cheng<br />

1170<br />

Treatment of Chronic HCV Genotype 1 (G1) Infection<br />

with Boceprevir: Frequency, Severity, Predictability and<br />

Management of Anemia in German Real-Life<br />

Gerlinde Teuber 1 , Peter Buggisch 2 , Hanns F. Loehr 3 , Hermann<br />

Steffens 4 , Michael R. R. Kraus 5 , Peter R. Geyer 6 , Bernd Weber 7 ,<br />

Thomas Witthoeft 8 , Uwe Naumann 9 , Elmar Zehnter 10 , Dagmar<br />

Hartmann 11 , Bernd Dreher 11 , Manfred Bilzer 11 ; 1 Gastroenterological<br />

Practice, Frankfurt, Germany; 2 IFI Liver Center Hamburg,<br />

Hamburg, Germany; 3 Gastroenterological Practice, Wiesbaden,<br />

Germany; 4 Practice of Internal Medicine, Berlin, Germany;<br />

5 Department II, Klinikum Burghausen, Burghausen, Germany;<br />

6 Gastroenterological Practice, Fulda, Germany; 7 Competence<br />

Center Addiction, Kassel, Germany; 8 Gastroenterological Practice,<br />

Stade, Germany; 9 Center of Medicine, Berlin, Germany; 10 Gastroenterological<br />

Practice, Dortmund, Germany; 11 MSD Pharma<br />

GmbH, Haar, Germany<br />

Background: Anemia is a frequent complication in patients<br />

(pts) undergoing triple therapy with the HCV protease inhibitor<br />

boceprevir (BOC). Data regarding the frequency, severity,<br />

predictability and management of anemia caused by triple<br />

therapy in real-life are still scarce and were evaluated in the<br />

present interim analysis. Methods: From April 2012 until January<br />

2014, 536 pts with HCV G1 infection were recruited in<br />

the ongoing NOVUS observational study by 97 practices and<br />

hospitals in Germany. Pts were treated with pegylated interferons<br />

(PegIFN) and ribavirin (RBV) together with BOC for 24 to<br />

44 weeks after a 4 weeks lead-in period with PegIFN/RBV. The<br />

present interim analysis was restricted to 265 treatment-naïve<br />

and 132 previously treated pts with documented hemoglobin<br />

(Hb) levels at baseline and during therapy. Results: During<br />

BOC triple therapy anemia (Hb


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 789A<br />

past regimens in achieving SVR in this difficult to treat population.<br />

Table 1. Main demographic, baseline clinical, virological and<br />

laboratory (values expressed as mean ± standard deviation) characteristics<br />

of HCV-MC patients<br />

or ombitasvir/paritaprevir/ritonavir/dasabuvir for GT1 and<br />

sofosbuvir/ribavirin for GT 2/3. Data was included from 13<br />

additional patients who are still awaiting medication approval<br />

for a total of 54 patients. Among these 54 patients, HCV<br />

medications were approved in 84% of patients with private<br />

insurance, 71% of Medicare and 50% of Medicaid patients<br />

(p=0.11). There was no difference in approval rate based<br />

on age, race, gender, genotype, treatment regimen or presence<br />

of cirrhosis. The median time spent per patient was 92.5<br />

minutes (IQR 80.0-105.0). The median cost spent per patient<br />

was $60.54 (IQR 53.00-72.90). There was no difference in<br />

time spent acquiring HCV medications or cost per patient by<br />

genotype, insurance status, treatment regimen or presence of<br />

cirrhosis. Conclusion: In a Hepatology practice with experienced<br />

staff using a specialty pharmacy to improve efficiency,<br />

the median provider time needed to obtain DAAs was 90 minutes<br />

per patient. This relatively labor intensive process will need<br />

to be streamlined to encourage more clinicians to engage in<br />

HCV therapy for the large number of potential candidates in<br />

the United States.<br />

Disclosures:<br />

The following authors have nothing to disclose: Veronica Loy, Tamara Benyashvili,<br />

Megan OMahony, Douglas Pavkov, William Adams, Robin Todus, Michelle<br />

Watts, Natasha Von Roenn, Scott Cotler<br />

HCV:hepatitis c virus<br />

MC: mixed cryoglobulinemia<br />

Disclosures:<br />

T. Craig Cheetham - Grant/Research Support: Gilead, BMS<br />

The following authors have nothing to disclose: Rasham Mittal, Zoe Bider-Canfield,<br />

Amandeep K. Sahota<br />

1172<br />

The Time and Cost Investment Required to Obtain and<br />

Initiate Direct Acting Antiviral Therapy<br />

Veronica Loy, Tamara Benyashvili, Megan OMahony, Douglas<br />

Pavkov, William Adams, Robin Todus, Michelle Watts, Natasha<br />

Von Roenn, Scott Cotler; Hepatology, Loyola Medical Center, Chicago,<br />

IL<br />

The biggest barrier to direct-acting antiviral agent (DAA) therapy<br />

for hepatitis C (HCV) is cost. Initiating treatment with DAAs<br />

can be time intensive for the provider, including obtaining<br />

prior authorization. The purpose of this study was to assess the<br />

amount of unbillable time and cost needed to obtain DAAs for<br />

HCV. Methods: We prospectively enrolled patients for whom<br />

DAA therapy was planned from 9/30/2014- 3/19/2015.<br />

Providers recorded the amount of time spent in minutes to<br />

obtain medication and to initiate HCV therapy for each patient.<br />

The position of each provider (APN, RN, LPN) was noted. Data<br />

included age, gender, race, HCV genotype, type of insurance<br />

carrier, and presence of cirrhosis. Patients who subsequently<br />

defered therapy and those with incomplete time data were<br />

excluded from data analysis. In our practice, 2 LPNs and 6<br />

RNs collect patient information and use a specialty pharmacy<br />

to obtain approval for DAAs. An APN reviews potential medication<br />

interactions and conducts a patient education session<br />

by phone. Time spent was converted to cost per patient based<br />

on average hourly medical wages in Illinois by Salary.com.<br />

Administrative time and cost per patient was compared by<br />

insurance carrier, HCV genotype, treatment regimen and presence<br />

of cirrhosis. Results: 79 patients consented to participate<br />

in the study. 25 were excluded due to incomplete time data<br />

(n=17) or subsequent deferral of therapy (n=8). 41 patients<br />

were started on therapy consisting of sofosbuvir/ledipasvir<br />

1173<br />

Health resource usage in peri-transplant patients during<br />

12 weeks of therapy with sofosbuvir (SOF), with or<br />

without ribavirin (RBV) and an NS5A inhibitor<br />

Suman Verma 1 , Graham R. Foster 2 , William Irving 3 , John McLauchlan<br />

5 , Michelle C. Cheung 2 , Benjamin E. Hudson 4 , Ivana Carey 1 ,<br />

Kosh Agarwal 1 ; 1 Institute of liver <strong>studies</strong>, Kings College Hospital,<br />

London, United Kingdom; 2 The Blizard Institute, Queen Marys University<br />

of London, London, United Kingdom; 3 Virology, University<br />

of Nottingham, Nottingham, United Kingdom; 4 University of<br />

Nottingham, Nottingham, United Kingdom; 5 MRC University of<br />

Glasgow, London, United Kingdom<br />

Introduction Direct acting antiviral (DAA) therapy for HCV<br />

decompensated disease is a reality. However, health economic<br />

benefit in this advanced peritransplant population remains<br />

unclear. This study aims to assess healthcare resource utilization<br />

and quality returns from a clinical and cost perspective<br />

during 12 weeks (wks) of DAA therapy in this challenging population.<br />

Methods The UK NHS England Early Access Program<br />

(EAP) mandated antiviral therapy to HCV positive patients with<br />

advanced liver disease via 17 centres. Twelve weeks of therapy<br />

with SOF and an NS5A inhibitor (Ledipasvir or Daclatasvir),<br />

with or without RBV was administered to a cohort of<br />

~500 decompensated cirrhostics, with 179 peritransplant (PT).<br />

PT was defined as meeting the NHSBT criteria for liver transplantation<br />

(LT) (n=56); on the national LT waiting list (n=50);<br />

or a previous LT with advanced HCV recurrence in graft and<br />

portal hypertension (n=45). Choice of therapy was clinician<br />

discretion. Data collection was prospective during the 12 wks<br />

of therapy via HCV Research UK on use of blood products;<br />

patient phone calls made to healthcare professionals including<br />

specialist hepatitis nurses (SHN) and consultants; unscheduled<br />

clinic appointments; and elective and emergency admissions<br />

into hospital and intensive care (ITU). Results Of the179 PT<br />

patients enrolled, 151 had completed 12 wks of therapy at<br />

the time of abstract submission. During this, 148 phone calls<br />

were made to SHN; 5 to hepatology consultants and 42 to<br />

other healthcare professionals (OHP) including GPs and clinical<br />

pharmacists, with a total duration of 993, 25, 280 minutes<br />

respectively. A total of 367 unscheduled clinic appointments


790A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

were required during the treatment period with 178 delivered<br />

by consultants, 112 by SHN and 77 by OHP. Support for<br />

anemia with erythropoietin, cosmofer or blood transfusion was<br />

required equally across the PT cohorts (previous LT 7, fulfilling<br />

NHSBT criteria 5, listed for LT 7). In addition there were<br />

133 hospital admissions with 25% electively for ascites drainage,<br />

and as an emergency 5% for variceal bleed and 9% for<br />

encephalopathy. Furthermore, 21 patients underwent LT, with<br />

complete inpatient stay data available for 13. This demonstrated<br />

an average stay of 18.07 days, with 2.85 days on ITU<br />

and 1.53 days ventilated. Further data will be presented at<br />

AASLD specifically addressing financial costs. Conclusion This<br />

preliminary analysis of health resource usage by PT patients<br />

during SOF and NS5A inhibitor +/- RBV demonstrates that this<br />

challenging advanced disease population still utilize significant<br />

healthcare resource despite better tolerability of the new DAA<br />

therapies.<br />

Disclosures:<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

William Irving - Advisory Committees or Review Panels: Novartis, MSD, Janssen<br />

Cilag, Bristol Myers Squibb; Grant/Research Support: GSK, Pfizer, Janssen<br />

Cilag, Gilead Sciences; Speaking and Teaching: Janssen Cilag, Roche<br />

Ivana Carey - Grant/Research Support: Gilead, Roche; Speaking and Teaching:<br />

BMS<br />

Kosh Agarwal - Advisory Committees or Review Panels: Gilead, Novartis,<br />

Abbott; Grant/Research Support: Roche, MSD; Speaking and Teaching: BMS,<br />

Astellas, Janssen<br />

The following authors have nothing to disclose: Suman Verma, John McLauchlan,<br />

Michelle C. Cheung, Benjamin E. Hudson<br />

1174<br />

Outcomes of ribavirin free regimens for treatment of<br />

recurrent hepatitis C (genotype1) infection post liver<br />

transplant<br />

Mohamed G. Shoreibah 1 , DeAnn Jones 2 , Sarah Baggett 2 , Brendan<br />

M. McGuire 1 , Omar Massoud 1 ; 1 GI and Hepatology, University<br />

of Alabama at Birmingham, Birmingham, AL; 2 Pharmacy,<br />

University of Alabama at Birmingham, Birmingham, AL<br />

Background. The development of direct-acting antivirals in<br />

the treatment of Hepatitis C Virus (HCV) is a rapidly evolving<br />

field. Data on the use of the newer agent in the treatment<br />

of recurrent HCV infection post liver transplant (LT) is scarce.<br />

Methods. We performed a retrospective analysis comparing<br />

two groups treated with RIBA free regimens for recurrent HCV<br />

infection, genotype 1, post LT: Group A, treated with SIM/<br />

SOF and Group B, treated with LDV/SOF (without RIBA).<br />

Patients received treatment between October 2014 and April<br />

2015. Patients were identified through hepatology department<br />

records. Data collection included genotype, prior HCV therapies,<br />

cirrhosis, and virologic response at 4 and 12 weeks on<br />

treatment. A one sided t-test was used to determine statistical<br />

significance. Results. A total of 78 patients met inclusion criteria<br />

with mean time from LT of 61 months. Descriptive statistics<br />

were: Group A (N:36, mean age: 58 years, 69% male, 58%<br />

treatment experienced, 73% GT1a, 11% cirrhotics, 1 patient<br />

with fibrosing cholestatic HCV) and Group B (N:42, mean<br />

age: 60 years, 69% male, 55% treatment experienced, 77%<br />

GT1a, 7% SOF failure, 10% cirrhotics, ). Group A patients<br />

were treated for 12 weeks compared to planned 12 weeks in<br />

76% and 24 weeks in 24% of patients in Group B. In group A,<br />

100% completed treatment. End of treatment response (EOTR)<br />

was 94% and available sustained virologic response (SVR12)<br />

rate was 91%. One patient with fibrosing cholestatic HCV<br />

achieved SVR12. All three patients who experienced relapse in<br />

group A are currently enrolled in Group B with a planned 24<br />

week treatment duration. A total of 50% of patients in Group B<br />

have completed treatment. Available EOTR is 100 (N:20) and<br />

available SVR12 is 100% (N:4). A total of 22% of patients<br />

had renal insufficiency with a mean glomerular filtration rate<br />

of 45 ml/min at the start of treatment. The mean albumin was<br />

3.64 gm/dL at the start of treatment and 3.9 gm/dL at SVR12<br />

(P=0.013), which suggests improved synthetic liver function.<br />

No discontinuation of treatment due to adverse events occurred<br />

in either group. Data collection is ongoing. Conclusion. The<br />

combination of SIM/SOF showed a high efficacy of 91% in the<br />

treatment of recurrent HCV post LT with excellent tolerability.<br />

The use of LDV/SOF in this group of patients (without RIBA) is<br />

being evaluated in our institution and appears to have excellent<br />

virologic response and tolerability. Additional SVR12 data will<br />

be reported at the time of the meeting.<br />

Disclosures:<br />

Mohamed G. Shoreibah - Advisory Committees or Review Panels: Gilead<br />

Brendan M. McGuire - Grant/Research Support: bayer healthcare, salix<br />

The following authors have nothing to disclose: DeAnn Jones, Sarah Baggett,<br />

Omar Massoud<br />

1175<br />

Predictors of Treatment Adherence and Discontinuation<br />

in Department of Defense (DoD) Health Care Beneficiaries<br />

Treated for Chronic Hepatitis C 2004-2013<br />

Dana Y. Teltsch 2 , David R. Walker 1 , Beth L. Nordstrom 2 , Kathy<br />

Fraeman 2 , Karl Kronmann 3 , Kristina St Clair 3 ; 1 HEOR, Abbvie,<br />

North Chicago, IL; 2 Retrospective Observational Studies, Evidera,<br />

Lexington, MA; 3 Naval Medical Center Portsmouth, Portsmouth,<br />

VA<br />

Background and objective: New treatments for chronic hepatitis<br />

C virus (HCV) promise excellent sustained virologic response<br />

and cure rate. Treatment effectiveness, however, depends on<br />

regimen compliance. Our goal was to identify predictors of<br />

adherence and discontinuation in a veteran, dependent and<br />

active duty military population receiving chronic HCV treatment.<br />

Methods: We conducted a retrospective cohort study<br />

of chronic HCV patients covered by the DoD health system<br />

who initiated HCV treatments during 2004-2013. The defined<br />

regimens included: peg-interferon (peg)+ribavirin (rib) alone<br />

or in combination with boceprevir (boc)/telaprevir (tel)/simeprevir<br />

(sim); or sofosbuvir (sof) in combination with rib/sim/<br />

rib+sim/peg+rib; all regimens were studied for 12 weeks after<br />

initiation. The time to regimen discontinuation was calculated<br />

allowing up to a 60 (peg and rib) or 15 (other drugs) day gap,<br />

and predictors of time to discontinuation identified using multivariate<br />

Cox models. Adherence was measured by proportion<br />

of days covered (PDC), with PDC > 80% considered adherent.<br />

Multivariate logistic regression models sought predictors of<br />

adherence. Potential predictors included patient demographics,<br />

liver disease status, comorbidities, and baseline resource<br />

utilization. Results: 6,196 patients initiated treatment with a<br />

regimen of interest (4,756 peg+rib; 212 peg+rib+boc; 631<br />

peg+rib+tel; 238 sof+sim; 24 sof+sim+rib; 178 sof+rib+peg;<br />

157 sof+rib; 0 sim+peg+rib ). Mean age was 49.9 years, and<br />

47.3% were female. 71.9% of patients had non-cirrhotic HCV;<br />

26.5% compensated or decompensated cirrhosis; 1.6% were<br />

post liver transplant. There was a small but consistent association<br />

between increased time from diagnosis in years and<br />

both earlier treatment discontinuation (hazard ratio (HR) 1.03;<br />

95% confidence interval (CI) 1.00-1.06) and lower adherence<br />

(odds ratio (OR) 0.96; (0.93-0.99)). Predictors of discontinua-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 791A<br />

tion included age >60 years at index date (1.18; 1.00-1.38,<br />

compared to age 51-60); chronic kidney disease (1.51; 1.13-<br />

2.02), diabetes (1.28; 1.09-1.50), HIV status (1.86; 1.24-<br />

2.79), and psychiatric disorders (1.14; 1.02-1.28). Predictors<br />

of adherence included decompensated cirrhosis (0.75; 0.61-<br />

0.94 compared to non-cirrhotic HCV); chronic kidney disease<br />

(0.70; 0.50-0.99); diabetes (0.80; 0.68-0.95); and HIV status<br />

(0.59; 0.36-0.96). Conclusions: Later treatment time from diagnosis,<br />

patient comorbidities, and older age were associated<br />

with poor adherence and with discontinuation prior to the recommended<br />

regimen duration. Results suggest that delaying<br />

treatment for many years after HCV diagnosis may adversely<br />

impact treatment compliance.<br />

Disclosures:<br />

Dana Y. Teltsch - Consulting: AbbVie; Employment: Evidera<br />

David R. Walker - Employment: Abbvie; Stock Shareholder: Abbvie, Baxter<br />

Healthcare<br />

The following authors have nothing to disclose: Beth L. Nordstrom, Kathy Fraeman,<br />

Karl Kronmann, Kristina St Clair<br />

study suggests as a result of a good tolerability profile, monitoring<br />

and AE related costs are minimal in LDV/SOF regimens.<br />

This study also suggests that, when the 8w regimen is used,<br />

the cost per SVR is significantly lower in naïve and NC when<br />

compared to TE and cirrhotic patients, indicating an economic<br />

benefit of selecting high effectiveness and well tolerated first<br />

line therapies and early treatment.<br />

Disclosures:<br />

Peter Buggisch - Advisory Committees or Review Panels: Janssen, AbbVie, BMS;<br />

Speaking and Teaching: Roche, MSD, Gilead, Merz Pharma<br />

Karsten Wursthorn - Grant/Research Support: BMS<br />

Albrecht Stoehr - Board Membership: MSD, Böhringer Ingelheim; Speaking<br />

and Teaching: Janssen, MSD, Gilead, Abbvie, Böhringer Ingelheim, BMS, VIIV<br />

Aline Gauthier - Consulting: Gilead<br />

Petar K. Atanasov - Consulting: Gilead<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

1176<br />

Real-world effectiveness and cost per SVR of ledipasvir/<br />

sofosbuvir chronic hepatitis C treatment<br />

Peter Buggisch 1 , Karsten Wursthorn 1 , Albrecht Stoehr 1 , Aline<br />

Gauthier 2 , Petar K. Atanasov 2 , Joerg Petersen 1 ; 1 Asklepios Klinik<br />

St. Georg Haus L, IFI Institut für Interdisziplinäre Medizin, Hamburg,<br />

Germany; 2 Amaris Consulting, London, United Kingdom<br />

Background and Aims: Ledipasvir/Sofosbuvir (LDV/SOF) single<br />

tablet regimen (STR) is approved in Europe and the US for<br />

the treatment of chronic hepatitis C (CHC) patients. With the<br />

emergence of novel, highly effective, and safe therapies and<br />

the expected demand for them, the need for optimal resource<br />

allocation is high. The cost per sustained viral response (SVR)<br />

is a measure which provides insights into the amount spent for<br />

the achievement of success in CHC therapy. This study aims<br />

to assess the safety, effectiveness, and the cost per SVR associated<br />

with LDV/SOF therapy in clinical practice in Germany.<br />

Methods: The first CHC patients treated with LDV/SOF in a<br />

single centre (and for whom SVR after 12 weeks of follow-up<br />

(SVR12) will be available in November 2015) were included<br />

in the analysis. Baseline characteristics, prior treatment history,<br />

safety, effectiveness and cost per SVR were investigated<br />

using descriptive statistics. Results: 115 patients met the inclusion<br />

criteria for this analysis. 8w (51.3%), 12w (44.3%) or<br />

24w (4.4%) treatment with LDV/SOF was initiated between<br />

21/11/2014 - 03/03/2015. 21% of patients had ribavirin<br />

(RBV) added to the STR (79% of which F4). The mean (SD) age<br />

was 52 (12.2) years, 60% were male, 89.6% had at least one<br />

comorbidity. Genotype (GT) distribution: 57%, 32%, 6% and<br />

4% for 1a, 1b, 3 and 4, respectively. METAVIR stage distribution:<br />

34%, 19%, 13%, 11% and 23% for F0, F1, F2, F3 and<br />

F4, respectively. Median (IQR) HCV RNA at baseline was 0.97<br />

(0.28-1.90) million IU/ml. 6% (2%) of patients were HIV (HBV)<br />

co-infected. In patients with available outcome data, the SVR4<br />

was 99% (n=92/93) and SVR12 was 100% (n=13/13). One<br />

treatment experienced (TE) F4 patient on 12w STR+RBV did not<br />

achieve SVR4. 6.9% (n=8) experienced grade 3 or 4 adverse<br />

events (AE) and 5.2% (n=6) were assessed as treatment-related.<br />

No AE lead to discontinuation. 1.0% of costs were<br />

non-therapy costs. The median cost per SVR4 was €53,025 (PP<br />

like approach). 73% of naïve patients and 66% of non-cirrhotic<br />

(NC) were on 8w duration; median cost per SVR4 was 59%<br />

lower in NC (€46,775) than in F4 patients and 58% lower in<br />

naïve (€46,272) versus TE patients. SVR12 and cost per SVR12<br />

will be available at the time of presentation. Conclusion: This<br />

1177<br />

Hope and Hopelessness in HCV patients in the era of<br />

DAA therapy<br />

Sophie K. Afdhal 1 , Michael Penn 1 , Zobair M. Younossi 2 , Michael<br />

P. Curry 3 ; 1 Psychology, Franklin & Marshall College, Charlestown,<br />

MA; 2 Inova Fairfax Hospital, Falls Church, VA; 3 Beth Israel Deaconess<br />

Medical Center, Boston, MA<br />

Hope and hopelessness are important patient responses to<br />

chronic illness and can influence recovery from disease and<br />

effect patient reported outcomes (PRO’s). In the era of Direct<br />

Acting Antiviral (DAA) therapy with high SVR rates greater<br />

than 90% and good tolerability we hypothesized that treatment<br />

expectations might alter the hope of patients with HCV. AIMS:<br />

To evaluate hope and hopelessness in a HCV population and<br />

to correlate it with disease variables and PRO’s such as fatigue,<br />

depression and QOL. PATIENTS and METHODS: Consecutive<br />

adult patients with chronic HCV were enrolled from the<br />

Liver Center at BIDMC and underwent education about the<br />

new DAA therapies and then completed the Beck Hopelessness<br />

survey (BHS), Beck Depression index (BDI), fatigue questionnaire<br />

(FACIT-F) and the SF-36 health survey. Demographic,<br />

social, educational and disease state data was collected. Data<br />

was analyzed by Chi square for comparison between groups,<br />

regression analysis and ANOVA. RESULTS: 158 patients were<br />

enrolled over June to August 2014 and were predominantly<br />

male (63%), Caucasian (75%) with over 56% having at least<br />

some college level education. 52% of patients believed that<br />

HCV was responsible for causing or worsening health related<br />

issues, 40% had a history of depression, 30% had a history of<br />

chronic pain, 17% had PTSD, 21% had current drug use and<br />

35% current alcohol use. Mode of acquisition of HCV was drug<br />

use in 38%. RESULTS: Mean hopelessness score (range 0 – 20)<br />

was 4.25 + 4.5 and 25 patients had severe hopelessness with<br />

a score > 10. BDI mean score 12.6 + 10.6 and 39 patients<br />

had moderate to severe depression. Patients with a history of<br />

depression or active depression were strongly correlated with<br />

hopelessness; P < 0.0001. Patients with PTSD were more likely<br />

to suffer from hopelessness (p


792A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

related to coexisting factors such as depression, PTSD and disease<br />

beliefs. The effect of successful treatment on hopelessness<br />

in this cohort is being evaluated.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Michael P. Curry - Advisory Committees or Review Panels: Bristol Meyers Squib,<br />

Abbvie; Grant/Research Support: Gilead Sciences, Mass Biologics, Merck,<br />

Salix, Conatus; Stock Shareholder: Achilion<br />

The following authors have nothing to disclose: Sophie K. Afdhal, Michael Penn<br />

1178<br />

Sofosbuvir is well tolerated and effective in chronic hepatitis<br />

C patients with advanced renal disease and/or on<br />

hemodialysis<br />

Geoffrey Kitzman 1 , Eric Davis 1 , Jesus Monico 2 , Evelyn Arendale 1 ,<br />

Brian B. Borg 1 ; 1 Digestive Disease, University of Mississippi Medical<br />

Center, Jackson, MS; 2 Pathology, University of Mississippi<br />

Medical Center, Jackson, MS<br />

Background: Sofosbuvir (Sof) is a nucleotide analog inhibitor<br />

of HCV NS5B polymerase. It is used in combination therapies<br />

with ribavirin or other DAAs. The safety and efficacy of sofosbuvir<br />

in patients with advanced renal disease (GFR


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 793A<br />

could potentially select F0–2 treatment-naïve G4 patients suitable<br />

to shorten SMV+PR therapy to 12 weeks.<br />

Disclosures:<br />

Tarik Asselah - Consulting: Jannsen, AbbVie, Achillion, BMS, Gilead, Merck,<br />

Roche; Speaking and Teaching: Jannsen, AbbVie, Achillion, BMS, Gilead,<br />

Merck, Roche<br />

Christophe Moreno - Consulting: Abbvie, Janssen, Gilead, BMS; Grant/Research<br />

Support: Janssen, Gilead, Roche, Astellas<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

Michael Gschwantler - Advisory Committees or Review Panels: Janssen, BMS,<br />

Gilead, AbbVie; Speaking and Teaching: Janssen, BMS, Gilead, AbbVie<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

Peter Buggisch - Advisory Committees or Review Panels: Janssen, AbbVie, BMS;<br />

Speaking and Teaching: Roche, MSD, Gilead, Merz Pharma<br />

Faisal M. Sanai - Advisory Committees or Review Panels: Merck Sharpe Dohme,<br />

Bristol Myers Squibb, Janssen Pharmaceuticals, Gilead Sciences; Grant/<br />

Research Support: Roche Pharmaceuticals, Bristol Myers Squibb; Speaking and<br />

Teaching: Roche Pharmaceuticals, Janssen Pharmaceuticals, Gilead Sciences,<br />

Bayer Schering<br />

Robert Ryan - Employment: Janssen Pharmaceuticals; Stock Shareholder: Janssen<br />

Pharmaceutical<br />

Oliver Lenz - Employment: Janssen; Stock Shareholder: Janssen (J&J)<br />

Gino Van Dooren - Stock Shareholder: Johnson and Johnson<br />

Isabelle Londjon-Domanec - Employment: Janssen<br />

Catherine Nalpas - Employment: Janssen pharmaceuticals<br />

Michael Schlag - Employment: Janssen Cilag Pharma; Stock Shareholder: Johnson<br />

& Johnson<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

The following authors have nothing to disclose: Antonio Craxi<br />

1180<br />

Rapid drop in serum glucose and hypoglycemia in<br />

chronic hepatitis C patients with diabetes during oral<br />

HCV therapy<br />

Laura M. Benítez-Gutiérrez 1 , Vincent Soriano 2 , Isabella Esposito 1 ,<br />

Pablo Barreiro 2 , Ana Maria Duca 1 , Silvia Requena 1 , Isolina<br />

Baños 1 , Ana Treviño 1 , Valentin Cuervas-Mons 1 , Carmen de Mendoza<br />

1 ; 1 Internal Medicine, Hospital Universitario Puerta de Hierro<br />

Majadahonda, Madrid, Spain; 2 Infectious Disease Unit,, La Paz<br />

University Hospital & IdiPAZ, Madrid, Spain<br />

Introduction: Chronic HCV infection is associated with metabolism<br />

abnormalities, including insulin resistance. Up to 15-25%<br />

of chronic hepatitis C patients will develop diabetes lifelong.<br />

New DAA allow HCV suppression within a very short-time<br />

frame. Little is known on the metabolic effects seen in chronic<br />

hepatitis C patients on DAA. Methods: All chronic hepatitis C<br />

patients that received sofosbuvir (SOF)-based regimens during<br />

at least 4 weeks at two reference hepatitis clinics in Madrid<br />

were retrospectively examined. Results: A total of 61 chronic<br />

hepatitis C patients begun SOF-based regimens and had<br />

weeks 0 and 4 controls. Regimens were as follows: SOF+simeprevir<br />

(16), SOF+daclatasvir (13), SOF+ledipasvir (27) and<br />

SOF+ribavirin (5). At 4 weeks 63% had undetectable viremia<br />

and the rest had


794A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and 8( 22.9) received everolimus. 31(24 genotype 1a, 9 genotype<br />

1b) patients completed a 24 week course of therapy,<br />

the remaining 4 patients completed an 8 (genotype 1a),12<br />

(genotype 1a),15 (genotype 1b) and 16 (genotype 1a) weeks<br />

of therapy individually. Side effects were reported in six of the<br />

patients and they included fatigue, somnolence, headache,<br />

anemia, photosensitivity and weight gain. In genotype 1a,<br />

4(16%) patients had RVR, 35 (100%) patients had ETR and<br />

SVR at 4 and 12 weeks. In genotype 1b, 2 (20%) patients had<br />

RVR, 8(80%) patients had ETR and SVR at 4 and 12 weeks.<br />

One of the non responders discontinued treatment before completion<br />

due to decompensation. Overall response rate for genotype<br />

1a and 1b was 94.3%. No patient required adjustment in<br />

IS and no episodes of rejection were documented during treatment.<br />

Conclusion: In patients with recurrent genotype 1 (a or b)<br />

hepatitis C after LT, the combination of once daily sofosbuvir<br />

and simeprevir is generally well tolerated with very high overall<br />

response rates. Discontinuation of treatment only occurred in<br />

two patients due to decompensation of cirrhosis. There were<br />

no interactions that required adjustments in IS.<br />

Disclosures:<br />

Nikolaos Pyrsopoulos - Advisory Committees or Review Panels: GILEAD, BMS,<br />

ABBVIE, VITAL THERAPIES, Quest; Grant/Research Support: ABBVIE<br />

The following authors have nothing to disclose: Nneoma O. Okoronkwo, George<br />

Protopopas, Kiran V. Rao, Vivek A. Lingiah, Maliha Ahmad, Lizza Bojito-Marrero,<br />

Kathleen Ruping, Raj Edula<br />

was very common in both groups (48%). RBV dose reduction<br />

occurred in 8 Asians (26%) and 20 non-Asians (19%). SAE<br />

occurred in 5 non-Asian and all were unrelated to treatment.<br />

Conclusion: Asian and non-Asian CHC-1 patients treated with<br />

SMV+SOF showed similar high treatment response and good<br />

tolerability, despite high rates of advanced disease and prior<br />

treatment failure. Anemia and fatigue were common in both<br />

Asians and non-Asians treated with SOF+RBV.<br />

1182<br />

Similar Tolerability and Effectiveness with SMV+SOF<br />

Therapy in Asians and Non-Asians with Chronic Hepatitis<br />

C Genotype 1 (CHC-1) but Anemia and Fatigue were<br />

Common with SOF+RBV in Both Groups<br />

Christine Y. Chang 1 , Nghia H. Nguyen 2,1 , Changqing Zhao 1,3 ,<br />

Glen A. Lutchman 1 , Aijaz Ahmed 1 , Tami Daugherty 1 , Gabriel<br />

Garcia 1 , Radhka Kumari 1 , W. Ray Kim 1 , Huy N. Trinh 4 , Vinh D.<br />

Vu 1 , Winston Ku 1 , My T. Nguyen 4 , Andrew Huynh 4 , Mindie H.<br />

Nguyen 1 ; 1 Division of Gastroenterology and Hepatology, Stanford<br />

University Medical Center, Palo Alto, CA; 2 Department of<br />

Medicine, University of California, San Diego, San Diego, CA;<br />

3 Institute of Liver Disease, Shugang Hospital, Department of Cirrhosis,<br />

Shanghai, China; 4 San Jose Gastroenterology, San Jose, CA<br />

Purpose: Asian CHC-1 patients can have much higher SVR<br />

with IFN-based therapy but more likely to develop anemia than<br />

non-Asians. Our goal is to investigate for possible ethnic-related<br />

differences in SVR12 or tolerability with DAA-based therapy.<br />

Methods: We evaluated outcomes of DAA-based therapy<br />

(SOF+SMV, SOF+RBV, or SOF+LDV±RBV for 8-24) in 401<br />

patients (330 non-Asians and 71 Asians) seen at 2 US centers<br />

from 12/2013-12/2014. Results: Asians were mostly Vietnamese<br />

(38%) and Chinese (24%). Non-Asians were mostly<br />

Caucasian (62%) and Hispanic (26%). Compared to non-<br />

Asians, Asians were older (mean age: 63±11 vs. 60±9 years,<br />

p=0.004) with lower BMI (26±7 vs. 28±5 kg/m 2 , p=0.01)<br />

and more likely to have HCV-2 (30% vs. 19%) or HCV-6 (8%<br />

vs. 0%), and less likely to have HCV-1 (54% vs. 68%) or 4 (0%<br />

vs. 1%), p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 795A<br />

sis. 14 patients were treated after liver transplantation, 2 of<br />

those with a cholestatic recurrence of HCV, one with advanced<br />

graft failure. 1 patient was treated post-kidney transplantation.<br />

Results: Efficacy: At end of treatment (EOT), 97% (160/165)<br />

were HCV negative, while 89% (92/103) reached SVR 12.<br />

Subgroup analysis: HCV negativity at EOT: SOF/RBV +/- PEG-<br />

IFN: 100% (54/54); SOF, DAC +/- RBV 98% (64/65); SOF,<br />

SMP +/- RBV 90% (27/30); SOF/LED +/- RBV 94% (14/15).<br />

Subgroup analysis for SVR 12: SOF/RBV +/- PEG-IFN: 93%<br />

(40/44); SOF, DAC +/- RBV 87% (27/31); SOF, SMP +/-<br />

RBV 89% (24/27); SOF, LED +/- RBV n.a.. SVR 12 in patients<br />

with cirrhosis: 74% (23/31), SVR 12 in pretreated patients:<br />

85% (44/52), SVR 12 in patients after liver transplantation<br />

75% (6/8). 5 patients discontinued prematurely (3 patients<br />

died during treatment, 2 patients had a virological relapse<br />

after premature discontinuation), 6 patients had a virological<br />

relapse after EOT. No virological breakthrough occurred.<br />

Safety: 1 patient being treated in the course of an advanced<br />

graft failure after liver transplantation died due to complications<br />

of transplant failure and re-transplantation during treatment.<br />

2 patients with advanced cirrhosis died during treatment<br />

(sepsis n=1, right heart failure n=1). Conclusions: New DAA<br />

are highly effective with respect to EOT, and SVR 12. Groups<br />

of patients with previous treatment failure, after liver transplantation,<br />

or cirrhosis achieve significantly higher SVR rates than<br />

with previous treatments. However, in “real-life”, patients with<br />

advanced liver cirrhosis still seem to be “difficult to treat” with<br />

respect to treatment efficacy. Since near all patients will have<br />

reached SVR 12 until September, 2015, uni - and multivariate<br />

analyses will be presented at the AASLD Liver Meeting 2015.<br />

Disclosures:<br />

The following authors have nothing to disclose: Christoph R. Werner, Daniel P.<br />

Egetemeyr, Nisar P. Malek, Ulrich M. Lauer, Christoph P. Berg<br />

1184<br />

Sofosbuvir Combination Therapy and Chronic Hepatitis<br />

C Infection: Is Cultural Background a Predictor of Sustained<br />

Virologic Response?<br />

Amber Tierney, John R. Lake, Amar Mahgoub, Carolyn Schmitt,<br />

Qi Wang, Mohamed A. Hassan; University of Minnesota, Minneapolis,<br />

MN<br />

Background: Treatment for hepatitis C infection has rapidly<br />

evolved over the past few years. Although sustained virologic<br />

response (SVR) rates in clinical trials are high, it is unknown<br />

if this response translates accurately to clinical practice. Our<br />

aim was to detail our initial experience treating patients with<br />

hepatitis C infection with sofosbuvir containing regimens with<br />

an emphasis on genotype 4 infection. Methods: A retrospective<br />

chart review was performed on all patients treated for<br />

hepatitis C infection from December 2013 to May 2015 at a<br />

single academic center with a large population of East African<br />

immigrants. Results: During the study period, 254 patients were<br />

treated for hepatitis C infection. Treatment regimens included<br />

either sofosbuvir/ribavirin (RBV) +/- pegylated interferon (PEG-<br />

IFN), sofosbuvir/simeprevir or ledipasvir/sofosbuvir +/- RBV.<br />

179 patients had 3 month follow-up after treatment with an<br />

overall SVR rate of 79%. Advanced fibrosis was associated with<br />

relapse in genotype 1 infection (p=0.02). Prior treatment, viral<br />

load, and duration of treatment did not have a statistically significant<br />

effect on SVR (p=0.64, 0.40, and 0.23, respectively).<br />

Successful response to treatment was associated with improvement<br />

in alanine and aspartate aminotransferases (p


796A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(14/42) were female, 21% (9/42) were Hispanic, and 9%<br />

(4/42) were Black. The median time from transplant to treatment<br />

initiation was 5.4 years (IQR: 2.1-8.8 years). Thirteen<br />

patients experienced one or more episodes of hepatic decompensation<br />

and/or SAE. Anemia requiring transfusion, the most<br />

common event, occurred in 8/13 (62%), while 7/13 (54%)<br />

decompensated. The cumulative incidence of hepatic decompensation/SAE<br />

was 31% [95% CI, 16-41%]. Risk factors for<br />

decompensation/SAE included pre-treatment hemoglobin [OR<br />

= 0.61 per g/dL; 95% CI: 0.40-0.88; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 797A<br />

All DAAs regimen were analyzed together. For TN and TE<br />

patients, 2 analyses evaluated the RBV use (regardless of the<br />

treatment duration and for a fixed 12 weeks treatment duration)<br />

and 3 analyses evaluated the treatment duration (regardless<br />

of, with, or without RBV). Another analysis compared 12<br />

wks with RBV vs. 24 wks without RBV. Meta-analyses were<br />

performed according to the Der Simonian and Laird method<br />

and reported weight-adjusted SVR12 gains. Results: 10 RCTs<br />

(1307 GT1 cirrhotic patients (573 TN, 734 TE)) were selected.<br />

The distribution of DAA combinations was as follows: SIM+SOF<br />

(COSMOS, N=168), 3D Abbvie (TURQUOISE II, N=380),<br />

DCV-TRIO BMS (UNITY 2, N=202), GPV+EBV (C-WOR-<br />

THY, N=253), and LDV+SOF (6 RCTs, N=496). In G1 TN<br />

cirrhotic patients, the use of ribavirin (ΔSVR=+2.45%;95%CI:<br />

-1.27 to +6.16%,NS,N=411) or an extended duration<br />

(ΔSVR=+0.64%;95%CI:-3.03 to +4.32%,NS,N=434) did not<br />

increase SVR12. In GT1 TE cirrhotic patients, the use of ribavirin<br />

did not increase SVR12 (ΔSVR=+0.23%;95%CI:-3.54<br />

to +4.01%,NS,N=494), even in analyses restricted to<br />

12wks treatment duration (ΔSVR=+1.85%;95%CI:-3.75 to<br />

+7.44%, NS,N=267). Conversely, extended duration was<br />

associated with higher SVR rates (ΔSVR=+6.35%;95%CI:<br />

+1.49 to +11.20%,p=0.018,N=532), even with ribavirin<br />

use (ΔSVR=+8.26%;95%CI: +2.29 to +14.23%, p=0.028,<br />

N=278). The magnitude of SVR gain was similar between the<br />

different DAA regimen suggesting that a 18 wks extended<br />

duration (as performed in the C-WORTHY study) would be<br />

sufficient for other combos. Only four <strong>studies</strong> (229 patients)<br />

compared 12wks +RBV vs. 24wks w/o RBV. Extended duration<br />

was associated with a 5.09% SVR gain that did not reach<br />

significance, probably because of a type II error and the overweight<br />

of the SIRIUS study. Conclusion: In GT1 TN patients with<br />

compensated cirrhosis, RBV as well as extending treatment<br />

duration are useless. In GT1 TE patients with compensated<br />

cirrhosis, RBV does not increase SVR rates but extending treatment<br />

duration (18 or 24 wks) significantly increases SVR rates<br />

Disclosures:<br />

Vincent Di Martino - Advisory Committees or Review Panels: Gilead, France,<br />

Abbvie, BMS France; Board Membership: MSD France; Consulting: Gilead,<br />

France; Speaking and Teaching: Janssen, BMS France, Gilead France<br />

Thong Dao - Board Membership: Schering-Plough, Schering-Plough, Schering-Plough,<br />

Schering-Plough; Speaking and Teaching: Roche, Gilead, Roche,<br />

Gilead, Roche, Gilead, Roche, Gilead<br />

Veronique Loustaud-Ratti - Board Membership: Gilead, Roche, Schering Plough<br />

MSD; Speaking and Teaching: Roche, Schering Plough MSD, Janssen, Bristol<br />

meyers squibb, gilead, Abbvie<br />

Odile Goria - Speaking and Teaching: Gilead, Janssen<br />

Eric Nguyen-Khac - Speaking and Teaching: Gilead, Abbvie, Janssen, Roche,<br />

MSD, BMS<br />

The following authors have nothing to disclose: Carine Richou, Jean Paul Cervoni,<br />

Delphine Weil, Claire Vanlemmens, Anne Minello, Christine Silvain, Thierry<br />

Thevenot<br />

1188<br />

Real World Effectiveness Of Sofosbuvir Based Therapy<br />

In Chronic HCV Infected Patients: The Results From Data<br />

Anlaysis Of 167 Patients In Qatar<br />

Moutaz F. Derbala 1 , Aliaa Amer 2 , Saad R. Alkaabi 1 , Nazeeh<br />

Z. Dweik 1 , Khaleel H. Sultan 1 , Yasser M. Kamel 1 , Khalid Al-Ejji 1 ,<br />

Hamid Wani 1 , Fuad I. Pasic 1 ; 1 GASTROENTEROLOGY & HEPA-<br />

TOLOGY, Hamad Hospitaland Weill Cornell Medical College in<br />

Qatar,Consultant Gastroenterology & Hepatology HMC, Doha,<br />

Qatar; 2 Hematopathology, Laboratory Medicine and pathology,<br />

Hematology Section, Hamad Medical Corporation, Doha, Qatar<br />

Background and Aims: However, limited data were available;<br />

Sofosbuvir (SOF) was approved in Qatar in early 2014. In both<br />

interferon (IFN) based triple therapy and IFN-free combinations,<br />

sofosbuvir have been used to treat different patient populations<br />

with various liver disease severity and comorbidities. Our aim<br />

is to assess the effectiveness of SOF based therapy in a real<br />

world setting. Methods: In the Qatar HCV registry, all patients<br />

who started on the following treatment regimens were included<br />

in the analysis: SOF/ribavirin (RBV), SOF/daclatasvir (DCV),<br />

SOF/simeprevir (SMV), and SOF/PegIFN/RBV. Demographics,<br />

history of liver disease were collected, and laboratory tests<br />

were performed at baseline. Clinic, adverse events and virological<br />

data were collected throughout treatment and post-treatment<br />

follow-up. Results: 167 patients were treated with SOF<br />

based therapy. 36%of the patients had cirrhosis and 52.8%<br />

were previously treated with PR or Protease inhibitor (PI) +PR.<br />

14.1% of the patients underwent liver transplantation.13.8 %<br />

patients were treated with SOF+PegIFN/RBV triple therapy.<br />

The overall SVR4 rate was 96%, 92%, and 87% in patients<br />

treated within genotype 1, 4 and 3 respectively. The overall<br />

SVR4 rate differed according to the prior history and fibrosis.<br />

The SOF based therapy achieved 89% SVR4 rate in cirrhotic<br />

patients and a 97% SVR4 in non-cirrhotic patients. One patient<br />

discontinued all medications due to viral breakthrough with no<br />

discontinuation due to adverse events. The most common side<br />

effects are Anemia and neutropenia, headache and myalgia<br />

in both SOF/RBV and SOF/Peg-IFN/RBV groups. Conclusions:<br />

SOF based Therapy is safe and effective in the treatment of<br />

HCV Genotypes 1, 2, 3 and 4 in a real world setting. However,<br />

both SOF/RBV and SOF/Peg-IFN/RBV regimens are less<br />

reliable in cirrhotic patients and genotype 3<br />

Demographic Data Of The Studied Patients<br />

Disclosures:<br />

The following authors have nothing to disclose: Moutaz F. Derbala, Aliaa Amer,<br />

Saad R. Alkaabi, Nazeeh Z. Dweik, Khaleel H. Sultan, Yasser M. Kamel, Khalid<br />

Al-Ejji, Hamid Wani, Fuad I. Pasic<br />

1189<br />

IFN and/or RBV-Free Therapy (Rx) with Simeprevir+Sofosbuvir<br />

(SMV+SOF) was Well-Tolerated and Effective<br />

for Chronic Hepatitis C Genotype 1 (CHC-1) Including<br />

Post-Liver Transplant (LT) and Decompensated Non-Liver<br />

Transplant (Non-LT) Patients: A Single-Center Experience<br />

Glen A. Lutchman 1 , Nghia H. Nguyen 1,3 , Christine Y. Chang 1 ,<br />

Aijaz Ahmed 1 , Tami Daugherty 1 , Gabriel Garcia 1 , Radhka<br />

Kumari 1 , W. Ray Kim 1 , Soumi Gupta 4 , Dilesh Doshi 2 , Mindie H.<br />

Nguyen 1 ; 1 Division of Gastroenterology and Hepatology, Stanford<br />

University Medical Center, Palo Alto, CA; 2 Health Economics<br />

& Outcomes Research, Janssen Scientific Affairs, Titusville, NJ;<br />

3 Department of Medicine, University of California, San Diego, San<br />

Diego, CA; 4 Medical Affairs, Janssen Scientific Affairs, Titusville,<br />

NJ<br />

Purpose: In the U.S., HCV is the leading cause of liver-related<br />

death and CHC-1 is the most common type. Until recently,<br />

effective and safe Rx options were limited. Our goal is to<br />

evaluate outcomes of SMV+SOF Rx in a diverse single-center<br />

patient cohort from 12/2013/12/2014. Methods: A total of


798A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

234 consecutive patients were identified, of which 36 were<br />

excluded due to prior SOF-based or current IFN-based Rx or<br />

pending SVR12 (n=25), yielding 198 for study inclusion: 50 LT<br />

and 148 non-LT. Results: Mean age was ~60 and the majority<br />

(80%) in both groups were Caucasian or Hispanic. In non-LT,<br />

10% had liver cancer, 71% cirrhosis (84% had decompensation<br />

with clinical symptoms or MELD≥10 or CPT≥7), while only<br />

8% of LT group had cirrhosis. In LT group, SVR12 was similar<br />

for both GT 1a (88%) and 1b (87%), and in cirhrosis (87%)<br />

but lower in those with prior-Rx (73%) (Fig 1). In non-LT group,<br />

SVR12 was higher for GT 1b, noncirrhosis, Rx-naïve compared<br />

to 1a, cirrhosis, and prior-Rx, though none of these differences<br />

were statistically significant (Fig 1). Rx was well tolerated in<br />

all groups with 6 Rx-unrelated SAEs. On multivariate analysis,<br />

also inclusive of sex and age, the presence of at least 2 difficult-to-treat<br />

features (prior Rx, cirrhosis, or GT 1a) was a significant<br />

negative predictor for SVR12 (HR=0.06, CI 0.01-0.73,<br />

p=0.027), but not the presence of just one of such factors or LT.<br />

Conclusions: SMV+SOF Rx (no RBV) was safe and effective for<br />

both LT and non-LT patients with SVR12 of ~80% for those with<br />

difficult-to-treat features and 90% for those without. Significant<br />

predictor for lower SVR12 was presence of 2 or more of such<br />

features.<br />

Disclosures:<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

Gabriel Garcia - Board Membership: Health Metrics Systems<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

Dilesh Doshi - Employment: Janssen<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Glen A. Lutchman, Nghia H.<br />

Nguyen, Christine Y. Chang, Tami Daugherty, Radhka Kumari, Soumi Gupta<br />

1190<br />

Virological and Clinical Response in Patients with<br />

HCV-Related Mixed Cryoglobulinemia Treated with<br />

Interferon-Free Regimens: Preliminary Results of a Prospective<br />

Pilot Study.<br />

Laura Gragnani, Alessia Piluso, Teresa Urraro, Alessio Fabbrizzi,<br />

Elisa Fognani, Luisa Petraccia, Monica Monti, Anna Linda Zignego;<br />

Dept. Experimental and Clinical Medicine, University of<br />

Florence, Firenze, Italy<br />

Background: Hepatitis C virus (HCV) infection is often associated<br />

with extrahepatic manifestations: the most frequent is<br />

represented by mixed cryoglobulinemia (MC). MC is an autoimmune/B-cell<br />

lymphoproliferative disorder, characterized by<br />

the presence of circulating immune complexes, called cryoglobulins.<br />

In 5–30% of cases, MC patients showed symptoms due<br />

to a systemic vasculitis of small/medium size vessels (mixed<br />

cryoglobulinemia syndrome, MCS). The etiologic therapy was<br />

considered the first-line option in MCS patients. However, interferon<br />

(IFN)-based therapy is frequently not tolerated. Recently,<br />

the introduction of direct acting antivirals (DAAs) substantially<br />

changed the treatment of HCV infection, allowing IFN-free<br />

regimens. No data at present exist about the use of IFN-free<br />

regimens in MC patients. Our aim was to analyze the efficacy<br />

and safety of new generation DAAs in IFN-free regimens in MC<br />

patients. Methods: Pilot prospective study, including 17 patients<br />

with HCV-associated MC with or without symptoms treated<br />

with new generation DAAs in IFN-free regimens. Patients were<br />

treated, for 12 or 24 weeks, with different all-oral antiviral<br />

regimen: ombitasvir, paritaprevir+ritonavir and dasabuvir ±<br />

ribavirin; sofosbuvir plus daclatasvir and sofosbuvir plus ribavirin.<br />

Results: Patients were divided in two different cohorts as follows:<br />

(a) 10 patients (3 [27.3%] males, mean age 64.73±6.6<br />

yrs) in the MCS-HCV group; (b) 7 patients (4 [57.1%] males,<br />

mean age 55.57±10.52 yrs) in the MC-HCV group. The main<br />

clinical baseline manifestation of MCS included purpura (80%),<br />

arthralgia (80%), asthenia (100%), peripheral neuropathy<br />

(90%), renal involvement (40%) and cutaneous ulcers (30%).<br />

Regarding previous anti-HCV treatment, 7 (41%) patients were<br />

naïve, 6 (35%) were partial or non responders, 1 (6%) patient<br />

was virological relapser and 3 (18%) discontinued previous<br />

treatment for severe adverse events. At week 8 of treatment, all<br />

patients were HCV RNA negative. Furthermore, a reduction of<br />

the cryocrit values was evident in all cases (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 799A<br />

1191<br />

Evaluation of the efficacy and safety to dual oral therapy<br />

with daclatasvir and asunaprevir in Japanese reallife<br />

settings<br />

Hitomi Sezaki 1 , Fumitaka Suzuki 1 , Tetsuya Hosaka 1 , Shunichiro<br />

Fujiyama 1 , Hideo Kunimoto 1 , Yushi Sorin 1 , Yusuke Kawamura 1 ,<br />

Norio Akuta 1 , Masahiro Kobayashi 1 , Yoshiyuki Suzuki 1 , Satoshi<br />

Saitoh 1 , Yasuji Arase 1 , Kenji Ikeda 1 , Mariko Kobayashi 2 , Hiromitsu<br />

Kumada 1 ; 1 Hepatology, Toranomon Hospital, Tokyo, Japan;<br />

2 Research institute for hepatology, Toranomon Hospital, Kawasaki,<br />

Japan<br />

Aim: The first IFN-free dual oral therapy in Japan with daclatasvir<br />

(DCV) and asunaprevir (ASV) for patients with chronic<br />

HCV-1b infection was approved since Sep. 2014. Since the<br />

real-world population is heterogeneous, it is important to indicate<br />

treatment outcomes in patients excluded from clinical trials<br />

(CTR). The aims of this study were to evaluate the efficacy<br />

and safety to 24-week DCV/ASV therapy in Japanese reallife<br />

settings. Methods: 604 GT1b patients started dual oral<br />

therapy with DCV/ASV until Jan. 31, 2015 in our hospital.<br />

423 patients achieved end of treatment (EOT) until May 16.<br />

180 of 423 patients met Phase3-CTR inclusion criteria and<br />

243 patients unmet it. This criteria is as following; age ≤ 75<br />

yr, absence of HCC, Hb ≥ 8.5g/dL, platelet counts ≥ 50 x<br />

10 3 /mL, albumin ≥ 3.5 g/dL, HCV RNA levels ≥ 5 logIU/mL,<br />

and absence of prior DAAs therapy. We evaluated SVR rates<br />

after treatment and the adverse events (AE) during therapy.<br />

Results: CTR-unmet patients were older than CTR-met patients<br />

(median 73 yr vs. 65 yr, p


800A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1193<br />

ITPA polymorphisms are predictors of anemia but no<br />

contribution to outcomes in chronic hepatitis C patients<br />

who treated with interferon plus ribavirin<br />

Xiumei Chi 1 , Mo-li Wang 3 , Yu Pan 1 , Jing Jiang 2 , Hongqing Yan 1 ,<br />

Ruihong Wu 1 , Xiaomei Wang 1 , Junqi Niu 1 ; 1 Hepatology, The first<br />

hospital of Jilin University, Changchun, China; 2 Epidemoology,<br />

The first hospital of Jilin University, Changchun, China; 3 The fourth<br />

hospital of Jilin University, Changchun, China<br />

Background and Aims: In a genome-wide association study of<br />

patients being treated for chronic hepatitis C (HCV), two single<br />

nucleotide polymorphisms (SNPs) within or adjacent to the<br />

inosine triphosphatase (ITPA) gene (rs1127354, rs6051702)<br />

were shown to protect against ribavirin induced hemolytic anemia<br />

during early stages of treatment. We then evaluated these<br />

polymorphisms/variants as predictors of anemia or treatment<br />

outcomes (SVR) in a large cohort of naive HCV patients who<br />

were treated with interferon plus ribavirin. Methods: Genetic<br />

material was available from 351 HCV patients. These patients<br />

were enrolled in the first hospital of Jilin University who underwent<br />

interferon plus ribavirin combination therapy and they had<br />

no history of any other interferon-based therapy. rs6051702<br />

in C20orf194, rs1127354 in ITPA and rs8099917 in IL-28B<br />

were genotyped and tested for their association with hemoglobin<br />

(Hb) reduction during treatment and with treatment<br />

outcomes. A stepwise multivariate regression analysis was<br />

performed to identify factors associated with outcome of the<br />

therapy. Results: The ITPA rs1127354 minor variants were<br />

significantly associated with protection against reduction Hb<br />

level at week 4 (p=2.65×10 -4 during HCV combination therapy.<br />

SNP rs6051702 was not associated with the hemoglobin<br />

decline to >3 g/dL at week 4 in our study (p=0.861).<br />

rs1127354 was strongly and independently associated with<br />

protection from quantitative Hb reduction at week 4. The minor<br />

alleles of each variant protected against HB reduction. The<br />

ITPase deficiency genotype was associated with lower rates of<br />

anemia over the entire treatment period (72 weeks). No association<br />

with sustained virological response (SVR) was observed.<br />

Conclusion: ITPA polymorphisms influences hemoglobin levels<br />

during ribavirin combination therapy. rs1127354 was conrmed<br />

to be a useful predictor of ribavirin-induced anemia in<br />

HCV patients but not a predictive factor for treatment outcomes.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xiumei Chi, Mo-li Wang, Yu Pan,<br />

Jing Jiang, Hongqing Yan, Ruihong Wu, Xiaomei Wang, Junqi Niu<br />

1194<br />

Cardiac arrhythmia in patients with chronic hepatitis C<br />

virus infection treated with sofosbuvir-including regimen<br />

Hélène Fontaine 1,2 , Arnaud Lazarus 3,4 , Caroline Pecriaux 5 ,<br />

François Bagate 3 , Philippe Sultanik 1,2 , Estelle Boueyre 1,2 , Marion<br />

Corouge 1,2 , Vincent Mallet 1,2 , Anaïs Vallet-Pichard 1,2 ,<br />

Philippe Sogni 1,2 , Denis Duboc 3 , Stanislas Pol 1,2 ; 1 Hepatology<br />

Unit, Cochin Hospital, Paris, France; 2 Université Paris Descartes,<br />

INSERM USM20, Institut Pasteur et Assistance Pubique - Hôpitaux<br />

de Paris, Paris, France; 3 Cardiology Unit, Hôpital Cochin, Assistance<br />

Publique -Hôpitaux de Paris, Paris, France; 4 Rhythmology<br />

Unit, Clinique Ambroise Paré, Neuilly Sur Seine, France; 5 Unité de<br />

Pharmacovigilance, Hôpital Cochin, Assistance Publique - Hôpitaux<br />

de Paris, Paris, France<br />

Introduction: Treatment of chronic hepatitis C virus (HCV)<br />

infection has dramatically improved since the advent of direct<br />

anti-viral agents, including Sofosbuvir. Sustained virological<br />

response, which corresponds to a viral cure, is achievable<br />

in more than 90 % of patients. To date, several thousands of<br />

patients have been treated with Sofosbuvir-based regimens<br />

worldwide. Since their approval in France, oral anti-viral therapies<br />

were prescribed for patients with complications, including<br />

cirrhosis and life-threatening extra-hepatic manifestations.<br />

Methods: We analyzed prospectively the incidence of arrhythmias<br />

and conduction disorder in a cohort of patients treated<br />

by Sofosbuvir (new antiviral HCV) in our hospital from January<br />

2, 2014 to December, 31, 2014. Results: We observed five<br />

cases (1.2 %) of severe arrhythmia, including three cardiac<br />

conduction defects, within the first days of Sosfobuvir-based<br />

therapy among 415 patients who had been treated in our<br />

unit. There were two females and three males. The median<br />

age was 58 (50-75) years. Two patients were co-infected with<br />

HIV. Arrhythmia relapsed when Sofosbuvir was reintroduced in<br />

one patient. Among the 3 conduction disorders, two patients<br />

were treated with beta-blocker and another with amiodarone.<br />

A pacemaker was implanted in all 3 patients with conduction<br />

disorders. The three patients who underwent a pacemaker<br />

implantation were not pacemaker-dependent on the controls<br />

after stopping antiviral treatments. Conclusion: Sofosbuvir<br />

seems to have cardiac toxicity. The mechanisme underlying<br />

Sofosbuvir cardiotoxicity has to be determined. A particular<br />

attention should be given to patients with a medical history of<br />

rhythm or conduction abnormalities, including careful monitoring<br />

(e.g., Holter ECG recordings) of cardiac rhythm during the<br />

first days of Sofosbuvir therapy.<br />

Disclosures:<br />

Hélène Fontaine - Board Membership: Gilead, BMS, Janssen, Abbvie; Speaking<br />

and Teaching: Gilead, BMS, Janssen, Abbvie, MSD<br />

Arnaud Lazarus - Board Membership: Boston Scientific, Medtronic; Consulting:<br />

Sorin Group; Employment: Biotronik<br />

Anaïs Vallet-Pichard - Independent Contractor: Schering Plough, Gilead, BMS,<br />

Roche<br />

Philippe Sogni - Board Membership: BMS, Gilead; Consulting: Gilead, Roche,<br />

MSD, BMS, Janssen, Mayoli-Spindler<br />

Stanislas Pol - Board Membership: Sanofi, Bristol-Myers-Squibb, Boehringer Ingelheim,<br />

Tibotec Janssen Cilag, Gilead, Glaxo Smith Kline, Roche, MSD, Novartis;<br />

Grant/Research Support: Glaxo Smith Kline, Gilead, Roche, MSD; Speaking and<br />

Teaching: Sanofi, Bristol-Myers-Squibb, Boehringer Ingelheim, Tibotec Janssen<br />

Cilag, Gilead, Glaxo Smith Kline, Roche, MSD, Novartis<br />

The following authors have nothing to disclose: Caroline Pecriaux, François<br />

Bagate, Philippe Sultanik, Estelle Boueyre, Marion Corouge, Vincent Mallet,<br />

Denis Duboc<br />

1195<br />

Utility of acoustic radiation force impulse (ARFI) for predicting<br />

treatment response of patients with chronic hepatitis<br />

C receiving triple therapy with protease inhibitor<br />

plus pegylated interferon and ribavirin<br />

Ryoko Yamada 1 , Naoki Hiramatsu 1 , Tsugiko Oze 1 , Ayako Urabe 1 ,<br />

Yuki Tahata 1 , Naoki Morishita 1 , Takayuki Yakushijin 1 , Yuichi<br />

Yoshida 1 , Tomohide Tatsumi 1 , Yasuharu Imai 2 , Tetsuo Takehara 1 ;<br />

1 Department of Gastroenterology and Hepatology, Osaka University<br />

Graduate School of Medicine, Suita, Japan; 2 Ikeda Municipal<br />

Hospital, Ikeda, Japan<br />

Background & Aim: With interferon-based therapy, liver fibrosis<br />

has been reported to correlate with the virologic response<br />

in patients with chronic hepatitis C (CH-C). Acoustic radiation<br />

force impulse (ARFI) is known as a non-invasive procedure for<br />

assessment of liver fibrosis. In this study, we investigated the<br />

utility of ARFI for CH-C patients receiving triple therapy with<br />

protease inhibitor (PI) plus pegylated interferon (Peg-IFN) and<br />

ribavirin (RBV). Patients & Methods: A total of 105 genotype 1<br />

CH-C patients were enrolled. They started triple therapy (telaprevir/simeprevir,<br />

26/79) between February 2011 and June<br />

2014, following liver biopsy and ARFI assessment. The base-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 801A<br />

line characteristics were as follows: age, 60.9 ± 10.0 years<br />

old; median HCV RNA, 6.6 log IU/mL; platelet count, 15.8 ±<br />

5.1 x10 4 /mm 3 ; ALT, 56.4 ± 42.0 IU/L; Vs value, 1.47 ± 0.46<br />

m/sec; Fibrosis score (F1/2/3/4), 55/18/15/10. Single<br />

nucleotide polymorphisms located near the IL28B (rs8099917)<br />

gene were determined using a real-time PCR system. Results:<br />

Vs values at baseline showed positive correlation with Fibrosis<br />

score (p


802A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

wave PI, telaprevir (TVR)-based triple therapy of patients with<br />

genotype 1b HCV infection. Methods: In this study, 23 patients<br />

infected with HCV genotype 1b who had not achieved sustained<br />

virologic response (SVR) after TVR-based therapy (17<br />

relapsers, 1 non-responder and 5 patients who discontinued<br />

treatment) received SMV-based therapy at Osaka University<br />

Hospital and institutions participating in the Osaka Liver Forum.<br />

Among them, 19 patients had experienced Peg-IFN plus RBV<br />

therapy (9 relapsers and 10 non-responders) before TVR-based<br />

therapy. Resistance testing of the NS3 region was performed<br />

by ultra-deep sequencing at the start of SMV-based therapy.<br />

Results: In per-protocol, the SVR12 rates were 52% (11/21), 2<br />

patients were non-responders and 8 patients relapsed. The only<br />

factor that was significantly associated with SVR was the virologic<br />

response to Peg-IFN plus RBV (relapser vs. non-responder,<br />

OR: 0.048, p = 0.024, SVR rates: 88% vs 25%) in univariate<br />

analysis, while that to TVR-based therapy had no association<br />

with SVR. Ultra-deep sequencing at the start of SMV-based<br />

therapy revealed that substitution at the R155 position, which<br />

is cross-resistant to TVR and SMV, was detected in only one<br />

patient with low frequency (R155G, 0.4%) and this patient<br />

achieved SVR. RAVs of D168 at the start of SMV-based therapy<br />

were detected in four patients with low frequency (~ 1%). The<br />

SVR rates of patients with and without RAVs of D168 were 50%<br />

(2/4) and 53% (9/17). In two non-response patients for SMVbased<br />

therapy, emergent TVR-resistant variants, V36G (92.2%)<br />

and T54A (82.9%), were undetectable at the start of SMVbased<br />

therapy, but de novo variants of D168 were detected<br />

during the therapy. Conclusions: In this study, no correlation<br />

was observed between baseline RAVs and the effect of SMVbased<br />

therapy. The efficacy of retreatment with SMV-based<br />

therapy after failure of TVR-based therapy was suggested to<br />

depend on previous virologic response to Peg-IFN plus RBV.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Naoki Morishita, Naoki Hiramatsu,<br />

Tsugiko Oze, Yuki Tahata, Ryoko Yamada, Takayuki Yakushijin, Akira<br />

Yamada, Eiji Mita, Toshihiko Nagase, Yukinori Yamada, Toshifumi Ito, Masahiko<br />

Tsujii, Masami Inada, Yasuharu Imai, Ryotaro Sakamori, Yuichi Yoshida,<br />

Tomohide Tatsumi, Norio Hayashi<br />

1198<br />

Better Work Productivity and Activity in Patients on<br />

Ombitasvir/ Paritaprevir/Ritonavir and Dasabuvir with<br />

or without Ribavirin (3D+RBV or 3D) in Treatment-Naïve<br />

Adults with Genotype 1(GT1) Chronic Hepatitis C<br />

Yan Liu 1 , Yan Luo 1 , Xuan Liu 1 , Danielle Sullivan 1 , Timothy R.<br />

Juday 1 , Brian Conway 2 ; 1 Abbvie Inc, North Chicago, IL; 2 Vancouver<br />

Infectious Diseases Centre, Vancouver, BC, Canada<br />

Background Interferon-containing regimens for treatment of<br />

Chronic Hepatitis C can further worsen the work productivity<br />

and activity impairment (WPAI) of patients. We report the<br />

WPAI of the 311 patients who participated in the phase 3b<br />

MALACHITE-I study. Methods GT1a-infected patients were randomized(2:1)<br />

to receive 12 weeks of 3D+RBV or 12 weeks<br />

of Telaprevir plus peginterferon/ribavirin(TPV+pegIFN+RBV)<br />

and 12-36 additional weeks of pegIFN/RBV. GT1b-infected<br />

patients were randomized(2:2:1) to 3D+RBV, 3D, or TPV+pegIFN/RBV.<br />

Baseline and change from baseline during treatment<br />

and post treatment(PT) periods were measured using<br />

the WPAI instrument (hepatitis C module, WPAI-HCV) for all<br />

five treatment groups The statistical significance of differences<br />

between treatment groups in mean change from baseline to<br />

Week 12 and to PT Week 12 was assessed by subgenotype,<br />

respectively. Results The analyses included 69 patients on<br />

3D+RBV and 34 on TPV+pegIFN/RBV in GT1a, and 84 on<br />

3D+RBV, 83 on 3D and 41 on TPV+pegIFN/RBV in GT1b.<br />

Mean baseline overall work impairment(OWI) and activity<br />

impairment scores for Week 12 and PT Week 12 analyses<br />

varied from 2.5 to 9.1 and 9.3 to 13.4 across treatment<br />

groups. Mean changes from baseline in WPAI-HCV scores<br />

are summarized in the Table. Numerically smaller additional<br />

impairments were observed in OWI at Week 12 with 3D+RBV<br />

in GT1b(p=0.004) and GT1a(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 803A<br />

pharmacokinetics <strong>studies</strong>,H2-blocker famotidine (40 mg q12<br />

hours) & PPI omeprazole (20 mg single dose) decreased LDV<br />

Cmax by 17-20% & 4-11% respectively but had no effect on<br />

AUC 1 . In the ION trials, which evaluated the effect of SOF/<br />

LDV in genotype 1 patients, patients were not allowed to be<br />

on any acid-suppressing agents for 28 days before initiation of<br />

therapy. Hence, currently there is no clinical data on the effect<br />

of acid suppressing agents on SVR in patients treated with<br />

SOF/LDV. Aim: To evaluate the effect of co-administration of<br />

PPI (at doses recommended in the package insert) with LDV/<br />

SOF on SVR12 in hepatitis C genotype 1 patients. Methods:<br />

Data from a combined treatment cohort of three Hepatology<br />

referral centers were obtained on all hepatitis C genotype1<br />

patients on PPI treated with SOF/LDV with or without ribavirin<br />

(RBV). Treatment duration (8-24 weeks) & addition of RBV was<br />

at the discretion of the treating hepatologist Results: Of the 50<br />

patients in this cohort,30 (60%) were male, 35 (70%)had genotype<br />

1a, 29 (58%) were Caucasians, & 68% had cirrhosis.<br />

Mean age was 58 years, with a mean BMI of 31 kg/m 2 . 25<br />

(50%) were treatment naïve & the remaining were treatment<br />

experienced with peg-interferon (peg-IFN) + RBV with or without<br />

Telaprevir/Bocepravir, peg-IFN + RBV + SOF, Simeprevir<br />

+SOF with or without RBV, SOF+RBV. PPI was co-administered<br />

at the recommended package insert dosing primarily<br />

for GERD (86%). 22/43 had undetectable viral load after 4<br />

weeks of treatment initiation, & 92% (22/24) who completed<br />

treatment achieved end-of treatment (EOT) response defined by<br />

undetectable viral load. Treatment was very well tolerated with<br />

treatment discontinuation in only one patient due to intractable<br />

nausea & vomiting requiring hospitalization. SVR12 results to<br />

follow Conclusion: The co-administration of a PPI with LDV/<br />

SOF at recommended doses does not seem to effect EOT<br />

response. Symptoms of GERD are also generally well tolerated<br />

at the package insert PPI dosing while on treatment 1 German<br />

P, et al. Effect of food & acid reducing agents on the relative<br />

bioavailability & pharmacokinetics of ledipasvir/sofosbuvir<br />

fixed dose combination tablet.15th International Workshop on<br />

Clinical Pharmacology of HIV & Hepatitis Therapy; Washington,<br />

D.C. 2014<br />

Disclosures:<br />

Hector E. Nazario - Advisory Committees or Review Panels: Gilead, Janssen;<br />

Speaking and Teaching: Gilead, Merck, Abbvie, Janssen<br />

Manjushree Gautam - Speaking and Teaching: Gilead<br />

Stevan A. Gonzalez - Consulting: AbbVie; Speaking and Teaching: Gilead,<br />

Salix, AbbVie<br />

Ruben Ramirez-Vega - Speaking and Teaching: Gilead<br />

Apurva A. Modi - Speaking and Teaching: Salix, Merck, Gilead, Abbvie, Janssen<br />

The following authors have nothing to disclose: Jamil S. Alsahhar, Mohammad<br />

Ashfaq, Milka Ndungu, Olga Teachenor, Gabriel R. Gonzales<br />

1200<br />

Pre-Existing Co-Morbidities and Co-Medications of<br />

Patients Undergoing Treatment of Chronic HCV G1<br />

Infection in German Real-Life<br />

Peter Buggisch 1 , Hanns F. Loehr 2 , Gerlinde Teuber 3 , Hermann<br />

Steffens 4 , Michael R. R. Kraus 5 , Peter R. Geyer 6 , Bernd Weber 7 ,<br />

Thomas Witthoeft 8 , Uwe Naumann 9 , Elmar Zehnter 10 , Dagmar<br />

Hartmann 11 , Bernd Dreher 11 , Manfred Bilzer 11 ; 1 IFI Liver Center<br />

Hamburg, Hamburg, Germany; 2 Gastroenterological Practice,<br />

Wiesbaden, Germany; 3 Gastroenterological Practice, Frankfurt,<br />

Germany; 4 Practice of Internal Medicine, Berlin, Germany;<br />

5 Department II, Klinikum Burghausen, Burghausen, Germany;<br />

6 Gastroenterological Practice, Fulda, Germany; 7 Competence<br />

Center Addiction, Kassel, Germany; 8 Gastroenterological Practice,<br />

Stade, Germany; 9 Center of Medicine, Berlin, Germany; 10 Gastroenterological<br />

Practice, Dortmund, Germany; 11 MSD Pharma<br />

GmbH, Haar, Germany<br />

Background: Information about co-morbidities of patients (pts)<br />

currently treated for chronic HCV genotype 1 (G1) infection in<br />

real-life is scarce. The present interim analysis of the NOVUS<br />

observational study was therefore aimed to investigate the<br />

frequency of pre-existing and ongoing co-morbidities of pts<br />

treated for chronic HCV G1 infection in German real-life and<br />

to determine the frequency of co-medications. Methods: From<br />

April 2012 until January 2014, 536 pts with HCV G1 infection<br />

were recruited in the ongoing NOVUS study by 97 practices<br />

and hospitals in Germany. Until now, pre-existing co-morbidities<br />

before triple therapy of HCV G1 infection with boceprevir<br />

(BOC) were documented for 469 pts. Results: Ongoing<br />

co-morbidities were reported for 329 of 469 pts (70%) before<br />

treatment of HCV G1 infection. Overall, 599 ongoing co-morbidities<br />

(multiple answers allowed) were documented. The most<br />

frequently reported co-morbidities were obesity (BMI >30 kg/<br />

m 2 ) (19%), cardiovascular diseases (18%), psychiatric disorders<br />

(14%), opiate substitution (13%), gastrointestinal diseases<br />

(11%), metabolic disorders (8%), thyroid gland diseases (7%),<br />

bone and joint diseases (6%), skin diseases (5%), HIV- co-infection<br />

(4%) and kidney diseases (2%). When co-morbidities<br />

were analyzed by gender (female vs. male), thyroid diseases<br />

occurred more frequently in females (13% vs. 3%, P


804A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Michael R. R. Kraus - Advisory Committees or Review Panels: Merck/MSD,<br />

Roche, Gilead, BMS, Janssen, ABBVIE; Consulting: Merck/MSD, Roche; Speaking<br />

and Teaching: Merck/MSD, Gilead, BMS, Janssen, ABBVIE<br />

Bernd Weber - Consulting: Mundipharma, Indivior, Hexal, Janssen, Bristol Myers<br />

Squibb, AbbVie, MSD, Gilead Sciences<br />

Uwe Naumann - Speaking and Teaching: MSD, Roche, BMS, Abbott, VIIV, Janssen,<br />

Boehringer Ingelheim, Gilead<br />

Dagmar Hartmann - Employment: MSD Germany<br />

Bernd Dreher - Employment: MSD<br />

Manfred Bilzer - Consulting: MSD Germany<br />

The following authors have nothing to disclose: Hanns F. Loehr, Hermann Steffens,<br />

Peter R. Geyer, Thomas Witthoeft, Elmar Zehnter<br />

1201<br />

High ribavirin dose and extended treatment can raise<br />

sustained virologic response in chronic hepatitis C genotype<br />

1 patients treated with simeprevir, pegylated interferon<br />

plus ribavirin<br />

Yuki Tahata 1 , Naoki Hiramatsu 1 , Tsugiko Oze 1 , Ayako Urabe 1 ,<br />

Naoki Morishita 1 , Ryoko Yamada 1 , Takayuki Yakushijin 1 , Yukiko<br />

Saji 2 , Yoshinori Doi 3 , Hideki Hagiwara 4 , Akira Kaneko 5 , Hiroyuki<br />

Fukui 6 , Masami Inada 7 , Kazuhiro Katayama 8 , Atsuo Inoue 9 , Yuichi<br />

Yoshida 1 , Tomohide Tatsumi 1 , Norio Hayashi 4 , Tetsuo Takehara 1 ;<br />

1 Department of Gastroenterology, Osaka University Graduate<br />

School of Medicine, Suita, Japan; 2 Itami City Hospital, Itami,<br />

Japan; 3 Otemae Hospital, Osaka, Japan; 4 Kansai Rosai Hospital,<br />

Amagasaki, Japan; 5 NTT West Osaka Hospital, Osaka, Japan;<br />

6 Yao Municipal Hospital, Yao, Japan; 7 Toyonaka Municipal Hospital,<br />

Toyonaka, Japan; 8 Osaka Medical Center for Cancer and<br />

Cardiovascular Diseases, Osaka, Japan; 9 Osaka General Medical<br />

Center, Osaka, Japan<br />

Background and Aim: Combination therapy with simeprevir<br />

(SMV), pegylated interferon (Peg-IFN) plus ribavirin (RBV) is<br />

recommended as first-line interferon-based therapy for patients<br />

with chronic hepatitis C (CH-C) genotype 1 in Japan. However,<br />

the factors associated with a sustained virologic response (SVR)<br />

have not been fully examined. In this study, we investigated<br />

factors associated with SVR in patients with CH-C treated with<br />

SMV, Peg-IFN plus RBV combination therapy. Patients and<br />

Methods: This multicenter study was conducted at Osaka University<br />

Hospital and institutions participating in the Osaka Liver<br />

Forum. A total of 531 patients with CH-C treated with SMV,<br />

Peg-IFN plus RBV were enrolled in this study (244 males, 287<br />

females; mean age: 62.2 ± 10.0 y.o.; 229 naïve patients, 212<br />

Peg-IFN plus RBV-experienced patients). SVR was evaluated by<br />

HCV-RNA negativity at 12 weeks after the end of treatment<br />

(SVR12). As a rule, naïve patients and prior relapsers were<br />

treated for 24 weeks, and patients with non-response (NR) to<br />

previous therapies were treated for 24 or 48 weeks (24 weeks:<br />

64 patients, 48 weeks: 23 patients). Results: The discontinuance<br />

rate due to adverse effect was 8.0% (42/523). In per-protocol,<br />

the SVR12 rates were 87% for naïve patients, 95% for<br />

relapsers and 42% for non-responders. Among naïve patients,<br />

IL28B genotype (rs8099917, TT vs. non-TT, OR: 0.044, p =<br />

0.001) and RBV dose (< 10/10-12/≥ 12 mg/kg/day, OR:<br />

4.513, p = 0.041) were significantly associated with SVR on<br />

multivariate analysis including baseline factors and drug adherence.<br />

In patients with IL28B TT genotype, the SVR12 rates were<br />

high regardless of RBV dose (93-100%), while in patients with<br />

IL28B non-TT genotype, stratified analysis of RBV dose showed<br />

that SVR12 rates were affected by the RBV dose (p = 0.046),<br />

being 44% (8/18) for patients given < 10 mg/kg/day of RBV,<br />

78% (14/18) for those given 10-12 mg/kg/day of RBV and<br />

100% (3/3) for those given ≥ 12 mg/kg/day of RBV. In NR<br />

patients with HCV-RNA negativity at 12 and 24 weeks, the<br />

SVR12 rates were significantly higher in patients treated for<br />

48 weeks than those treated for 24 weeks (88%, 15/17 vs.<br />

58%, 25/43; p = 0.03). As for SVR12 rates according to the<br />

timing of HCV-RNA disappearance, the SVR12 rates of 24-<br />

and 48-week treatment were 91% (10/11) and 100% (1/1) in<br />

patients with undetectable HCV-RNA at 2 weeks, 38% (9/24)<br />

and 89% (8/9) at 4 weeks, 43% (6/14) and 83% (5/6) at<br />

5-12 weeks, respectively. Conclusion: Higher RBV dose for<br />

naïve patients with unfavorable IL28B genotype and extended<br />

treatment for NR patients can raise SVR in CH-C genotype<br />

1 patients treated with SMV, Peg-IFN plus RBV combination<br />

therapy.<br />

Disclosures:<br />

Kazuhiro Katayama - Consulting: Nobel pharma; Speaking and Teaching:<br />

Bayer, MSD, Otsuka pharmaceutical Co.,Ltd., Glaxo, Zeria pharmaceutical Co.,<br />

Ltd., Bristol Meyers<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yuki Tahata, Naoki Hiramatsu,<br />

Tsugiko Oze, Ayako Urabe, Naoki Morishita, Ryoko Yamada, Takayuki Yakushijin,<br />

Yukiko Saji, Yoshinori Doi, Hideki Hagiwara, Akira Kaneko, Hiroyuki Fukui,<br />

Masami Inada, Atsuo Inoue, Yuichi Yoshida, Tomohide Tatsumi, Norio Hayashi<br />

1202<br />

Hepatitis C Cure Results in a Modest, but Durable,<br />

Decrease in Liver Stiffness<br />

Sweta Chekuri, Kian Bichoupan, Sanders Chang, Neal M. Patel,<br />

Ponni V. Perumalswami, David P. Del Bello, Rachna Yalamanchili,<br />

Alyson Harty, Douglas Dieterich, Andrea D. Branch; Division of<br />

Hepatology, Mount Sinai, New York, NY<br />

Background & Aims: We investigated the impact of HCV cure<br />

on liver stiffness (LS). Methods: FibroScan was used to measure<br />

LS of 61 patients treated with an interferon (IFN)-containing or<br />

IFN-free regimen (2008-2015) who achieved a sustained virological<br />

response (SVR24). Scores ≥ 14 kPa denoted cirrhosis.<br />

Mann-Whitney and Wilcoxon signed-rank tests were used for<br />

analysis. Results: The median age was 60 years (± 11). The<br />

median pre-treatment LS score was 9.1 kPa (IQR: 6.1 - 15 kPa),<br />

the median post-SVR score was 7.3 kPa (IQR: 5.3 – 10 kPa),<br />

and the median change was -2.1 kPa (IQR: -5.3 – -0.4 kPa), a<br />

modest, but statistically significant decrease, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 805A<br />

C trough<br />

levels were higher in patients with more advanced liver<br />

fibrosis; they correlated particularly strongly with type 4 collagen<br />

7S, WFA + -M2BP, and FIB-4 levels. ASV C trough<br />

prediction<br />

of hepatic transaminase elevations or VBT was difficult, and<br />

elevated ASV C trough<br />

levels were not observed even in patients<br />

with compromised renal function.<br />

Disclosures:<br />

Itaru Ozeki - Speaking and Teaching: BMS<br />

Disclosures:<br />

Kian Bichoupan - Consulting: Janssen Pharmaceuticals, Gilead Sciences<br />

Alyson Harty - Advisory Committees or Review Panels: Gilead; Consulting: Gilead,<br />

Jannsen, Acaria Pharmacy, Abbvie<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

The following authors have nothing to disclose: Sweta Chekuri, Sanders Chang,<br />

Neal M. Patel, Ponni V. Perumalswami, David P. Del Bello, Rachna Yalamanchili<br />

1203<br />

Clinical Significance of Blood Asunaprevir Concentrations<br />

Levels during Combined Daclatasvir and Asunaprevir<br />

Therapy<br />

Itaru Ozeki; Department of Gastroenterology, Sapporo Kosei General<br />

Hospital, Sapporo, Japan<br />

[Objectives] This study evaluated the clinical significance of<br />

blood asunaprevir (ASV) concentrations levels during combined<br />

daclatasvir+ASV therapy in patients with genotype 1<br />

chronic hepatitis C. [Subjects and Methods] The study included<br />

116 patients (36 males [31.0%]; median age, 73 years) starting<br />

this combination therapy. Liver cirrhosis (LC) was present<br />

in 62 (53.4%) patients. The median estimated glomerular filtration<br />

rate (eGFR) was 65.5 mL/min/1.73 m 2 . ASV C trough<br />

levels<br />

were analyzed by background variables, and then analyzed<br />

according to the presence/absence of hepatic transaminase<br />

elevations, virological breakthrough (VBT), and renal dysfunction.<br />

Liver fibrosis was determined using Fibroscan ® EI values as<br />

well as FIB-4 and Wisteria floribunda agglutinin-positive human<br />

Mac-2-binding protein (WFA + -M2BP) levels. Serum WFA + -<br />

M2BP levels were measured with HISCL ® M2BPGi reagent (Sysmex<br />

Co., Kobe, Japan). ASV C trough<br />

levels were measured using<br />

HPLC in patients receiving blood sampling 8-12 hours after<br />

oral administration of ASV 1 week after starting ASV treatment.<br />

Hepatic transaminase elevations was defined as cases higher<br />

than twice the upper limit of the normal range. Each parameter<br />

was expressed as the median. [Results] 1) Analysis of ASV<br />

C trough<br />

levels by background variables revealed that the levels<br />

did not vary with age, gender, body mass index, or eGFR, but<br />

it was significantly higher in the LC group (119 ng/mL; n=62)<br />

than in the non-LC group (52 ng/mL; n=54) (p=0.001) . 2)<br />

Analysis of the correlation of ASV C trough<br />

levels with factors possibly<br />

associated with liver fibrosis (platelet, albumin, hyaluronic<br />

acid, type 4 collagen 7S, FIB-4, WFA + -M2BP, and FS EI values)<br />

revealed that all were significant (p


806A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

The following authors have nothing to disclose: Bernhard Scheiner, Philipp<br />

Schwabl, Sebastian Steiner, David Chromy, Theresa Bucsics, Berit A. Payer,<br />

Maximilian C. Aichelburg, Armin Rieger, Katharina Grabmeier-Pfistershammer<br />

1205<br />

Real-world effectiveness of ledipasvir/sofosbuvir 8<br />

weeks chronic hepatitis C treatment<br />

Peter Buggisch 1 , Karsten Wursthorn 1 , Albrecht Stoehr 1 , Aline<br />

Gauthier 2 , Petar K. Atanasov 2 , Joerg Petersen 1 ; 1 Asklepios Klinik<br />

St. Georg Haus L, IFI Institut für Interdisziplinäre Medizin, Hamburg,<br />

Germany; 2 Amaris Consulting, London, United Kingdom<br />

Background and Aims: Ledipasvir/Sofosbuvir (LDV/SOF) single<br />

tablet regimen (STR) is approved in Europe for chronic<br />

hepatitis C (CHC) patients with genotypes (GT) 1, 3 and 4. In<br />

the ION-3 study 8 weeks (8w) of LDV/SOF was non-inferior<br />

to 12w in previously untreated GT1 patients without cirrhosis<br />

with no benefit for the addition of ribavirin (RBV). According<br />

to the summary of product characteristics (SmPC) 8w may<br />

be considered in treatment naïve, non-cirrhotic patients with<br />

HCV RNA< 6million IU/mL. The aim of the present analysis<br />

is to characterise the population receiving 8w LDV/SOF and<br />

to describe outcomes in clinical practice. Methods: The first<br />

CHC patients treated with 8w LDV/SOF in a single centre in<br />

Germany, and for whom SVR12 will be available for presentation,<br />

were included. Baseline characteristics, prior treatment<br />

history, safety and effectiveness were investigated. Analysis<br />

was performed using descriptive statistics. Results: 86 patients<br />

met the inclusion criteria and initiated 8w treatment with LDV/<br />

SOF between 21/11/2014 to 07/04/2015. No patient<br />

had ribavirin added to the STR. The mean (SD) age was 49.8<br />

(12.1) years and 44.2% were males. Genotype distribution<br />

was 52.3%, 45.4% and 2.3% for GT1a, GT1b and GT4. The<br />

METAVIR stage distribution at baseline was 51.2%, 26.7%,<br />

17.4% and 4.7% for F0, F1, F2 and F3. No patients had cirrhosis.<br />

Median HCV RNA at baseline was 794,328 (Q1-Q3:<br />

169,824-2,041,738, Min-Max: 11 -18,620,871) IU/ml, two<br />

patients had HCV RNA ≥ 6 million IU/mL (11,481,536 and<br />

18,620,871 IU/ml, both METAVIR stage F0). 4.7% of patients<br />

were HIV co-infected and no patients were HBV co-infected.<br />

Overall, 99% of the patients were treatment-naïve. One patient<br />

had relapsed after previous interferon (INF)/RBV therapy. To<br />

date, one patient discontinued therapy due to lack of adherence.<br />

Two patients experienced adverse events: one assessed<br />

as unrelated and one as possibly related to treatment. Among<br />

patients for whom data is available, SVR4 was 100% (69/69)<br />

and SVR12 was 100% (14/14). Complete results for SVR12,<br />

adverse events and discontinuations will be available at the<br />

time of presentation. Conclusion: 8w LDV/SOF is predominantly<br />

prescribed according to the SmPC for treatment-naïve<br />

non-cirrhotic CHC patients with HCV RNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 807A<br />

Disclosures:<br />

Paul Martin - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Merck, Gilead, Janssen, Abbvie<br />

The following authors have nothing to disclose: Eduardo A. Rodriguez, Cynthia<br />

Levy, Kalyan R. Bhamidimarri<br />

1207<br />

Assessment of serum hepatitis C virus RNA 24 weeks<br />

after the end of treatment using TaqMan PCR is more<br />

relevant than after 12 weeks for predicting sustained<br />

virological response by simeprevir- versus telaprevir-based<br />

triple therapy<br />

Tatsuo Kanda, Shingo Nakamoto, Reina Sasaki, Yuki Haga,<br />

Masato Nakamura, Shin Yasui, Makoto Arai, Fumio Imazeki,<br />

Osamu Yokosuka; Department of Gastroenterology and Nephrology,<br />

Chiba University, Graduate School of Medicine, Chiba,<br />

Japan<br />

Background and Aims: Hepatitis C virus (HCV) infection causes<br />

acute and chronic hepatitis and results in cirrhosis and hepatocellular<br />

carcinoma (HCC). Direct-targeting antiviral agents<br />

(DAAs) against HCV with or without peginterferon plus ribavirin<br />

result in higher rates of eradication of this virus and shorter<br />

treatment duration. We examined which is better for predicting<br />

persistent virological response, the assessment of serum HCV<br />

RNA 12 or 24 weeks after the end of treatment for predicting<br />

sustained virological response (SVR12 or SVR 24, respectively)<br />

in patients treated with peginterferon plus ribavirin and HCV<br />

NS3/4A protease inhibitor. Patients and Methods: In all, 145<br />

Japanese patients infected with HCV genotype 1 treated with<br />

peginterferon plus ribavirin with HCV NS3/4A protease inhibitor<br />

were retrospectively analyzed: 59 patients (57.6 years;<br />

male, 42; and IL28B rs8099917TT, 40) and 86 patients (60.8<br />

years; male, 42; and IL28B rs8099917TT, 58) were treated<br />

by telaprevir- and simeprevir-including regimens, respectively.<br />

Written informed consent was obtained from all patients, and<br />

this study conformed to the ethical guidelines of the 1975<br />

Declaration of Helsinki and was approved by the Ethics Committee<br />

of Chiba University, School of Medicine (No.523 and<br />

No.1462). HCV RNA was measured by TaqMan HCV Test,<br />

version 2.0, real-time PCR assay. IL28B rs8099917 was analyzed<br />

by TaqMan SNP assay. SVR12 and SVR24, respectively,<br />

were defined as HCV RNA negativity at 12 week and 24 week<br />

after stopping treatment. Results: Total SVR rates were 78.0%<br />

and 70.7%, in telaprevir and simeprevir groups, respectively.<br />

In telaprevir group, SVR rates of treatment-naïve, previous-treatment<br />

relapsers and others were 76.7%, 87.0% and 50.0%,<br />

respectively. In simeprevir group, SVR rates of treatment-naïve,<br />

previous-treatment relapsers and others were 75.0%, 100%<br />

and 25.0%, respectively. In telaprevir group, all 46 patients<br />

with SVR12 finally achieved SVR24. In simeprevir group, only<br />

4 (7.1%) of total 56 patients with SVR12 achieved SVR24, and<br />

these 4 patients were all previous-treatment relapsers (67.5<br />

years; male, 2; and IL28rs8099917TT, 2). Conclusion: SVR12<br />

were suitable for predicting persistent virological response in<br />

almost cases. In simeprevir-including regimens, SVR12 could<br />

not always predict persistent virological response. Clinician<br />

should use SVR24 for predicting persistent virological response<br />

in the use of DAA treatment for a group of real-world patients<br />

chronically infected with HCV.<br />

Disclosures:<br />

Tatsuo Kanda - Grant/Research Support: MSD K.K.; Speaking and Teaching:<br />

AJINOMOTO PHARMACEUTICALS CO., LTD, CHUGAI PHARMACEUTICAL<br />

CO., LTD, GlaxoSmithKlein, BMS, Jansen, Daiichisankyo; Stock Shareholder:<br />

Mitsubishi Tanabe Pharma<br />

Osamu Yokosuka - Grant/Research Support: Chugai, Taiho, Bristol Myers,<br />

Takeda<br />

The following authors have nothing to disclose: Shingo Nakamoto, Reina Sasaki,<br />

Yuki Haga, Masato Nakamura, Shin Yasui, Makoto Arai, Fumio Imazeki<br />

1208<br />

Reduced Survival in Elderly Liver Transplant Recipients:<br />

How Old is Too Old?<br />

Nae-Yun Heo 1,2 , Ajitha Mannalithara 1 , Prowpanga Udompap 1 ,<br />

Donghee Kim 1 , Waldo Concepcion 3 , Carlos O. Esquivel 3 , W.<br />

Ray Kim 1 ; 1 Division of Gastroenterology and Hepatology, Stanford<br />

University School of Medicine, Stanford, CA; 2 Department<br />

of Internal Medicine, Inje University College of Medicine, Busan,<br />

Korea (the Republic of); 3 Department of Surgery, Stanford University<br />

School of Medicine, Stanford, CA<br />

Background/Aims: Liver transplantation (LTx) is increasingly<br />

performed in patients over 65 years of age, an age group<br />

traditionally considered to be high risk. Prognostic models specifically<br />

designed to predict post-LTx outcome in elderly patients<br />

may help select optimal candidates. Methods: Records of all<br />

adult (>18 years) LTx recipients were extracted from the Organ<br />

Procurement and Transplantation Network (OPTN) database<br />

for 2004-2013. After exclusion of recipients of multi-organ or<br />

repeated LTx, recipients were divided into three groups:


808A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1209<br />

The Small Molecule TLR9 Antagonist COV08-0064<br />

Protects from Warm Ischemia Reperfusion Injury of the<br />

Liver<br />

Mohamed E. Shaker 1,2 , Bobby N. Trawick 3 , Wajahat Z. Mehal 1 ;<br />

1 Section of Digestive Diseases, Yale School of Medicine, New<br />

Haven, CT; 2 Department of Pharmacology & Toxicology, Faculty of<br />

Pharmacy, Mansoura University, Mansoura 35516, Egypt; 3 Center<br />

for Organic Chemistry, Mallinckrodt Pharmaceuticals, St. Louis,<br />

MO<br />

Warm ischemia/reperfusion (IR) injury is a major reason for<br />

failure of liver surgeries, and also limits the use of steatotic livers<br />

for transplantation. A significant portion of the tissue injury<br />

in IR is mediated by activation of an innate immune response<br />

upon reperfusion. Minimizing this innate immune response is an<br />

attractive goal to limit IR injury. Here, we present a novel and<br />

selective small molecule Toll-like receptor (TLR9) antagonist with<br />

the ability to protect from IR injury. Methods: Male C57Bl/6<br />

mice (8-10 weeks) were subjected to 1h of warm hepatic ischemia<br />

by occlusion of the vasculature supplying the left and<br />

median lobes of the liver. Control mice had the same surgical<br />

procedure, but without interruption of liver blood flow. COV08-<br />

0064 was administered at a dose of 80 mg/kg intraperitoneally<br />

1h prior starting ischemia. At 3 h and 12 h after resuming<br />

blood flow, the livers were examined by standard histology,<br />

scored for injury, and qPCR for pro- and anti-inflammatory cytokines<br />

performed. The ability of COV08-0064 to regulate inflammatory<br />

cytokine production was also tested in vitro on mouse<br />

macrophages and plasmacytoid dendritic cells (pDCs). Results:<br />

COV08-0064 pretreatment for I/R-mice resulted in limiting the<br />

rise in serum alanine aminotransferase activity (595 ± 124 vs<br />

3233 ± 785, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 809A<br />

Cox proportional hazards analysis to assess post-transplant<br />

survival. Results Between 2002 and 2014, the mean age of<br />

liver transplant registrants increased from 51.2 to 55.7, with<br />

a more prominent increase in HCV-positive (50.9 to 57.9)<br />

than HCV-negative (51.3 to 54.3) registrants. The proportion<br />

of registrants aged ≥ 60 years increased from 19% to 41%<br />

and the proportion aged ≥ 65 years increased from 8.1% to<br />

17%. Almost identical trends were observed among liver transplant<br />

recipients. Among transplant registrants, increasing age<br />

was associated with increased mortality before transplantation<br />

(adjusted sub-hazard ratio 1.30 for age 50-59, 1.49 for age<br />

60-64, 1.73 for age 65-69 and 2.04 for age ≥70 relative<br />

to age 18-49) and decreased likelihood of transplantation.<br />

Among transplant recipients, increasing age was associated<br />

with increased post-transplant mortality (adjusted hazard ratio<br />

1.16 for age 50-59, 1.34 for age 60-64, 1.61 for age 65-69<br />

and 1.87 for age ≥70 relative to age 18-49). Conclusions<br />

Dramatic aging of liver transplant recipients and registrants<br />

occurred from 2002 to 2014, most prominently in HCV-infected<br />

patients. The combination of higher waitlist mortality and<br />

higher post-transplant mortality in elderly patients suggests that<br />

listing of these patients should be pursued cautiously.<br />

Proportion of liver transplant registrants and recipients >=60<br />

years of age by HCV status, 2002-2014<br />

Grade 1 [47(38.2%)], ACLF Grade 2 [47(38.2%)], and ACLF<br />

Grade 3 [29(23.6%)]. Results: ACLF patients were younger<br />

(52yrs. Vs 54 yrs., p=0.018) and have significantly higher<br />

MELD score at LT (30.47±6.98 vs. 19.74±4.44, p=0.001).<br />

Higher proportion of underlying liver disease included alcohol,<br />

autoimmune hepatitis and Wilson’s disease in ACLF patients.<br />

ACLF recipients needed longer hospitalization (20.13±23.87<br />

vs. 11.51±11.41 days, p=0.001), ICU stays (9.43±17.91<br />

vs. 4.23±8.99 days, p=0.001), higher early (≤90 days) readmissions<br />

(0.51±0.73 vs. 0.35±0.68, p=0.008), and higher<br />

early (≤90 days) surgical re-explorations (26 vs 14.8%,<br />

P=0.002). Graft loss was significant in the ACLF group (35.6<br />

% vs. 27.5%, p=0.044, median follow up =1476 days), and<br />

was pronounced at 1 year follow up (18.7 % vs. 10.3%,<br />

p=0.009). Graft (p=0.035, Figure 1A) and patient survival<br />

(p=0.025, Fig.1B) was significantly lower in the ACLF group,<br />

with incremental poor graft (p=0.013, Fig.1A) and patient<br />

survival (p=0.022, Figure 1B) in higher grades of ACLF. Graft<br />

(p=0.086, Fig. 1C) and patient survival (p=0.118, Fig. 1D)<br />

based on MELD severity categories were not significant by<br />

Kaplan-Meier analysis. Conclusion: ACLF recipients defined by<br />

EASL-CLIF score have inferior overall graft and patient survival,<br />

and their length of post-LT hospitalization, length of ICU stay,<br />

early readmissions, and early surgical re-explorations were<br />

significantly higher.<br />

Figure 1. Cumulative graft (1A) and patient survival (1B) in ACLF<br />

grades. Cumulative graft (1C) and patient survival (1D) in various<br />

MELD categories.<br />

Disclosures:<br />

Charles S. Landis - Grant/Research Support: Gilead, Abbvie, BMS<br />

The following authors have nothing to disclose: Feng Su, Lei Yu, Kristin Berry, Iris<br />

W. Liou, Jorge Reyes, Stephen C. Rayhill, George N. Ioannou<br />

1212<br />

Outcomes of acute-on-chronic liver failure in liver transplant<br />

recipients based on EASL-CLIF consortium definition<br />

Anuj Sharma 1 , William J. Goldkamp 1 , Satheesh Nair 2,1 , Jason<br />

Vanatta 2,1 , James Eason 2,1 , Sanjaya K. Satapathy 2,1 ; 1 University<br />

of Tennessee Health Science Center, Memphis, TN; 2 Methodist<br />

Transplant Institute, Memphis, TN<br />

Background: Outcomes of liver transplant (LT) recipients with<br />

acute-on-chronic liver failure (ACLF) with recently defined<br />

EASL-CLIF consortium definition has not been studied. We<br />

aimed to study the long term outcomes of LT patients with<br />

ACLF. Methods: After retrospective review of 833 consecutive<br />

LT recipients (04/2006 to 03/2013), and exclusion of<br />

patients without chronic liver disease, re-LTs, SLK recipients,<br />

and acute liver failure, 123 ACLF and 582 without ACLF were<br />

included. LT recipients with ACLF were divided into ACLF<br />

Disclosures:<br />

Satheesh Nair - Advisory Committees or Review Panels: Jansen; Grant/Research<br />

Support: Gilead; Speaking and Teaching: Abbvie, Valeant, BMS<br />

Sanjaya K. Satapathy - Advisory Committees or Review Panels: Gilead, Intercept;<br />

Grant/Research Support: Genfit, Gilead, Biotest<br />

The following authors have nothing to disclose: Anuj Sharma, William J. Goldkamp,<br />

Jason Vanatta, James Eason


810A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1213<br />

Prediction of Posthepatectomy Liver Failure based on<br />

Preoperative Liver Fibrosis Assessment: Comparison of<br />

Mac-2 Binding Protein Glycosylation Isomer and Liver<br />

Stiffness Measurement<br />

Takahiro Nishio, Kojiro Taura, Kazutaka Tanabe, Gen Yamamoto,<br />

Yukihiro Okuda, Yoshinobu Ikeno, Satoru Seo, Kentaro Yasuchika,<br />

Etsuro Hatano, Hideaki Okajima, Toshimi Kaido, Shinji Uemoto;<br />

Department of Surgery, Graduate School of Medicine, Kyoto University,<br />

Kyoto, Japan<br />

Posthepatectomy liver failure (PHLF) is one of the major<br />

causes of hepatectomy-related mortality, and accurate prediction<br />

of PHLF based on preoperative liver function test is<br />

essential. Recently, Mac-2 binding protein glycosylation isomer<br />

(M2BPGi) has been proposed as a useful liver fibrosis<br />

marker, while liver stiffness (LS) measurement was also well<br />

accepted as a noninvasive assessment tool for liver fibrosis.<br />

We evaluated the usefulness of these novel liver fibrosis markers,<br />

M2BPGi and LS, for predicting PHLF. One hundred and<br />

thirty-eight patients with hepatocellular carcinoma undergoing<br />

hepatectomy between August 2011 and October 2014 were<br />

analyzed. Preoperative serum M2BPGi level was measured<br />

in addition to a routine blood test, and LS was measured by<br />

acoustic radiation force impulse-based Virtual Touch Tissue<br />

Quantification (VTTQ) technology. Remnant liver volume rate<br />

(REM) was calculated by computed tomography volumetry.<br />

PHLF was diagnosed according to the definition of the International<br />

Study Group of Liver Surgery and was graded as A,<br />

B, or C. The accuracy of predictors was assessed by the area<br />

under the receiver operating characteristic curve (AUROC). The<br />

mean serum M2BPGi level (cut off index) was 1.0±0.5 in F0-1,<br />

1.7±1.1 in F2-3, and 4.5±3.6 in F4; and the mean VTTQ value<br />

(m/s) was 1.4±0.3, 1.7±0.6, and 2.7±0.9, respectively; both<br />

of which significantly increased according to the progression<br />

of liver fibrosis stage (p < .01). M2BPGi had AUROCs of 0.80<br />

for the diagnosis of F≥2, 0.82 for F≥3, and 0.83 for F4, which<br />

were better than any other fibrosis markers including the VTTQ<br />

value. PHLF occurred in 34 patients (24.6%): grade A, 15<br />

patients (10.9%); grade B, 14 patients (10.1%); grade C, 5<br />

patients (3.6%). The AUROCs of M2BPGi were 0.61 for the<br />

prediction of PHLF grade ≥A, 0.70 for grade ≥B, and 0.64<br />

for grade C; while those of VTTQ value were 0.66, 0.77, and<br />

0.72, respectively. Multivariate stepwise selection identified 3<br />

best fit significant factors for the prediction of PHLF: M2BPGi<br />

(odds ratio [OR], 1.6; 95% confidence interval [CI], 1.1-2.4,<br />

p = .013), platelet count (OR, 0.42; 95% CI, 0.23-0.72, p <<br />

.01), and REM (OR, 0.37; 95% CI, 0.23-0.59, p < .01). The<br />

predictive model including these 3 variables had AUROCs<br />

of 0.77 for predicting PHLF grade ≥A, 0.81 for grade ≥B,<br />

and 0.80 for grade C; and this model enabled the estimation<br />

of safe resection liver volume according to the risk of PHLF<br />

by the preoperative M2BPGi and platelet count. Conclusions.<br />

M2BPGi and LS were useful markers for predicting PHLF as<br />

well as liver fibrosis progression.<br />

Disclosures:<br />

Etsuro Hatano - Speaking and Teaching: Bayer<br />

The following authors have nothing to disclose: Takahiro Nishio, Kojiro Taura,<br />

Kazutaka Tanabe, Gen Yamamoto, Yukihiro Okuda, Yoshinobu Ikeno, Satoru<br />

Seo, Kentaro Yasuchika, Hideaki Okajima, Toshimi Kaido, Shinji Uemoto<br />

1214<br />

Two-Year Results of a Pilot Program in Early Liver Transplantation<br />

for Severe Alcoholic Hepatitis<br />

Brian P. Lee 1 , Po-Hung Chen 1 , Ruben Hernaez 1 , Ahmet Gurakar 1 ,<br />

Benjamin Philosophe 2 , Zhiping Li 1 ; 1 Medicine, Johns Hopkins University<br />

School of Medicine, Baltimore, MD; 2 Surgery, Johns Hopkins<br />

University School of Medicine, Baltimore, MD<br />

OBJECTIVE: Six months of alcohol abstinence is typically<br />

required before a patient is considered for liver transplant (LT).<br />

In October 2012, our LT center initiated a pilot program to<br />

transplant selected patients with acute alcoholic hepatitis. The<br />

purpose of this study was to assess the results of this pilot program<br />

and the effects on organ allocation. METHODS: Data<br />

was collected from all patients with alcohol-related liver disease<br />

(ALD) since initiation of our pilot program in October<br />

2012. These patients were stratified into three groups based<br />

on indication for LT: severe alcoholic hepatitis as first liver<br />

decompensation (Group 1), alcoholic cirrhosis with ≤6 months<br />

abstinence (Group 2), alcoholic cirrhosis with ≥6 months<br />

abstinence (Group 3). Follow-up period was defined as time<br />

between date of transplant and date of last follow-up or death.<br />

Recidivism was any evidence of alcohol consumption following<br />

LT. The number of organs allocated for ALD within 2 years of<br />

the pilot program was compared to the 2-year period immediately<br />

prior to implementation. RESULTS: 36 patients underwent<br />

LT. All patients in Group 1 had Maddrey’s Discriminant<br />

Function greater than 32. Table 1 has patient characteristics<br />

and outcomes. 18 of 36 (50%) LTs performed for ALD had<br />

less than 6 months of abstinence prior to listing. Compared<br />

to the 2-year period prior to the pilot program, there was an<br />

increase in the proportion of organs indicated for ALD (20%<br />

vs. 12%; p=0.13). 6-month survival was 100%, 100%, 83%<br />

in Group 1, 2, 3, respectively (p=0.16). CONCLUSIONS: For<br />

carefully selected patients presenting with alcoholic hepatitis,<br />

early LT resulted in similar 6-month survival and alcohol relapse<br />

rates compared to patients undergoing LT after 6 months of<br />

abstinence. Patients transplanted for acute alcoholic hepatitis<br />

(Group 1) had significantly higher MELD score and shorter<br />

length of time on the transplant list compared to both other<br />

groups. The implementation of a policy allowing early LT for<br />

ALD resulted in an increase in proportion of organs allocated<br />

for alcohol-related indications, but did not reach statistical significance.<br />

Table 1: Patient Characteristics and Outcomes<br />

*Excluding two patients who died during the post-operative<br />

period<br />

Disclosures:<br />

Ahmet Gurakar - Advisory Committees or Review Panels: Gilead; Grant/<br />

Research Support: BMS<br />

The following authors have nothing to disclose: Brian P. Lee, Po-Hung Chen,<br />

Ruben Hernaez, Benjamin Philosophe, Zhiping Li


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 811A<br />

1215<br />

Early Liver Transplantation in Severe Alcoholic Hepatitis:<br />

A Multi-Center, Retrospective, Cohort Study<br />

Brian P. Lee 1 , David W. Victor 2 , Darryn R. Potosky 3 , Ibrahim<br />

A. Hanouneh 4 , Mary E. Rinella 5 , Michael D. Voigt 6 , Sheila L.<br />

Eswaran 7 , Howard P. Monsour 2 , R. Mark Ghobrial 2 , Jessica A.<br />

Keppel 7 , Zhiping Li 1 ; 1 Johns Hopkins University School of Medicine,<br />

Baltimore, MD; 2 Houston Methodist Hospital, Houston,<br />

TX; 3 University of Maryland School of Medicine, Baltimore, MD;<br />

4 Cleveland Clinic, Cleveland, OH; 5 Northwestern University Feinberg<br />

School of Medicine, Chicago, IL; 6 University of Iowa Carver<br />

College of Medicine, Iowa City, IA; 7 Rush Medical College, Chicago,<br />

IL<br />

OBJECTIVE: Liver transplant (LT) centers typically require 6<br />

months of alcohol abstinence before a patient is considered<br />

for LT. Severe alcoholic hepatitis (SAH) refractory to medical<br />

treatment has a 6-month survival estimated at 30%. A European<br />

study showed that early LT of SAH in a highly selected<br />

group of patients could improve survival. The purpose of this<br />

study was to examine the outcomes of early LT for SAH in U.S.<br />

transplant centers. METHODS: All 111 U.S. LT centers were surveyed<br />

for patients who underwent LT for acute SAH (Maddrey’s<br />

Discriminant Function >32) as the first presentation of liver<br />

decompensation, and listing for LT prior to 6 months of alcohol<br />

abstinence. 48 centers responded to the survey. 10 centers<br />

reported such experience, of which 7 submitted retrospective<br />

data. All patients presented from October 2006 to November<br />

2014. Alcohol relapse was defined as any alcohol consumption<br />

following LT. Binge drinking was defined as 6 units of<br />

alcohol in a day for men, 4 units for women. Frequent drinking<br />

was defined as any alcohol consumption more than 4 days a<br />

week. RESULTS: 41 patients underwent early LT for acute SAH.<br />

Median follow-up was 2.4 years. Median period of alcohol<br />

abstinence prior to LT listing was 53 days. Median MELD at<br />

listing was 34. Median Maddrey’s discriminant function was<br />

69. Only 12 (29%) received steroids. Of those excluded from<br />

steroid therapy (n=29), 11 (38%) had confirmed infection,<br />

6 (21%) had presumed infection without confirmed source,<br />

5 (17%) had GI bleed, 7 (24%) were excluded due to provider<br />

choice. Six-month and 1-year survival were both 100%.<br />

Overall survival was 89%. 9 (22%) had alcohol relapse; of<br />

these patients (n=9), 6 (67%) reported binge drinking, 7 (78%)<br />

reported frequent drinking, 4 (44%) reported both binge and<br />

frequent drinking, and 4 (44%) had re-achieved sobriety by<br />

time of last follow-up. Compared to the cohort of Mathurin et<br />

al, steroid use was lower (29% vs. 92%, p


812A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Table 1<br />

Disclosures:<br />

James W. Ferguson - Advisory Committees or Review Panels: Astellas, Novartis<br />

The following authors have nothing to disclose: Suman Verma, Fiona M. Hand,<br />

Matthew J. Armstrong, Marie de Vos, Douglas Thorburn, Terry Pan, John R.<br />

Klinck, Rachel Westbrook, Georg Auzinger, Andrew Bathgate, Steven Masson,<br />

Andrew Holt, Diarmaid D. Houlihan<br />

1217<br />

Sublingual administration of tacrolimus is a feasible<br />

route and requires a lower dose to achieve similar drug<br />

exposition in adult liver transplant recipients.<br />

Sandra Solari 2 , Alejandra Cancino 1 , Blanca M. Norero 1 , Rodrigo<br />

Wolff 1 , Francisco Barrera 1 , Alejandro Soza 1 , Marco Arrese 1 , Juan<br />

Francisco Guerra 3 , Nicolás Jarufe 3 , Jorge A. Martínez 3 , Carlos E.<br />

Benítez 1 ; 1 Gastroenterology, Pontificia Universidad Católica de<br />

Chile, Santiago, Chile; 2 Laboratorio Clínico, Pontificia Universidad<br />

Católica de Chile, Santiago, Chile; 3 Digestive Surgery, Pontificia<br />

Universidad Católica de Chile, Santiago, Chile<br />

Introduction: Oral administration of immunosuppressive drugs<br />

is not always feasible after liver transplantation and parenteral<br />

administration of tacrolimus (TAC) implies a significantly higher<br />

expense of resources. The aim of this study is to evaluate the<br />

feasibility of sublingual (sl) administration of tacrolimus in adult<br />

liver transplant recipients. Methods: In patients on per oral<br />

administration (po) serum TAC levels were determined at different<br />

time points (0, 0.5, 1, 2, 4, 6 and 12 hours), then patients<br />

were switched to sl administration and TAC dose was tapered<br />

(starting with a 20-30% reduction of the po dose) to obtain a<br />

similar trough level that in po administration. Then TAC levels<br />

were determined at the same time points. The exposition to<br />

TAC (AUC) was estimated and compared for SL and po administration.<br />

Results: Seventeen recipients were enrolled. The mean<br />

time required to taper de sl dose to obtain a proper trough level<br />

was 16.41±10.5 days (range 6-40 days). There were no differences<br />

on trough levels on po and sl administrations (7.06±2.2<br />

vs 7.19±2.0 ng/ml respectively, p=ns). Remarkably, when<br />

AUC was evaluated, no differences were found in both groups<br />

(111.4±32.23 ng/ml/h vs 111.4±35.65 ng/ml/h, p=ns).<br />

However, the dose required to achieve a similar trough level<br />

was significantly higher when TAC was administered on po vs<br />

sl route (4.68±2.32 vs 2.94±1.7 mg/d, p=0.018), meaning a<br />

38% reduction. The maximum concentration after TAC administration<br />

(Cmax) was similar in sl and po groups (19.65±9.97<br />

vs 15.88±6.91 ng/ml, p=ns). No severe adverse effects were<br />

registered on sl group and was well tolerated. In one patient<br />

very few and small blisters were transiently observed for a few<br />

days and in 4 patients a spicy taste was transiently observed<br />

with the meals on sl group. Conclusion: Sublingual administration<br />

of tacrolimus is a feasible alternative to po administration<br />

and can also be monitored employing trough levels requiring<br />

a lower dose to achieve the same drug exposition. These findings<br />

also suggest that parenteral TAC administration could be<br />

replaced by sl route. These findings could imply a significantly<br />

reduction on the cost of immunosuppression therapy.<br />

Disclosures:<br />

Alejandro Soza - Advisory Committees or Review Panels: Merck, Roche, Gilead;<br />

Speaking and Teaching: Merck, BMS, Roche; Stock Shareholder: Gilead,<br />

Achillion<br />

The following authors have nothing to disclose: Sandra Solari, Alejandra Cancino,<br />

Blanca M. Norero, Rodrigo Wolff, Francisco Barrera, Marco Arrese, Juan<br />

Francisco Guerra, Nicolás Jarufe, Jorge A. Martínez, Carlos E. Benítez<br />

1218<br />

Pre-transplant echocardiographic parameters as a tool<br />

to evaluate post-transplant survival and cardiac outcomes<br />

in liver transplant recipients<br />

Daniel W. Bushyhead 1 , James N. Kirkpatrick 2 , David S. Goldberg<br />

3 ; 1 Internal Medicine, Hospital of the University of Pennsylvania,<br />

Philadelphia, PA; 2 Cardiology, Hospital of the University of<br />

Pennsylvania, Philadelphia, PA; 3 Gastroenterology, Hospital of the<br />

University of Pennsylvania, Philadelphia, PA<br />

Background Despite advances in liver transplantation and<br />

pre-operative risk stratification, there remains significant<br />

post-transplant morbidity and mortality from cardiovascular<br />

disease. There are limited and conflicting data on the role of<br />

pre-transplant echocardiography in risk stratifying patients with<br />

cirrhosis. Moreover, the relationship between pre-transplant<br />

cardiac function and post-transplant cardiovascular and renal<br />

disease is unknown. The purpose of our study was to determine<br />

if pre-transplant echocardiography predicts post-transplant outcomes<br />

in liver transplant recipients. Methods We conducted a<br />

retrospective analysis of adult liver transplant recipients at the<br />

University of Pennsylvania from January 1, 2005 to September<br />

30, 2014. We included only patients with pre-transplant echocardiograms<br />

that were read by a cardiologist at our institution.<br />

Patients with acute liver failure and those without a diagnosis<br />

of cirrhosis were excluded. Pre-transplant echocardiographic<br />

parameters were: ejection fraction, diastolic dysfunction, aortic<br />

stenosis, regional wall motion abnormalities, mitral and tricuspid<br />

regurgitation, left atrial size, right ventricular size and<br />

function, pulmonary artery systolic pressure (PASP) and right<br />

atrial pressure. Post-transplant outcomes were: patient survival,<br />

major adverse cardiac events (MACE; myocardial infarction<br />

and heart failure) and chronic kidney disease (CKD) stage 4 or<br />

5 (estimated glomerular filtration rate


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 813A<br />

1219<br />

Reporting Physical Function in UNOS: The Weakness of<br />

the Karnofsky Performance Status Scale<br />

Connie W. Wang, John P. Roberts, Jennifer C. Lai; University of<br />

California, San Francisco, San Francisco, CA<br />

Background: Currently, physical function (fxn) is reported in<br />

UNOS using the Karnofsky Performance Status (KPS) scale,<br />

which ranges from 0-100% as determined by the liver transplant<br />

(LT) team. Given its subjective nature, we aimed to evaluate<br />

factors influencing the KPS score in the UNOS registry.<br />

Methods: Included were all adult LT recipients from 9/08-<br />

9/13. Physical fxn was categorized as low (0-20%), medium<br />

(30-70%), or high (80-100%) on the KPS. Cox models assessed<br />

the relationship between KPS and post-LT mortality, ordinal<br />

logistic models the relationship between KPS with patient (pt)/<br />

center (ctr)-level factors. Results: Of 25,632 LT pts, 21%, 56%<br />

and 23% had low, medium, and high physical fxn at LT. Low,<br />

medium, and high physical fxn-ing pts differed by: %female<br />

(41 vs. 33 vs. 26%), %hospitalized in ICU (54 vs. 4 vs. 1%),<br />

median lab MELD (35 vs. 20 vs. 15), and median Child score<br />

(12 vs. 10 vs. 8) [p


814A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1221<br />

Renal Recovery Outcomes after Liver Transplantation<br />

in Patients Requiring Pre-operative Renal Replacement<br />

Therapy<br />

Rex Cheng 1 , Kimberly Muczynski 2 , Lena Sibulesky 3 , Renuka Bhattacharya<br />

1 ; 1 Gastroenterology/Hepatology, University of Washington,<br />

Seattle, WA; 2 Nephrology, University of Washington, Seattle,<br />

WA; 3 Transplant Surgery, University of Washington, Seattle, WA<br />

Background: The purpose of this study is to characterize the<br />

long-term outcomes of renal recovery in liver transplant (OLT)<br />

recipients who required preoperative renal replacement therapy<br />

(RRT). Methods: A single center retrospective review was<br />

performed for OLT recipients who were transplanted from January<br />

2009 to April 2015 and required RRT at least one day<br />

prior to transplant. Patients were excluded if they received a<br />

combined liver-kidney transplant or required a repeat transplant.<br />

Patient demographics, clinical data, pre- and post-transplant<br />

laboratory data, number of days on RRT, time to RRT<br />

cessation were extracted from an electronic medical record.<br />

Outcomes measured include early renal recovery (cessation<br />

of RRT within 30 days after OLT), delayed renal recovery (cessation<br />

of RRT more than 30 days after OLT) or development<br />

of end-stage renal disease, as well as residual renal function,<br />

characterized by serum creatinine at 30 and 90 days after<br />

cessation of RRT. Results: Excluding combined liver-kidney<br />

transplants (n=17) and patients requiring repeat transplant (n<br />

= 14), 454 patients underwent liver transplantation during this<br />

time period. Sixty patients (mean age 51.9 ± 9.7 years, 55.0%<br />

male, 62% with acute tubular necrosis, mean Model of End-<br />

Stage Liver Disease score 36.7 ± 4.7 at transplant) received<br />

RRT prior to OLT. 34 (57%) achieved early renal recovery, 15<br />

(25%) achieved delayed renal recovery over a mean of 57.6<br />

days and 11 (18%) were deemed to have end-stage renal<br />

disease with RRT dependence. Patients who achieved early<br />

renal recovery had a shorter duration of pre-transplant RRT<br />

than those who did not, 17 days, [95% confidence interval<br />

(CI), 12-21 days] versus 26 days [(95% CI, 19-33 days), p =<br />

0.044]. At 30 days after cessation of RRT (n= 48), 70.5% of<br />

patients with early renal recovery vs 40.0% of patients with<br />

delayed renal recovery attained a serum creatinine < 2 (p =<br />

0.046). At 90 days after cessation of RRT (n=44), 93.3% of<br />

patients with early renal recovery vs 78.5% of patients with<br />

delayed renal recovery attained a serum creatinine < 2 (p =<br />

0.148). Conclusion: Over three-quarters of patients requiring<br />

pre-transplant RRT achieve renal recovery. Shorter duration<br />

of RRT is associated with earlier renal recovery. Early renal<br />

recovery is associated with more rapid improvement in residual<br />

renal function. The majority of patients who achieve renal<br />

recovery have improved residual renal function at 90 days<br />

after RRT cessation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Rex Cheng, Kimberly Muczynski,<br />

Lena Sibulesky, Renuka Bhattacharya<br />

1222<br />

Coffee consumption promotes survival and protects<br />

against hepatocellular carcinoma recurrence following<br />

liver transplantation: putative mechanisms involve modulation<br />

of adenosinergic signalling<br />

Georg Wiltberger 1 , Ruoyu Miao 2 , Undine G. Lange 1 , Rudi<br />

Ascherl 1 , Hans-Michael Hau 1 , Felix Krenzien 1 , Georgi Atanasov 1 ,<br />

Christian Benzing 1 , Elliot B. Tapper 2 , Linda Feldbrügge 2 , Johannes<br />

Broschewitz 1 , Michael Bartels 1 , Sven Jonas 3 , Johann Pratschke 4 ,<br />

Yan Wu 2 , Simon C. Robson 2 , Moritz Schmelzle 4 ; 1 Department of<br />

Visceral, Transplant, Thoracic and Vascular Surgery, University<br />

Hospital Leipzig, Leipzig, Germany, Leipzig, Germany; 2 Department<br />

of Medicine, Beth Israel Deaconess Medical Center, Harvard<br />

Medical School, Liver Center and The Transplant Institute, Div. of<br />

Gastroenterology, Boston, MA; 3 Department of Hepato-, Pancreato-<br />

and Biliary Surgery, 310Klinik, Nürnberg, Nürnberg, Germany;<br />

4 Department of General, Visceral, and Transplant Surgery,<br />

Charité - Universitätsmedizin Berlin, Campus Virchow Klinikum,<br />

Berlin, Germany<br />

Background: Increasing evidence suggests that coffee consumption<br />

might decrease liver cirrhosis death risk and reduce<br />

the risk for hepatocellular carcinoma (HCC). Caffeine is a natural<br />

antagonist to adenosine-mediated receptor signaling and<br />

has tumoricidal activity in experimental settings. Aim: To determine<br />

whether coffee consumption decreases HCC recurrence<br />

and promotes survival following orthotopic liver transplantation<br />

(OLT) and to examine putative mechanisms of action.<br />

Methods: Patients who underwent OLT for HCC from January<br />

2002 to December 2012 were interviewed on the numbers<br />

of consumed cups of coffee per day before and after OLT.<br />

Coffee consumption was correlated with patients’ data with<br />

regard to HCC recurrence and patient survival. For in vitro<br />

mechanistic assays, human HepG2 HCC cells were analyzed<br />

for expression of adenosine receptors and ectoenzymes. Cells<br />

were also treated with adenosine, in the presence or absence<br />

of 8-(3-Chlorostyryl)-caffeine (CSC) or alloxazine, followed by<br />

analyses of proliferation, metastasis and alterations in key adenosine-mediated<br />

cancer pathways inclusive of MAPK (ERK and<br />

JNK) and NF-kappa B. Results: 90 patients with a median<br />

age of 60 years (range 73 – 37) and a median BMI of 26.95<br />

(range 18.3 – 39.4) underwent OLT for histologically confirmed<br />

HCC. 16 (17.8 %) patients experienced HCC-recurrence after<br />

median time of 11.5 months (range 1 - 40.5) post-transplant.<br />

In patients with a high preoperative or postoperative coffee<br />

intake HCC recurrence was observed less frequently than<br />

in those with low coffee intake (p=0.018). In a multivariate<br />

Cox’ proportional hazard model for recurrence-free survival<br />

only postoperative coffee intake ≥3 cups daily showed a significant<br />

relevance. HepG2 cells express abundant levels of<br />

A2A and A2B as well as several key ectoenzymes (including<br />

ENTPD3, ENTPD5, ENTPD8, CD73, and adenosine deaminase<br />

1 (ADA1)). These cells show intrinsic capacity to generate<br />

extracellular adenosine. Exogenous adenosine promotes HCC<br />

cell growth and metastasis in vitro, which is blocked by CSC,<br />

but not by alloxazine. Moreover, adenosine induces activation<br />

of MAPK (ERK and JNK) and NF-kappaB pathways, which can<br />

be prevented by antagonism of A2A and A2B receptors. Conclusions:<br />

Coffee consumption is associated with a decreased<br />

risk of HCC recurrence and increased survival following OLT.<br />

Experimental data suggest that these results might be, at least<br />

in part, associated with the antagonist activity of caffeine on<br />

adenosine-A2AR mediated growth-promoting effects in HCC<br />

cells. If validated further, our findings have implications for<br />

future recommendations for HCC patients after OLT.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 815A<br />

Disclosures:<br />

Simon C. Robson - Grant/Research Support: Pfizer, NIH, Dainippon; Independent<br />

Contractor: Biolegend, EMD Millipore, Mersana; Management Position:<br />

eBioscience; Speaking and Teaching: ACP, Elsevier, ATC; Stock Shareholder:<br />

Nanopharma, Puretech<br />

Moritz Schmelzle - Grant/Research Support: Novartis GmbH<br />

The following authors have nothing to disclose: Georg Wiltberger, Ruoyu Miao,<br />

Undine G. Lange, Rudi Ascherl, Hans-Michael Hau, Felix Krenzien, Georgi Atanasov,<br />

Christian Benzing, Elliot B. Tapper, Linda Feldbrügge, Johannes Broschewitz,<br />

Michael Bartels, Sven Jonas, Johann Pratschke, Yan Wu<br />

1223<br />

Recipient obesity is an independent risk factor for<br />

unplanned reoperations and prolonged operative time<br />

in orthotopic liver transplant<br />

Eric S. Wise 1 , Malena Outhay 2 , Michael Maggart 2 , Saad Rehman<br />

2 , Sir Norman T. Melancon 2 , James Leathers 2 , Kristina Jacomino<br />

2 , Michelle Izmaylov 2 , Amol Utrankar 2 , Kyle M. Hocking 1 , Sunil<br />

K. Geevarghese 1,3 ; 1 Surgery, Vanderbilt, Baltimore, MD; 2 School<br />

of Medicine, Vanderbilt, Nashville, TN; 3 Transplant Center, Vanderbilt,<br />

Nashville, TN<br />

Background: Selection of orthotopic liver transplant recipients<br />

is increasingly inclusive of patients with higher body-mass indices<br />

(BMI), as BMI has not been implicated as a determinant<br />

of patient or graft survival. As larger patients present a surgical<br />

challenge, we aim to characterize the influence of obesity<br />

on important, yet ill-characterized surgery-specific outcomes.<br />

Methods: 382 included OLT patients with BMI values reflecting<br />

obesity (>30 kg/m 2 ) or normal weight (20-25 kg/m 2 ) from<br />

2005-2014, from a single instituiton, were reviewed for rationally<br />

chosen recipient-based preoperative factors potentially<br />

influencing the need for an unplanned reoperation (UR), or<br />

the length of operation. Factors included BMI, demographics,<br />

serum laboratory <strong>studies</strong>, comorbidities, etiology and subjective<br />

measures of disease severity (e.g. ascites, inpatient status).<br />

Factors that trended (P30 kg/m 2 (P=.07), younger age<br />

(P=.04), higher MELD (P=.05) and recent renal replacement<br />

(P=.08). Younger age (OR 0.97/year, P=.01) and BMI>30<br />

kg/m 2 (OR 1.72, P=.03) remained independently significant.<br />

Conclusions: Obesity may make OLT more technically challenging,<br />

and thus represents an independent risk factor for UR<br />

and prolonged operative time.<br />

Disclosures:<br />

The following authors have nothing to disclose: Eric S. Wise, Malena Outhay,<br />

Michael Maggart, Saad Rehman, Sir Norman T. Melancon, James Leathers,<br />

Kristina Jacomino, Michelle Izmaylov, Amol Utrankar, Kyle M. Hocking, Sunil<br />

K. Geevarghese<br />

1224<br />

Ursodeoxycholic Acid as Adjunctive Therapy to Endoscopic<br />

Management of Non Anastomotic Biliary Lesions<br />

after Liver Transplant: Effectiveness Depends on Etiology<br />

Mark Pedersen 1 , David W. Backstedt 2 , Bobby Kakati 2 , Myunghan<br />

Choi 3 , Anil B. Seetharam 4 ; 1 Internal Medicine, University of<br />

Arizona College of Medicine-Phoenix, Phoenix, AZ; 2 Gastroenterology,<br />

University of Arizona College of Medicine-Phoenix, Phoenix,<br />

AZ; 3 Arizona State University College of Nursing and Health<br />

Innovation, Phoenix, AZ; 4 Banner Transplant and Advanced Liver<br />

Disease Center, University of Arizona College of Medicine, Phoenix,<br />

AZ<br />

Rationale: Non anastomotic biliary lesions (NABLs) are a cause<br />

of morbidity after liver transplant (LT) and are postulated to<br />

occur as a result of diverse mechanisms including insufficiency<br />

of allograft arterial supply and occult cytomegalovirus (CMV)<br />

infection. Ursodeoxycholic acid (UDCA) abrogates bile duct<br />

injury in chronic cholestatic conditions and study is needed<br />

to evaluate utility as adjunctive therapy for NABLs after LT.<br />

Aim: Evaluate effect of post-operative UDCA use on markers of<br />

cholestasis and management of NABLs after LT. Methods: Retrospective,<br />

unmatched cohort study identifying consecutive LT<br />

recipients over a 5 year period (2009-2014) with 1 year follow<br />

up. Recipients with Roux-en-Y biliary reconstruction excluded.<br />

Biliary complications were defined as endoscopically assessed<br />

non anastomotic stricture or leak. “UDCA group” was defined<br />

as subjects with uninterrupted usage of UDCA (300mg PO<br />

TID) for a minimum of one month post-LT. “Non-UDCA group”<br />

had no recorded exposure after LT. Baseline demographics<br />

included: age, gender, indication for LT, Model for End Stage<br />

Liver Disease (MELD) score at LT, blood type, body mass index<br />

(BMI), cytomegalovirus (CMV) donor status, need for take-back<br />

surgery for non biliary complication 1 month from LT, warm<br />

ischemic time, cold ischemic time, and donation after cardiac<br />

death (DCD). Outcomes compared between UDCA and non-<br />

UDCA groups included: number of biliary complications, number<br />

of ERCPs, number of stents placed, and serum bilirubin at:<br />

hospital discharge post-LT, first ERCP, 3, 6, 9, and 12 months.<br />

Categorical variables were analyzed using Fisher’s exact tests.<br />

Continuous variables were compared with student’s t test with<br />

p


816A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Mark Pedersen, David W. Backstedt,<br />

Bobby Kakati, Myunghan Choi<br />

The following authors have nothing to disclose: Craig Haifer, Lijia Yu, Julie Pavlovic,<br />

Paul Gow, Kumar Visvanathan, Adam Testro<br />

1225<br />

Quantiferon-CMV testing after cessation of antiviral<br />

therapy predicts late CMV reactivation after Liver Transplantation<br />

Siddharth Sood 2,1 , Craig Haifer 1 , Lijia Yu 3 , Julie Pavlovic 1 , Paul<br />

Gow 1 , Robert M. Jones 1 , Kumar Visvanathan 3 , Peter W. Angus 1 ,<br />

Adam Testro 1 ; 1 Department of Gastroenterology and Liver Transplant<br />

Unit, Austin Health, Heidelberg, VIC, Australia; 2 Department<br />

of Gastroenterology & Hepatology, Royal Melbourne Hospital,<br />

Melbourne, VIC, Australia; 3 Innate Immunity Laboratory, University<br />

of Melbourne, Melbourne, VIC, Australia<br />

Introduction CMV is the most common infection to reactivate<br />

following liver transplantation (OLT). Although antivirals are<br />

effective in preventing reactivation, some patients develop late<br />

CMV after prophylaxis or treatment has ceased. Quantiferon–<br />

CMV (QFN-CMV, Qiagen, USA) measures immune response<br />

towards CMV. We have previously shown that early non-reactive<br />

(n.r.) QFN-CMV (suggesting insufficient immunity) at 2<br />

weeks post-OLT predicts early CMV reactivation, usually within<br />

the first 2 months post-OLT. This study investigates whether<br />

n.r. QFN-CMV at 6 months (M6) is associated with late CMV<br />

after antiviral cessation. Methods 75 OLT recipients were<br />

monitored as part of a prospective blinded study. Absolute<br />

QFN-CMV1000 copies/mL). Treatment<br />

or prophylaxis was continued until M6. After M6, patients<br />

were monitored with CMV PCR monthly and treated for late<br />

CMV if they developed recurrent DNAemia. Fisher’s exact test<br />

was used. No organs were donated from prisoners. Results<br />

Fifty-six OLT recipients (D+ and/or R+) were monitored to 12<br />

months post-OLT. 47/56(84%) had a reactive QFN-CMV by<br />

M6. 24/56(43%) received prophylaxis or treatment before M6<br />

(either high-risk or developed early CMV DNAemia). 18/24<br />

had developed a reactive QFN-CMV by M6, of whom 78%<br />

had already developed a reactive QFN-CMV by 4 months.<br />

3/56(5%) of patients developed late CMV (median 251 days<br />

post-OLT). 1 asymptomatic patient commenced treatment with<br />

viral load 1100 copies/mL, while two developed CMV disease<br />

with very high viral loads (79,970 and 276,380 copies/mL).<br />

All 3 had a n.r. M6 QFN-CMV, suggesting they were high risk<br />

of late CMV following cessation of therapy. A M6 n.r. QFN-<br />

CMV was significantly associated with late CMV p=0.003,<br />

with an OR of 51.2 (sens 1.00, spec 0.89, PPV = 0.33, NPV =<br />

1.00). Conclusion Only 5% of recipients developed late CMV,<br />

but 2 of 3 suffered CMV disease. A n.r. M6 QFN-CMV is significantly<br />

associated with risk of late CMV, and QFN-CMV-ve<br />

patients likely require either extension of antiviral treatment<br />

or more regular monitoring. Conversely, M6 QFN-CMV has<br />

a high NPV, suggesting patients with robust ex-vivo immune<br />

response towards CMV at M6 can cease further monitoring.<br />

We also show that 78% of patients on antivirals may have<br />

protective immunity towards CMV before 4 months post-OLT,<br />

and antiviral prophylaxis could be discontinued early in this<br />

group with potential to reduce both drug side-effects and costs.<br />

Disclosures:<br />

Siddharth Sood - Grant/Research Support: Cellestis Ltd<br />

Robert M. Jones - Speaking and Teaching: Novartis, Astellas, Roche, Novartis,<br />

Astellas, Roche, Novartis, Astellas, Roche, Novartis, Astellas, Roche<br />

Peter W. Angus - Advisory Committees or Review Panels: Gilead Sciences, BMS;<br />

Grant/Research Support: Gilead sciences<br />

1226<br />

Everolimus is Associated with Lower Weight Gain Two<br />

Years Following Liver Transplantation – Results of a<br />

Randomized Multicenter Study<br />

Michael R. Charlton 1 , Mary E. Rinella 2 , Julie Heimbach 3 ; 1 Intermountain<br />

Medical Center, Salt Lake City, UT; 2 Northwestern University,<br />

Chicago, IL; 3 Mayo Clinic, Rochester, MN<br />

Background: The prevalence and severity of obesity increase<br />

following liver transplantation. Complications of obesity, including<br />

posttransplant metabolic syndrome, are an important cause<br />

of post-transplant morbidity and mortality. mTOR inhibitors have<br />

been shown to lead to lower body mass when compared to<br />

calcineurin inhibitors in animal <strong>studies</strong>. In contrast, calcineurin<br />

inhibitor-based immunosuppression is associated with steady<br />

weight gain in liver transplant recipients. Aims: To determine<br />

the comparative impact of mTOR inhibition on the course of<br />

post-transplant weight gain in humans following liver transplantation<br />

in the context of a randomized, controlled trial. Methods:<br />

Per protocol data on change in weight from baseline during<br />

a 24-month period were analyzed from a prospective, randomized,<br />

multicenter, open-label study, in which de novo liver<br />

transplant patients were randomized at 30 days to one of three<br />

immunosuppression protocols: 1) everolimus (EVR) + reduced<br />

tacrolimus (TAC; n = 245), or 2) TAC control (n = 243) or 3)<br />

TAC elimination (n = 231). Results: The treatment groups were<br />

well balanced in terms of age, gender, baseline weight, BMI,<br />

pre-transplant HCV infection and the incidence of diabetes.<br />

Randomization to TAC elimination was stopped prematurely<br />

due to a significantly higher rate of treated biopsy-proven acute<br />

rejection (tBPAR). Among patients who remained on treatment,<br />

mean increase in weight was significantly greater at month 12<br />

and 24 months from baseline in the TAC control arm than in the<br />

EVR + reduced TAC and the TAC elimination group (p=0.037-<br />

0.001, see table). Study medication was discontinued due to<br />

adverse events in 28.6% of EVR + reduced TAC and 18.2%<br />

of TAC control patients. Average daily prednisone doses were<br />

similar between treatment arms. The frequency of cardiovascular<br />

events and new onset diabetes mellitus was similar between<br />

treatment arms. Hyperlipidemia was more frequent in the EVR<br />

containing arms (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 817A<br />

1227<br />

Glycine is graft protective and preserves kidney function<br />

after liver transplantation: Data of the HEGPOL-Trial<br />

[ISRCTN69350312]<br />

Peter Schemmer 1 , Arash Nickkholgh 1 , Georgios Polychronidis 1 ,<br />

Steffen P. Luntz 2 , Ertan Mayatepek 3 , MarkusW. Buechler 1 , Ernst<br />

Klar 4 ; 1 General, Abdominal and Transplant Surgery, Ruprecht-<br />

Karls University of Heidelberg, Heidelberg, Germany; 2 Coordination<br />

Center for Clinical Trials, Ruprecht-Karls University of<br />

Heidelberg, Heidelberg, Germany; 3 Department of General Pediatrics,<br />

Heinrich-Heine-University,, Düsseldorf, Germany; 4 Department<br />

of Surgery, University of Rostock, Rostock, Germany<br />

Background. Glycine, a non-essential amino acid, prevents<br />

tissues from several types of injury in various experimental<br />

models. Thus the HEGPOL-Trial was designed to assess both<br />

the efficacy and safety of glycine in liver transplantation (LTx).<br />

Methods/Materials. A total of 130 patients undergoing primary<br />

LTx were randomized in a two-arm placebo-controlled<br />

multicenter double-blinded trial. While sixty six recipients<br />

were treated with i.v. glycine (4.4 %; 250 mL) controls (n=64)<br />

received injectable water (250 mL) during the anhepatic phase.<br />

The same infusion was repeated once a day for the following<br />

7 postoperative days (POD 1-7). Primary endpoints were peak<br />

levels of aspartate-aminotransaminase (AST), alanine-aminotransaminase<br />

(ALT) and graft survival. Secondary endpoints<br />

were liver injury based on histology, liver perfusion, AUC of<br />

both AST and ALT, early graft dysfunction and cyclosporine<br />

A-induced nephrotoxicity. Results. Per protocol analysis has<br />

shown no difference between groups; however, patients with<br />

cut off plasma glycine values of ≥7000 mmol/L (n=29) prior to<br />

reperfusion showed significantly lower serum ALT levels during<br />

the first 25 hours after LTx. Further the eGFR was significantly<br />

better with glycine during the study. Patient survival was 86.2%<br />

compared with 72.6% in controls (p=0.08). Conclusion.<br />

Although the per protocol analysis could not verify the hypothesized<br />

effects of glycine, very high plasma concentrations of<br />

glycine achieved after its i.v. administration at the anhepatic<br />

phase and early after liver transplantation proved not only to<br />

be safe, but also hepatoprotective and nephroprotective. Trial<br />

Registration: HEGPOL; ISRCTN69350312.<br />

Disclosures:<br />

The following authors have nothing to disclose: Peter Schemmer, Arash Nickkholgh,<br />

Georgios Polychronidis, Steffen P. Luntz, Ertan Mayatepek, MarkusW.<br />

Buechler, Ernst Klar<br />

1228<br />

The Combination of Strongly (+) C4d Staining and<br />

High Titer Donor Specific Antibody (DSA) Suggests that<br />

Plasma Cell Hepatitis is a Result of Antibody-Mediated<br />

Rejection<br />

M. Isabel Fiel 1 , Sander S. Florman 2 , Thomas D. Schiano 2 ; 1 Department<br />

of Pathology, Icahn School of Medicine at Mount Sinai, New<br />

York, NY; 2 Recanati-Miller Transplant Institute, Icahn School of<br />

Medicine, New York, NY<br />

Background and Aim: Plasma cell hepatitis (PCH), is a form<br />

of post liver transplant (LT) allograft dysfunction. Antibody-mediated<br />

rejection (AMR) is being increasingly recognized as a<br />

cause of graft failure and is typified by high DSA titers, and<br />

histologically by (+) C4d immunostaining. Since PCH is felt to<br />

be a variant of rejection, we sought to determine whether it is<br />

a result of AMR. Design: PCH (characterized by >30% plasma<br />

cell infiltrate and severe necroinflammation) cases were identified;<br />

cases of acute (ACR) and chronic rejection (CR) served as<br />

controls. Immunostaining for C4d (ARP, Waltham MA; 1:200<br />

dilution) was performed on FFPE tissue using a Dako Autostainer.<br />

Assessment of C4d staining was performed blindly by<br />

a single hepatopathologist using the following scoring system:<br />

0=negative, 1=equivocal; 2 = 10-50% (+); 3 = moderate/<br />

strong diffuse (+) in endothelial cells lining the portal venules<br />

and portal capillaries. Testing for DSA was performed for each<br />

patient; an MFI >2000 was considered significant (+). Results:<br />

A total of 30 (15 PCH, 15 control) cases were included; 9/15<br />

(60%) PCH, 4/15 (27%) controls were (+) for both C4d and<br />

DSA (p=0.065). A combination of strong C4d staining and<br />

high DSA MFI (DSA+) was associated with the diagnosis of<br />

PCH when compared to C4d+/DSA- cases (p=0.03). Six of 9<br />

(67%) PCH cases and 2/4 controls had (+) HLA Class 2 DSA<br />

alone; 2/9 PCH and 2/4 controls had both HLA Class 1 and<br />

2 (+) DSA (See Table). Four PCH patients had been receiving<br />

interferon therapy for recurrent HCV. One control case was<br />

documented to have rejection one week post-LT and had strong<br />

(3+) C4d staining and extremely high DSA MFI titers (class<br />

1=21167, class 2=22276) suggestive of true AMR. Conclusions:<br />

Patients with PCH have high titer DSA and strongly positive<br />

C4d staining of their liver allograft biopsy. Although some<br />

cases of ACR and CR may have high DSA MFI, C4d staining<br />

is less frequent and has less intensity in comparison to PCH<br />

cases. These findings suggest that PCH is a result of AMR. All<br />

HCV patients diagnosed with PCH while on interferon therapy<br />

showed (+) C4d staining and high DSA, suggesting that AMR<br />

played a role in their allograft dysfunction. Based on this data,<br />

the histologic diagnosis of PCH should prompt a work-up for<br />

AMR.<br />

Disclosures:<br />

Thomas D. Schiano - Advisory Committees or Review Panels: salix, merck, gilead,<br />

pfizer; Grant/Research Support: galectin, massbiologics, biotest<br />

The following authors have nothing to disclose: M. Isabel Fiel, Sander S. Florman<br />

1229<br />

The coming of age of the graft in pediatric orthotropic<br />

liver transplant (OLT) recipients<br />

Shewit Giovanni, J. Michael Millis, Andrew I. Aronsohn, John F.<br />

Renz, Ruba K. Azzam, Pamela Boone, Kathleen Dasgupta, Helen<br />

S. Te; University of Chicago Medicine, Chicago, IL<br />

Background: While graft survival is increasing in adults and<br />

children after OLT, data on long term graft function >15 years<br />

after OLT is scant. Aim: To describe long term graft and patient<br />

outcomes (>15 years) of patients who had OLT at 15<br />

years prior were identified from the OLT database. Patient<br />

demographic and laboratory data, as and radiographic and<br />

histologic data when available, were collected. Current immunosuppression<br />

and clinical outcomes were also noted. Results:<br />

25 patients who received OLT at a median age of 3.9 years<br />

(range, 0.2-16.2) were included in the study. Biliary atresia<br />

was the indication for OLT in 48% of cases, followed by inborn<br />

errors of metabolism in 16%. 22 (88%) patients received cadaveric<br />

grafts, of which 16 were whole and 6 were partial. The<br />

median duration of follow-up was 21.4 yrs (range, 15.1-28.1).<br />

Current immunosuppression were tacrolimus-based in 60%,<br />

cyclosporine-based in 12%, sirolimus alone in 8%, prednisone<br />

alone in 4%, and none in 16%. The latest laboratory tests are<br />

summarized in Table 1. 68% of patients had abnormal liver<br />

tests, 47% of which was cholestastic, 18% hepatocellular and


818A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

the rest was a mixed pattern. Radiologic cholangiopathy was<br />

seen in 9 (36%) of patients. Of 17 patients who had liver biopsies,<br />

chronic rejection was seen in 6 and nonspecific hepatitis<br />

in 4; stage 3-4 fibrosis was present in 7 patients. A total of<br />

76% of patients had either liver test, radiographic, or histologic<br />

abnormalities of the graft. There were 4 graft failures, leading<br />

to death in 2 patients and a repeat OLT in 1, while one is<br />

awaiting repeat OLT. Another patient died from sepsis related<br />

to hepatic abscesses. Conclusions: Liver grafts in children continue<br />

to function >15 years later in adulthood, but abnormal<br />

liver tests, radiographic or histologic findings are present in<br />

majority of these grafts. Chronic rejection was common, and<br />

graft failure that required repeat OLT or led to death occurred<br />

in a few patients. The presence of donor-specific antibodies in<br />

the patients with graft abnormalities will be investigated further.<br />

Disclosures:<br />

J. Michael Millis - Advisory Committees or Review Panels: Vital Therapies; Board<br />

Membership: Vital Therapies<br />

Helen S. Te - Advisory Committees or Review Panels: BMS; Grant/Research<br />

Support: Abbvie, Conatus, BMS<br />

The following authors have nothing to disclose: Shewit Giovanni, Andrew I.<br />

Aronsohn, John F. Renz, Ruba K. Azzam, Pamela Boone, Kathleen Dasgupta<br />

1230<br />

Lack of effect of CMV reactivation in the outcome of<br />

HCV-infected liver transplant(LT)patients<br />

Victoria Aguilera 1,2 , Tommaso Di Maira 1 , Isabel Conde 3 , Angel<br />

Rubin 1,2 , Carmen Vinaixa 1,2 , Salvador Benlloch 1,2 , Martin Prieto<br />

1,2 , Marina Berenguer 4,2 ; 1 Hepatology Section, Hospital Universitari<br />

La Fe, Valencia, Spain; 2 CIBERehd, Instituto de Salud Carlos<br />

III, Grant PI13/01770, Valencia, Spain; 3 Hepatologia, Instituto<br />

de Investigación Sanitaria Hospital Universitari La Fe, Valencia,<br />

Spain; 4 Facultad de Medicina, Universidad de Valencia, Valencia,<br />

Spain<br />

Background: The role of CMV reactivation in recurrent hepatitis<br />

C after LT is controversial. Other aspects such as development of<br />

infections, diabetes mellitus (DM), rejection and cardio-vascular<br />

events (CVE) have also been associated with CMV reactivation<br />

post-LT. Aims: To evaluate if CMV reactivation is associated<br />

with (1) aggressive recurrent hepatitis C and (2) post-LT infections,<br />

rejection, DM and CVE. Patients and methods: Patients LT<br />

from Jan 2011 to Dec 2013 due to HCV-cirrhosis with HCV-positive<br />

viremia. CMV Reactivation was defined as CMV viral<br />

load(VL) >400c/ml. CMV VL was determined in a clinical basis<br />

during the first 12 months after LT. Outcome variables were:<br />

(i) aggressive recurrent hepatitis C (defined as F>1, F=0 with<br />

moderate inflammatory activity or cholestatic hepatitis), acute<br />

rejection (Banff criteria), infections, de novo DM post-LT and<br />

CVE. Demographics, donor, immunosuppression(IS), surgical,<br />

hematological, biochemical and virological related variables<br />

were also recorded. Results: 77 patients (mean age:56yr (25-<br />

68), 67% men, 46% HCC) were included. Donor and recipient<br />

CMV serologies were as follows: D+/R+:n=61,79%; D -/<br />

R+:n=6,8%; D+/R-:4, 5%; D-/R-:n=6, 8%. CMV prophylaxis<br />

(during 90 days) was administered in 7(9%) patients. Reasons<br />

for prophylaxis were: CMV-D +/R - n=3, Methylprednisolone<br />

boluses or basiliximab use, n=4. CMV reactivation occurred in<br />

26 patients(33%), 24(34%) without prophylaxis and 2(28%)<br />

with prophylaxis (p=ns). Time to reactivation was 138 vs 45<br />

days in those with and without prophylaxis (P=.026). Reactivation<br />

led to anti-CMV treatment in 16(61%) pts. Severe recurrent<br />

hepatitis C occurred in 55 pts (71%), 9 (12%) pts developed<br />

rejection, 39 pts (51%) infections, 15(20%) de novo DM and<br />

6 (8%) a CVE. Nor CMV reactivation nor CMV treatment were<br />

associated with aggressive recurrent hepatitis C, rejections,<br />

infections, de novo DM or CVE. Graft and patient survival were<br />

not related either with CMV reactivation. There was a trend<br />

towards a potential protection effect of sirulimus-SRL- based IS<br />

and CMV reactivation (20% in SRL group vs 39% in non SRL<br />

group, p=0.1) but triple therapy or basiliximab use were not<br />

associated with reactivation. Conclusions: CMV reactivation<br />

occurs frequently after LT. About half of cases require therapy.<br />

In patients with prophylaxis, CMV reactivation tends to occur<br />

at later time points. Given the lack of association between<br />

CMV reactivation and post-transplant outcome, we believe it<br />

is unnecessary to prolong CMV prophylaxis for more than 3<br />

months, as suggested by some authors. mTor inhibitors seem to<br />

protect against CMV reactivation; larger <strong>studies</strong> are needed to<br />

confirm this finding..<br />

Disclosures:<br />

The following authors have nothing to disclose: Victoria Aguilera, Tommaso Di<br />

Maira, Isabel Conde, Angel Rubin, Carmen Vinaixa, Salvador Benlloch, Martin<br />

Prieto, Marina Berenguer<br />

1231<br />

Blood Phosphatidyl ethanol identifies alcohol use<br />

among liver transplant recipients<br />

Michael F. Fleming 2 , Jenny Vue 1 , Michael R. Lucey 1 ; 1 Medicine,<br />

Univ. of Wisconsin School of Medicine and Public Health, Madison,<br />

WI; 2 Family Medicine, Northwestern University, Chicago, IL<br />

All patients with liver failure due to alcohol use disorder (AUD)<br />

who undergo liver transplantation (LT) are advised to abstain<br />

from alcohol. Monitoring alcohol use in AUD -LT recipients is<br />

hampered by lack of easily usable methods to identify alcohol<br />

relapses. Phosphatydylethanol (PEth), a red cell-membrane<br />

phospholipid that is released into circulation after alcohol use,<br />

accurately identifies recent alcohol use (i.e. in the prior 30<br />

days) in healthy subjects and AUD patients. The whole blood<br />

concentration of PEth varies in a dose-response fashion with<br />

consumption, with >8ng/ml being considered a positive test<br />

for recent exposure. There are no known false positive tests<br />

and PEth is considered 100% specific. PEth can detect daily<br />

drinking of > 2 drinks or intermittent binge drinking. We report<br />

a prospective study of blood PEth as a marker of alcohol use<br />

among 213 LT recipients at 2 US liver transplant centers. All<br />

subjects gave informed consent. The protocol was approved by<br />

the IRBs at each institution. Methods: The sample included 147<br />

men and 66 women with a mean age of 59 years. 131 had<br />

a history of alcohol dependence prior to LT and 82 served as<br />

non-alcohol dependent controls. Subjects participated in a face<br />

to face interview to assess recent alcohol use using a 30-day<br />

timeline follow-back calendar procedure and dried blood spot<br />

collection at study entry, and at 6 and 12 months thereafter.<br />

PEth was determined using an Agilent Zorbax chromatography-tandem<br />

mass spectrometry procedure and was performed<br />

by the United States Drug Testing Laboratories in Chicago.<br />

Results: among the available data, 49 subjects (23% of the<br />

total sample), including 39 AUD subjects (30% of the AUD<br />

group) and 10 controls (12% of the control group), had a positive<br />

test of >8 ng/ml, either at baseline or during the follow-up<br />

period. 149 patients reported no alcohol use and had a negative<br />

PEth assays. 32 subjects (15% of the total sample) reported<br />

no alcohol use and had a positive test indicating recent alcohol<br />

use. Of the 25 subjects who reported recent alcohol use, 11


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 819A<br />

had a positive test and 14 were negative. The 14 drinking subjects<br />

who were PEth-negative reported low levels of use. This<br />

study confirms that, despite advice to abstain, more than one<br />

quarter of alcohol-addicted LT recipients return to alcohol use.<br />

PEth identifies alcohol use in LT recipients who deny drinking.<br />

Liver transplant centers may want to include PEth as part of<br />

routine post-LT care in order to detect alcohol use in recipients<br />

who deny drinking and thereby provide them with early interventions,<br />

designed to restore sobriety.<br />

Disclosures:<br />

Michael R. Lucey - Advisory Committees or Review Panels: Galectin; Grant/<br />

Research Support: Vertex, Abbvie, Gilead, Salix<br />

The following authors have nothing to disclose: Michael F. Fleming, Jenny Vue<br />

1232<br />

A New Score based on explant pathology allows an<br />

individualized prediction of HCC-recurrence after liver<br />

transplantation<br />

Charlotte E. Costentin 1 , Giuliana Amaddeo 1 , Christophe Duvoux 1 ,<br />

Julien Calderaro 2 , Alexis Laurent 3 , Ariane Mallat 1 , Françoise<br />

Roudot-Thoraval 4 ; 1 Hepatology, Hôpital Henri Mondor, Creteil,<br />

France; 2 pathology, Henri Mondor Hospital, Creteil, France;<br />

3 Surgery, Henri Mondor Hospital, Creteil, France; 4 Public Health,<br />

Henri Mondor Hospital, Creteil, France<br />

Aim: After liver transplantation (LT) for hepatocellular carcinoma<br />

(HCC), a standardized prediction of recurrence would<br />

valuable in order to define patients who might benefit from post<br />

LT adjuvant therapy or early changes in immunosuppressive<br />

regimens. We recently compared the accuracy of 4 explantbased<br />

models for prediction of recurrence (EASL 2015). In<br />

this study, the “Up to seven model”, using 2 variables (size<br />

of largest nodule and number of nodules) appeared to be the<br />

best tool to identify patients with a high risk of recurrence at<br />

5 years post LT. The aim of this study was to design a new<br />

explant-based model including vascular invasion and tumor differentiation<br />

as predictors of recurrence and to test its accuracy<br />

against the “Up to seven” model. Methods: Following uni- and<br />

multi-variate analysis of pathological predictors of recurrence<br />

in a series of 372 liver explants of patients transplanted for<br />

HCC between 2003 and 2005 and followed prospectively for<br />

5 years, we built 2 predictive models. New score 1 included<br />

all independent pathological predictors (number of nodules,<br />

size of the largest, tumor differentiation, micro/macrovascular<br />

invasion, tumor burden (uni or bilobar)) and ranged from 0 to<br />

9 points. New score 2 included all above mentioned variables<br />

except tumor differentiation (a variable that is not available<br />

when tumor is totally inactivated after bridging therapy) and<br />

ranged from 0 to 11 points. The accuracy of these 2 scores<br />

and the “Up to seven” model without vascular invasion to predict<br />

5 year-HCC recurrence was assessed by comparison of<br />

AUCs of ROC curves which were subsequently compared using<br />

the Hanley&McNeil method. Results: The “Up to seven model”<br />

identified two distinct risk groups for 5 year-recurrence of HCC<br />

(10.8±2.2% if score ≤7 ; 54.5±4.5% si score >7). New scores<br />

1 and 2 identified 4 groups of patients with different levels<br />

of risks of recurrence ranging from 10 to >70% at 5 years.<br />

AUCs were 0.7915 [0.73394 - 0.84907] for the “Up to seven<br />

model, 0.7881 [0.73135 - 0.8448] for New score 1 and<br />

0.7889 [0.7352 - 0.8427] for New score 2 (p=0.743). Tumor<br />

differentiation did not improve accuracy to predict 5 year-HCC<br />

recurrence (New score 1 vs New score 2, p=0.49). Using<br />

New score 2, a nomogram was built, giving the probability of<br />

recurrence at 1, 3 and 5 years post transplantation, according<br />

to the score value (fig) Conclusion: A new score using number<br />

of nodules, size of the largest, vascular invasion and tumor<br />

burden on explant has similar accuracy compared to the “Up<br />

to seven” model to predict 5 year-HCC recurrence but allows a<br />

more individualized prediction of recurrence based on patients’<br />

profile of explant findings.<br />

Disclosures:<br />

Christophe Duvoux - Advisory Committees or Review Panels: Novartis, Roche,<br />

Novartis, Roche, Novartis, Roche, Novartis, Roche; Speaking and Teaching:<br />

Astellas, Astellas, Astellas, Astellas<br />

Françoise Roudot-Thoraval - Advisory Committees or Review Panels: Roche; Consulting:<br />

LFB biomedicaments; Speaking and Teaching: gilead, Janssen, BMS,<br />

Roche<br />

The following authors have nothing to disclose: Charlotte E. Costentin, Giuliana<br />

Amaddeo, Julien Calderaro, Alexis Laurent, Ariane Mallat<br />

1233<br />

Comparative analysis of renal dysfunction in recipients<br />

with NASH compared to controls with alcoholic liver<br />

disease after liver transplantation<br />

Cheri Ogwo 1 , Satheesh Nair 2,1 , Jason Vanatta 2,1 , Vinaya Rao 2,1 ,<br />

Chibuzor Iwelu 1 , James Eason 2,1 , Sanjaya K. Satapathy 2,1 ; 1 Surgery,<br />

University of Tennessee Health Sciences Center, Memphis,<br />

TN; 2 Surgery, Methodist University Hospital, Memphis, TN<br />

BACKGROUND: Association between NAFLD/NASH and<br />

chronic kidney disease (CKD) has been well recognized.<br />

Release of pro-inflammatory mediators from steatotic liver or<br />

through the contribution of metabolic syndrome (Obesity, type<br />

2 DM, dyslipidemia, and hypertension) has been postulated<br />

as the etiology. AIM: We aim to assess if the predisposition<br />

for renal dysfunction persists after liver transplantation (LT) in<br />

recipients with NASH compared to controls with alcoholic liver<br />

disease (ALD). METHODS: Charts of 806 adult LT recipients<br />

(6/2006- 12/2012) reviewed; 108 recipients with NASH,<br />

and 122 with ALD were identified. After exclusion of SLK transplants<br />

(9/11), re-transplants (3/5), death ≤ 90 days (3/7); 92<br />

NASH recipients were compared with 99 recipients with ALD.<br />

eGFR before LT, at 3 months, 1 yr, and 3 yrs. were compared.<br />

RESULTS: NASH patients are more likely to be older (57.7 yrs.<br />

vs 55 yrs., p=0.01), obese (32±5.4 vs. 28.4±5.6, p=0.001),<br />

and diabetic (42.4% vs. 38.4%, p=0.003), and have lower<br />

MELD score at listing (20.63±7.32 vs. 22.08±6.12, p=0.04).<br />

Both groups received our standard steroid-free protocol with<br />

rabbit ATG induction with low-dose Tacrolimus. eGFR pre-LT in<br />

NASH and ALD group were comparable (76±45 vs. 78±33,<br />

p=0.19). A significant decline in eGFR (Figure 1) was noted<br />

in NASH recipients at 3 mos. (61±30 vs. 69±36, p=0.15), 1<br />

yr. (60±20 vs. 70±25, p=0.03), and 3 yrs. (51±17 vs.68±18,<br />

p=0.001). Based on eGFR groups (Gr 1: eGFR < 45, Gr II:<br />

eGFR ≥45-60, eGFR >60), a distinct pattern of either inferior<br />

recovery of renal function or decline in renal function was noted<br />

in NASH recipients post LT (Figure 1). Declining eGFR in the<br />

NASH group was more likely in diabetics (46±16 vs. 67±23,<br />

p=0.02), but not in non-diabetics (57±17 vs. 68±17, p=0.17)<br />

at 3 yrs. of follow up. CONCLUSION: LT recipients with NASH<br />

phenotype have an increased propensity for renal dysfunction<br />

compared to ALD particularly in diabetics. Early conversion to<br />

calcineurin inhibitor sparing agents in LT recipients with NASH<br />

and diabetes should be considered as a strategy to preserve<br />

renal function.


820A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Satheesh Nair - Advisory Committees or Review Panels: Jansen; Grant/Research<br />

Support: Gilead; Speaking and Teaching: Abbvie, Valeant, BMS<br />

Sanjaya K. Satapathy - Advisory Committees or Review Panels: Gilead, Intercept;<br />

Grant/Research Support: Genfit, Gilead, Biotest<br />

The following authors have nothing to disclose: Cheri Ogwo, Jason Vanatta,<br />

Vinaya Rao, Chibuzor Iwelu, James Eason<br />

1234<br />

Long-Term Outcome of Extrahepatic Biliary Atresia into<br />

adult life<br />

Javaid Sadiq 1 , Aditi Kumar 4 , Hifsa Sohail 3 , Carla Lloyd 3 , James<br />

W. Ferguson 4 , Darius Mirza 2 , Khalid Sharif 1 , Gideon Hirschfield 4 ,<br />

Deirdre A. Kelly 3 ; 1 Paediatric Hepatobiliary Transplant Surgery,<br />

Birmingham Children’s Hospital Birmingham UK, Birmingham,<br />

United Kingdom; 2 Liver Surgery, University Hospital Birmingham,<br />

Birmingham, United Kingdom; 3 Liver unit, Birmingham Children’s<br />

Hospital, Birmingham, United Kingdom; 4 Liver unit, University Hospital<br />

Birmingham, Birmingham, United Kingdom<br />

Background:Biliary Atresia(BA) is the single commonest cause<br />

of neonatal cholestasis leading to cirrhosis,portal hypertension<br />

and liver failure and is the main indication for pediatric liver<br />

transplant(LT). Aim:Evaluate the long-term outcome of children<br />

with BA transitioning to adult life Subjects & Methods:Records<br />

of patients of BA managed over a period of 34 years(1980-<br />

2014) at a single institution were retrospectively reviewed<br />

.Patients with more than 10 years of follow-up were included in<br />

the study.Data collection included demographics,age at Kasai<br />

Portoenterostomy(KPE),associated malformations, survival with<br />

native liver or post-LT,mortality, current education/work/marital/family<br />

status. Results: 493 BA patients were managed<br />

during this period(260 F & 233 M).Median age at kasai was<br />

53 days(range:7-183 days). 92 % had isolated BA while 8 %<br />

had BA polysplenia malformation syndrome.332 patients were<br />

included in this study (1980 – 2004). 11 patients were lost to<br />

follow-up. Median patient survival is 17.3 yrs(0.32- 34.6) &<br />

median survival with native liver is 2.25 yrs(0.07-34.6). 53<br />

patients(16.5%) died in pediatric care; 26 with their native<br />

livers & 27 after LT.135 patients(50.3%) are still in pediatric<br />

care(Group A).57 are surviving with their native liver(A1)<br />

while 78 children have been transplanted(A2).7 patients are<br />

awaiting transplant in Group A1.133(49.6 %) patients were<br />

transferred to adult services(Group B); 49 with native livers(B1)<br />

and 84 after LT(B2). 28 patients in group B1 had portal hypertension(PH);20<br />

treated with beta blockers,esophageal banding<br />

or shunts. 9 patients transferred to adult services with native<br />

liver(B1) subsequently required LT & 7 are listed for LT due to<br />

decompensated liver disease.6 patients in group B2 required<br />

retransplant. After transfer to adult care, 3 patients in Group B1<br />

died( one due to ruptured splenic aneurysm & 2 due to decompensated<br />

liver disease) while 5 patients in Group B2 died from<br />

post-transplant lymphoproliferative disorders(PTLD),Hepatopulmonary<br />

syndrome, ruptured psoas cyst and bleeding & chronic<br />

rejection).Out of 268 patients in this series, majority participated<br />

in normal school education while 32(12 %) required<br />

special needs support. 29 transferred went to university, 18<br />

obtained non-vocational qualifications and 33 joined various<br />

training courses. Conclusion: Improved medical and surgical<br />

techniques have improved the outcome and quality of life for<br />

patients with BA, allowing them to live into adult life, complete<br />

their education & function as useful members of the society<br />

Disclosures:<br />

James W. Ferguson - Advisory Committees or Review Panels: Astellas, Novartis<br />

Gideon Hirschfield - Advisory Committees or Review Panels: Intercept Pharma;<br />

Consulting: Dignity Sciences, GSK, NGM Bio, Lumena, J & J; Grant/Research<br />

Support: BioTie; Speaking and Teaching: Falk Pharma<br />

Deirdre A. Kelly - Consulting: Sanofi Pasteur; Grant/Research Support: BMS,<br />

Astellas, Acitllion, MSD, Roche<br />

The following authors have nothing to disclose: Javaid Sadiq, Aditi Kumar, Hifsa<br />

Sohail, Carla Lloyd, Darius Mirza, Khalid Sharif<br />

1235<br />

Assessment of Energy Metabolism in Liver Transplant<br />

Recipients<br />

Ajay Singhvi 1 , H. Steven Sadowsky 2 , Ayelet Cohen 1 , Alysen<br />

Demzik 1 , Mary E. Rinella 1 , Lisa B. VanWagner 1 , Josh Levitsky 1 ;<br />

1 Division of Gastroenterology & Hepatology, Northwestern Memorial<br />

Hospital, Chicago, IL; 2 Physical Therapy & Human Movement<br />

Sciences, Feinberg School of Medicine, Chicago, IL<br />

Background: The reasons for accelerated weight gain and metabolic<br />

syndrome following liver transplantation (LT) are not<br />

well understood. We hypothesized that this may be related<br />

to ineffective metabolic rates compared to healthy patients,<br />

particularly during exercise. The purpose of this pilot study was<br />

to assess energy expenditure during exercise in LT recipients.<br />

Methods: We consented ten subjects who were transplanted for<br />

NASH or cryptogenic cirrhosis (>1 year post) to undergo analysis<br />

of body composition, resting energy expenditure (REE), and<br />

peak oxygen uptake (VO 2 max<br />

). The latter was conducted using<br />

a ramped-Bruce protocol. VO 2 max<br />

was assessed by standard<br />

parameters, including respiratory exchange ratio and a rating<br />

of perceived exertion. Predicted VO 2<br />

was determined using the<br />

Wasserman-Hansen equations and REE was determined using<br />

the Harris-Benedict equation. Measured VO 2 max<br />

values were<br />

then compared to data from the NHANES database for ageand<br />

sex-matched healthy individuals. Results: Subjects (50%<br />

male) were mean age 60.1±4.3 years, BMI 30.7±3.22 kg/<br />

m 2 , time post-LT 5.2±2.6 years. Average percent body fat was<br />

22.6±8.4% in males and 28.5±2.9% in females. Mean REE<br />

was 97.3% of predicted REE in males and 77.9% of predicted<br />

REE in females. Average VO 2 max<br />

(mL/kg/min) in LT males was<br />

16.9±5.0 compared to a predicted value of 21.0±3.1 (80.4%<br />

of predicted VO 2 max<br />

, p=0.04). LT females had a mean VO 2<br />

max<br />

(mL/kg/min) of 15.1±2.2 compared to a mean predicted<br />

value of 20.1±1.1 (75.4% of predicted VO 2 max<br />

, p=0.004).<br />

These values were lower than those of historical healthy controls<br />

who had a mean VO 2 max<br />

(mL/kg/min) of 31 in males<br />

and 23 in females (Figure 1). Discussion: Our results suggest<br />

that LT recipients, particularly females, have lower resting and<br />

exercise energy expenditure compared to predicted values and<br />

healthy controls. This may contribute to post-LT weight gain and<br />

inability to lose adequate weight with diet and exercise. Further<br />

study is necessary to investigate the pathophysiology of this<br />

phenomenon, the metabolic change from pre- to post-LT, and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 821A<br />

interventional exercise/weight loss trials tailored to LT recipients’<br />

metabolic capacities.<br />

Disclosures:<br />

Josh Levitsky - Consulting: Transplant Genomics Incorporated<br />

The following authors have nothing to disclose: Ajay Singhvi, H. Steven Sadowsky,<br />

Ayelet Cohen, Alysen Demzik, Mary E. Rinella, Lisa B. VanWagner<br />

1236<br />

Retained HLA Class II Donor Specific Antibody after<br />

Liver Transplantation Increases Risk of Acute Cellular<br />

Rejection: A Case Control Study<br />

Mohamed Elfeki 1,3 , Mahmoud El-Bendary 3 , Petrina V. Genco 2 ,<br />

Raghda Farag 3 , Youssef Mosaad 4 , Justin H. Nguyen 1 , Denise M.<br />

Harnois 1 , Surakit Pungpapong 1 ; 1 Transplant, Mayo Clinic, Jacksonville,<br />

FL; 2 Laboratory Medicine and Pathology, Mayo Clinic,<br />

Jacksonville, FL; 3 Tropical Medicine and Hepatology, Mansoura<br />

University, Faculty of Medicine, Mansoura, Egypt; 4 Clinical pathology<br />

and Immunology, Mansoura University, Faculty of Medicine,<br />

Mansoura, Egypt<br />

Background: Liver is an organ that is well-known for its immune<br />

tolerant capacity. Conflicting results regarding liver graft<br />

function and rejection risk affected by pre-transplant positive<br />

lymphocyte cross match (LCM) or post-transplant circulating<br />

donor specific antibody (DSA) have been reported according<br />

to individual institutions. We aimed to evaluate the clinical<br />

significance of positive LCM coupled with positive DSA at time<br />

of and following orthotopic liver transplant (OLT) on outcome<br />

of graft function as regards to biopsy-proven acute cellular<br />

rejection (BPACR) and early allograft dysfunction (EAD). EAD<br />

was defined as the presence of one or more of the following<br />

postoperative laboratory analyses reflective of liver injury and<br />

fuction: bilirubin ≥10mg/dl on day 7, International normalized<br />

ratio ≥1.6 on day 7, and alanine or aspartate aminotransferases<br />

〉2000 IU/L within the first 7 days. Methods: This is a<br />

retrospective case control study of prospectively collected data<br />

from 586 OLT performed at our institution between December<br />

2009 and December 2013. Total of 37 cases with positive<br />

LCM and DSA at the time of OLT were identified. Each case<br />

was matched to two controls (74 controls) based on diagnosis<br />

of liver disease, age and sex of recipients. Results: Majority of<br />

the cases (90%) were females and they had a higher number<br />

of parity when compared to female controls (mean 2.6 vs.<br />

1.7, p=0.003). BPACR occurred significantly more among the<br />

cases than the controls (51% vs. 24%, p=0.006) with an odds<br />

ratio (OR) of 3.28 and 95% confidence interval (CI) of 1.42-<br />

7.57. Higher number of the cases experienced EAD compared<br />

to the controls (22% vs. 9%, p=0.08) with an OR of 2.64 (95%<br />

CI = 0.88-7.96). Despite similar immunosuppressive regimen<br />

per our institutional protocol, the cases who retained HLA class<br />

II DSA post-OLT experienced BPACR more often than the controls<br />

(80% vs. 24%, p


822A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tion and patient and graft survival. Implications for the clinical<br />

setting are conscientiously monitoring of the GFR in the first<br />

months after liver transplantation and taking action as early<br />

as possible to prevent worsening of the creatinine clearance.<br />

Disclosures:<br />

Faouzi Saliba - Advisory Committees or Review Panels: Novartis, Roche, Genzyme,<br />

Vital therapies; Grant/Research Support: Astellas; Speaking and Teaching:<br />

Gambro, MSD, Gilead<br />

Rene Adam - Grant/Research Support: Merck Serono, Roche, Sanofi aventis,<br />

astellas, novartis, Merck Serono, Roche, Sanofi aventis, astellas, novartis, Merck<br />

Serono, Roche, Sanofi aventis, astellas, novartis, Merck Serono, Roche, Sanofi<br />

aventis, astellas, novartis<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

The following authors have nothing to disclose: Noémie Minczeles, Valérie Delvart,<br />

Teresa Antonini, Rodolphe Sobesky, Philippe Ichai, Audrey Coilly, Bruno<br />

Roche, Marc Boudon, Eric Vibert, Denis X. Castaing, Daniel Cherqui, Gilles<br />

Pelletier, Jean-Charles Duclos-Vallee<br />

1238<br />

Value of Magnetic Resonance Cholangiography in<br />

Assessment of Non-Anastomotic Biliary Strictures after<br />

Liver Transplantation<br />

A. Claire den Dulk 1 , Martin N.J.M. Wasser 2 , Francois Willemssen<br />

3 , Melanie A. Monraats 2 , Marianne de Vries 3 , Rivka van den<br />

Boom 2 , Jan Ringers 4 , Hein W. Verspaget 1 , Herold J. Metselaar 5 ,<br />

Bart van Hoek 1 ; 1 Gastroenterology and Hepatology, Leiden University<br />

Medical Center, Leiden, Netherlands; 2 Radiology, Leiden<br />

University Medical Center, Leiden, Netherlands; 3 Radiology, Erasmus<br />

Medical Center, Rotterdam, Netherlands; 4 Transplant Surgery,<br />

Leiden University Medical Center, Leiden, Netherlands; 5 Gastroenterology<br />

and Hepatology, Erasmus Medical Center, Rotterdam,<br />

Netherlands<br />

Background Non-anastomotic biliary strictures (NAS) remain<br />

a frequent complication after orthotopic liver transplantation<br />

(OLT). The aim of this study was to evaluate whether Magnetic<br />

Resonance Cholangiography (MRCP) could be used to detect<br />

or exclude NAS and grade the severity of biliary strictures.<br />

Methods In total, 58 patients after OLT from two transplantation<br />

centres in whom endoscopic (ERCP) or percutaneous transhepatic<br />

cholangiography (PTC) and MRCP were performed<br />

within less than 6 months were included in the study. Of these<br />

patients, 41 had NAS and 17 were without NAS based on<br />

ERCP or PTC and follow-up. Four radiologists – two in each<br />

center – used an adapted validated classification –termed<br />

Leiden Biliary Stricture Classification (LBSC)– to evaluate the<br />

MRCP independently and assess NAS severity on a scale from<br />

0 to 3 points in three hepatobiliary regions.(Fig 1) A maximum<br />

of 15 points could be obtained. Interobserver agreement of the<br />

severity score and intra-observer agreement between ERCP/<br />

PTC and MRCP for each region was calculated with the kappa<br />

(κ) statistic. Results Optimal cut-off value of the LBSC to detect<br />

the presence of NAS with MRCP was calculated at ≥ 3 points<br />

for all readers. Applying this cut-off, sensitivity for each reader<br />

was >90%, with a corresponding specificity of 50-82%, positive<br />

predictive value (PPV) of 86-91%, and negative predictive<br />

value (NPV) of 80-100%. When the cut-off value was applied<br />

to the radiologists’ mean scores sensitivity was 98%, specificity<br />

65%, PPV 87% and NPV 92%. MRCP performance was better<br />

in evaluation of the intrahepatic bile ducts than of the extrahepatic<br />

bile ducts. The additional value of MRCP for grading<br />

severity (κ= 0.2 – 0.7) and localizing NAS (κ= 0.2 – 0.9) was<br />

limited. Conclusion MRCP is a reliable tool to detect or exclude<br />

non-anastomotic biliary strictures after OLT. MRCP cannot be<br />

used to reliably grade the severity of these strictures.<br />

Disclosures:<br />

Bart van Hoek - Advisory Committees or Review Panels: Janssen-Cilag, Bristol<br />

Meyers Squib, Gilead, Merck, Abbvie<br />

The following authors have nothing to disclose: A. Claire den Dulk, Martin N.J.M.<br />

Wasser, Francois Willemssen, Melanie A. Monraats, Marianne de Vries, Rivka<br />

van den Boom, Jan Ringers, Hein W. Verspaget, Herold J. Metselaar<br />

1239<br />

Are cirrhotic patients awaiting liver transplantation protected<br />

against vaccine preventable diseases?<br />

Mazzola Alessandra 1,2 , Margherita Tran Minh 1,3 , Raluca Pais 1 ,<br />

Pascal Lebray 1 , Denis Bernard 4 , Claire Goumard 5 , Yvon Calmus 1 ,<br />

Filomena Conti 1 ; 1 Unité Médicale de Transplantation Hépatique,<br />

APHP, Hôpital Pitié Salpétrière, Paris, France; 2 Sezione di Gastroenterologia,<br />

University of Palermo, Palermo, Italy; 3 Clinical<br />

and experimental Medecine, Università del Piemonte Orientale,<br />

Novara, Italy; 4 Service d’Anesthésie Réanimation, APHP, Hôpital<br />

Pitié Salpétrière, Paris, France; 5 Sevice de Chirurgie Hépatobiliaire<br />

et de Transplantation, APHP, Hôpital Pitié Salpétrière, Paris,<br />

France<br />

Background: Cirrhotic patients are spontaneously immunocompromised<br />

and are at increased risk of infection with higher<br />

morbidity and mortality. Vaccination may reduce the mortality<br />

related to infectious complications in those patients. Only<br />

few data regarding protection against vaccine preventable<br />

diseases and the results of vaccination are available in cirrhotic<br />

patients awaiting liver transplantation (LT). Aims: The aims of<br />

this prospective non-randomized study were to prospectively<br />

assess the prevalence of hepatitis A, B and varicella viruses<br />

(HAV,HBV,VZV) serological markers and to the evaluate the<br />

immunization status in a cohort of cirrhotic patient awaiting LT<br />

to determine the need of vaccination in this cohort. Patients and<br />

methods: All the cirrhotic patients registered on the LT waiting<br />

list in a single center were questioned about the prevalence of<br />

previous diseases and vaccinations by a questionnaire. The<br />

enzyme linked immunosorbent assay (ELISA) method was used<br />

to assess the presence of HBsAg, anti-HBc, anti-HBs, anti-HAV<br />

and anti-VZV antibodies. The positivity of serology was defined<br />

as: HBsAg ≥1 mUI/mL, anti-HBs ≥10 mUI/L, anti-HBc≥1 mUI/L,<br />

anti-HAV(IgG)≥1mUI/mL and anti-VZV(IgG)≥165 mUI/mL for<br />

HBV, HAV and VZV viruses, respectively. Results: 201 patients<br />

(male: 74%, mean age: 54±11years) have been evaluated.<br />

The indication for LT was HCV-related cirrhosis in 42.4%, HBV<br />

cirrhosis in 9%, alcoholic cirrhosis in 43.1%, NASH in 6.1%,<br />

hepatocellular carcinoma in 41.9%, and other indications in<br />

21.2%. Forty-five percent were class A, 33.2% B and 18.9%<br />

C according to Child-Pugh score; median MELD score was 12<br />

(6-33). Only 2.5% of patients had a vaccination book, 50%<br />

were anaware of having or having not a pasthistory of HBV,<br />

HAV or VZV infection. Twenty-nine percent, 6.7% and 4.0%,<br />

respectively, reported a vaccination against HBV, HAV and<br />

VZV. Seroprevalence was 11.1% for HBsAg, 32.2% for anti-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 823A<br />

HBc, 34% for anti-HBs, 19.8% for anti-HBc-/anti-HBs+, 79.4%<br />

for anti-HAV(IgG), and 91.7% for anti-VZV(IgG). Conclusion:<br />

Despite vaccination recommendations in cirrhotic patients<br />

awaiting LT, only few patients are informed and vaccinated<br />

while on the LT waiting list in a large Parisian center. Serological<br />

protection for VZV and HAV was high (91.7 and 79.4<br />

respectively), but it was low for HBV (34%), demonstrating the<br />

urgent need of vaccination management in this population.<br />

Disclosures:<br />

Pascal Lebray - Grant/Research Support: Merck, astellas, Biotest, BMS; Speaking<br />

and Teaching: Janssen, MSD, Gilead<br />

The following authors have nothing to disclose: Mazzola Alessandra, Margherita<br />

Tran Minh, Raluca Pais, Denis Bernard, Claire Goumard, Yvon Calmus, Filomena<br />

Conti<br />

1240<br />

Increase in Red Cell Distribution Width Post Liver Transplant<br />

is Associated with Patient Mortality<br />

Thure Caire 1 , Kirbylee K. Nelson 2 , R. Todd Stravitz 4 , Ambuj<br />

Kumar 2 , Nyingi M. Kemmer 3 ; 1 Gastroenterology, USF, Brandon,<br />

FL; 2 Internal Medicine, USF, Tampa, FL; 3 Hepatology, Tampa General<br />

Medical Group, Tampa, FL; 4 VCU, Richmond, VA<br />

Background: Following liver transplantation (LT), patients are at<br />

significant risk of mortality, however, prognostication following<br />

LT is limited. The red cell distribution width (RDW) is a measure<br />

of anisocytosis, and has been associated with morbidity and<br />

mortality in several chronic diseases associated with chronic<br />

systemic inflammation. The association of changes in RDW with<br />

mortality in patients following transplant surgery is unknown.<br />

As a potential surrogate for immunopathology, we hypothesize<br />

an increase in RDW may be positively correlated with mortality<br />

risk, therefore, the aim of this study is to determine the association<br />

between changes in RDW following LT and patient mortality.<br />

Methods: We performed a retrospective cohort study of all<br />

LT recipients performed at our center from Jan 1 2012 – Dec<br />

31 2012. The data collected included demographics (age,<br />

gender, ethnicity/race, indication for LT), pre-transplant RDW<br />

lab data (MELD score, hematologic parameters including MCV<br />

and RDW, Liver panel, renal profile) with a 24 month follow-up<br />

period including lab data at 6 months post-LT, and mortality 2<br />

years post-LT. We retrospectively analyzed data for associations<br />

with survival. Our primary end point was association of<br />

change in RDW (pre-LT to 6 months post-LT) with post-transplant<br />

mortality at 2 years. Statistical analysis was done using chisquare,<br />

fischer exact test, regression analysis. A p-value 30kg/m 2 (OR 1.089, p = 0.048),<br />

use of rapamycin (OR 9.898 p < 0.001), prior intra-arterial<br />

HCC therapy (OR 3.761, p < 0.001), take back surgery (OR<br />

9.038, p < 0.001), and cold ischemic time (OR 1, p = 0.209)<br />

were considered significant. On multivariate analysis, only biologic<br />

MELD at LT (OR 0.896, p = 0.028), BMI>30kg/m 2 (OR<br />

1.122, p = 0.043), take back within 1 st month (OR 14.559,<br />

p < 0.001), and rapamycin use (OR 12.450, p < 0.001)<br />

remained significant. Pre-LT intra-arterial therapy was not a<br />

significant predictor on multivariate analysis (OR 2.546, p =<br />

0.140). Conclusion: Pre-transplant intra-arterial HCC therapies<br />

to downstage into or maintain potential LT recipients within<br />

Milan criteria are not associated with increased prevalence of<br />

HAT post-LT. With respect to pre-LT HCC intra-arterial treatment<br />

factors, no increased risk for HAT is conferred with modality<br />

(chemo vs. radio-embolization), number of sessions, or number<br />

of HCC foci.<br />

Disclosures:<br />

Anil B. Seetharam - Advisory Committees or Review Panels: Bristol Meyers<br />

Squibb; Speaking and Teaching: Gilead, Janssen, Merck, Bayer<br />

The following authors have nothing to disclose: Mark Pedersen, Meera Ramanathan,<br />

Myunghan Choi


824A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1242<br />

The Association of Pre-Transplant Sarcopenia and Cirrhotic<br />

Cardiomyopathy with Postoperative Complications<br />

after Liver Transplant.<br />

Tripti Gupta 1,2 , Ryan M. Durel 1 , Qingyang Luo 1 , Daniel Devun 1 ,<br />

John Seal 1 , Emily Ahmed 1 , David S. Bruce 1 , Trevor W. Reichman 1 ,<br />

Shobha Joshi 1 , Natalie H. Bzowej 1 , Nigel Girgrah 1 , Humberto E.<br />

Bohorquez 1 , Ari J. Cohen 1 , Ian C. Carmody 1 , George E. Loss 1 ,<br />

George Therapondos 1 ; 1 Ochsner Clinic Foundation, New Orleans,<br />

LA; 2 University of Queensland, Brisbane, QLD, Australia<br />

Background: Sarcopenia in patients listed for liver transplantation<br />

(LT) is associated with increased death on the wait list<br />

and may affect post-operative recovery. These patients also<br />

have cirrhotic cardiomyopathy but its effect on post-operative<br />

outcomes remains unclear. Aims: To describe the prevalence of<br />

sarcopenia and cirrhotic cardiomyopathy in a cohort of adult<br />

patients with cirrhosis who are waitlisted for LT. To explore a<br />

correlation between these two conditions. To investigate their<br />

effect on post-operative outcomes. Methods: This is a retrospective<br />

study of 327 cirrhotic patients who underwent LT at<br />

Ochsner Medical Center between Jan 2012-Dec 2013. Pre-LT<br />

data such as demographics, clinical characteristics, etiology<br />

of liver disease, MELD score, 2D echocardiographic data (Left<br />

ventricular volume index- LAVI), E/A ratio, QTc interval were<br />

collected. CT abdomen/pelvis was used to calculate psoas<br />

muscle cross-sectional area (at L4 level). Post-LT data were also<br />

collected- no of days in hospital (first admission), no of ICU<br />

admissions, no of infections and creatinine level. Results: The<br />

M:F ratio was 68/32 and the mean age was 55 yrs. The mean<br />

MELD score was 23.5. Data for the first postoperative year<br />

were collected. The MELD scores and other demographics as<br />

well as 1-yr patient and graft survivals in the sarcopenia vs<br />

non-sarcopenia groups were similar. Cardiac assessment: 36%<br />

of our cohort had an abnormal LAVI. Sarcopenia: 19.5% of<br />

patients were sarcopenic. There was no correlation between<br />

the presence of sarcopenia and any of the abnormal cardiac<br />

measurements. 12.5% of patients with sarcopenia were discharged<br />

to a rehab facility vs only 6.8% of patients without<br />

sarcopenia. See Table for other post-LT outcomes. Summary:<br />

There was no correlation between the presence of sarcopenia<br />

and cirrhotic cardiomyopathy. Sarcopenia was shown to<br />

be associated with prolonged hospital stay, more episodes of<br />

infection and was more likely to lead to a discharge to a rehabilitation<br />

facility rather than home. The presence of cirrhotic<br />

cardiomyopathy pre-LT was associated with worse renal function<br />

at 1 year post LT. Conclusions: Sarcopenia is associated<br />

with increased post-LT morbidity and increased health care<br />

utilization while cirrhotic cardiomyopathy is associated with<br />

post-LT renal impairment. Although both conditions are seen<br />

in patients with cirrhosis, they do not appear to be associated<br />

with each other.<br />

Post-LT outocmes<br />

Disclosures:<br />

Shobha Joshi - Grant/Research Support: Salix, Eisai; Speaking and Teaching:<br />

Merck, Gilead, Bristol-Myers Squibb<br />

Natalie H. Bzowej - Grant/Research Support: Gilead Sciences, Ocera Therapeutics<br />

Nigel Girgrah - Speaking and Teaching: Merck, Bayer, Vertex, Gilead, Merck,<br />

Bayer, Vertex, Gilead, Merck, Bayer, Vertex, Gilead, Merck, Bayer, Vertex,<br />

Gilead<br />

The following authors have nothing to disclose: Tripti Gupta, Ryan M. Durel,<br />

Qingyang Luo, Daniel Devun, John Seal, Emily Ahmed, David S. Bruce, Trevor<br />

W. Reichman, Humberto E. Bohorquez, Ari J. Cohen, Ian C. Carmody, George<br />

E. Loss, George Therapondos<br />

1243<br />

Outcomes Post-Liver Transplant for Primary Sclerosing<br />

Cholangitis: Colitis Activity and Malignancy<br />

Astrid-Jane Greenup 1 , Gokulan Pavendranathan 1,2 , Mathew J.<br />

Keegan 1 , Sean Griffin 1 , David Bowen 1,4 , Warwick Selby 1,3 , Nicholas<br />

A. Shackel 1,4 , Simone I. Strasser 1,3 , Geoff McCaughan 1,4 ,<br />

David J. Koorey 1,3 ; 1 A W Morrow GI and Liver Centre, Royal<br />

Prince Alfred Hospital, Sydney, NSW, Australia; 2 University of<br />

New South Wales, Sydney, NSW, Australia; 3 University of Sydney,<br />

Sydney, NSW, Australia; 4 Centenary Institute of Cancer Medicine<br />

& Cell Biology, Sydney, NSW, Australia<br />

Background The reported risk of colorectal cancer(CRC) in<br />

patients transplanted for primary sclerosing cholangitis(PSC)<br />

is higher than in other transplant indications and non-transplanted<br />

patients with PSC.Yearly endoscopic surveillance is<br />

recommended.Furthermore, inflammatory bowel disease(IBD)<br />

activity varies despite immunosuppressive therapy. Aims The<br />

aims of this retrospective observational study were to examine<br />

IBD colitis activity and rates of colonic dysplasia and cancer<br />

in patients post-liver transplantation for PSC. Methods Data<br />

for patients undergoing liver transplantation for PSC at Royal<br />

Prince Alfred Hospital from January 1986 to December 2014<br />

was collected and analysed. Results One hundred and fourteen<br />

patients were transplanted for PSC during this 29-year period.<br />

Thirty-one patients were excluded for less than 2 years of follow<br />

up post-transplant (deceased(n=20;19 prior to 2008) or transplantation<br />

since January 2013(n=11)).Of the remaining 83<br />

patients,14 patients had a pre-transplant colectomy,care of 4<br />

patients was transferred to another interstate transplant centre<br />

and 3 had limited available medical records.Eighteen of 62<br />

patients did not have IBD pre-transplant,of whom 1 required<br />

a colectomy(for sporadic CRC) and 4 subsequently developed<br />

IBD.Accordingly,post-transplant IBD outcomes were available<br />

for 48 patients with mean age 58 years,male to female ratio<br />

2:1 and diagnosis of ulcerative colitis in 28 and Crohn’s<br />

disease in 20.Mean follow-up was 11 years.Forty-seven of<br />

48 had had at least one colonoscopy;median interval 1.74<br />

years.Colitis was histologically active in 68%(n=32) post-transplant,including<br />

22 patients with quiescent disease preceding<br />

transplant.Three patients received tumour necrosis factor<br />

antagonist therapy,inducing remission in two whilst the other<br />

required colectomy.A further four patients needed colectomy<br />

for refractory colitis.In the remaining 15,colitis was quiescent<br />

post-transplant.Six patients had a total or hemi colectomy for<br />

the indications of CRC(3),invisible dysplasia(1),endoscopically<br />

unresectable polypoid dysplasia(1) and caecal volvulus(1).An<br />

additional 4 patients had a history of invisible dysplasia and<br />

4 endoscopically resectable polypoid dysplasia. Conclusions<br />

Colitis has been active in 68% of patients post-transplant for<br />

PSC.Dysplasia and/or CRC have been detected in 14 patients.<br />

Colectomy has been necessary in 11 patients overall.These<br />

findings emphasise the need for aggressive surveillance and<br />

colitis management post-transplantation, including vigilance<br />

for de novo IBD development.Median colonoscopic interval<br />

of 1.74 years remains longer than recommended and is an<br />

opportunity for quality improvement.<br />

Disclosures:<br />

Simone I. Strasser - Advisory Committees or Review Panels: Janssen, AbbVie,<br />

Roche Products Australia, MSD, Bristol-Myers Squibb, Gilead, Norgine, Bayer<br />

Healthcare; Speaking and Teaching: Bayer Healthcare, Bristol-Myers Squibb,<br />

MSD, Roche Products Australia, Gilead, Janssen, Abbvie


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 825A<br />

Geoff McCaughan - Advisory Committees or Review Panels: Gilead<br />

The following authors have nothing to disclose: Astrid-Jane Greenup, Gokulan<br />

Pavendranathan, Mathew J. Keegan, Sean Griffin, David Bowen, Warwick<br />

Selby, Nicholas A. Shackel, David J. Koorey<br />

1244<br />

The Gap Between Assessed Physical Capacity And<br />

Actual Activity In Advanced Cirrhosis<br />

Michael A. Dunn 1,2 , Deborah A. Josbeno 3 , Amy R. Schmotzer 1 ,<br />

Doug Landsittel 1,2 , Anthony Delitto 3 ; 1 Medicine, University of Pittsburgh,<br />

Pittsburgh, PA; 2 Thomas E Starzl Transplantation Institute,<br />

University of Pittsburgh, Pittsburgh, PA; 3 Physical Therapy, University<br />

of Pittsburgh, Pittsburgh, PA<br />

Background: Muscle wasting in cirrhosis is a major cause of<br />

deterioration, transplant waitlist removals and deaths. It is<br />

associated with deconditioning and impaired physical performance,<br />

and becomes apparent as anatomic sarcopenia on<br />

imaging. Physical inactivity is a known driver of muscle wasting<br />

in chronic diseases. We instruct waitlisted patients to maintain<br />

regular activity and mandate the assessment of physical performance<br />

capacity for listing. However, very little is known about<br />

whether such assessments predict or reflect the extent of physical<br />

activity that patients actually perform to sustain health while<br />

awaiting transplantation. Methods: We asked 43 waitlisted<br />

patients to self-assess their ability to perform ordinary physical<br />

tasks using the Rosow-Breslau survey. We also recorded their<br />

Karnofsky physical capacity assessments from UNOS listing<br />

records. We then measured actual physical activity using a<br />

wearable armband with accelerometer and thermal sensing to<br />

record (a) percentage of waking time spent in sedentary (0-1.5<br />

METs), light (1.5-3.0 METS) and moderate-to-vigorous (>3.0<br />

METS) activity, (b) daily steps, and (c) daily energy expenditure<br />

in kcal. Results: We were able to acquire analyzable<br />

data for a 4 to 7 day wear period in 30 of the 43 patients<br />

(21 men, 9 women; age 57.5 ± 8.9; MELD 17.2 ± 6.2). Their<br />

measured levels of physical activity were among the lowest<br />

reported in chronic disease, comparable with that of COPD<br />

patients using oxygen or chronic renal failure patients on dialysis.<br />

Their percentages of waking hours spent in sedentary,<br />

light, and moderate-to-vigorous activity were 80.8 ± 14.7%,<br />

16.6 ± 13.0%, and 2.5 ± 2.7%, respectively. Compared with<br />

a range of 7,000-13,000 steps/day in healthy adults, their<br />

mean steps/day were 3132 ± 2022, range 552-8414. Both<br />

the activity percentage and step data were typical of other<br />

severely inactive populations. Mean daily energy expenditure<br />

was 2287 ± 468 kcal, as expected for their age range. Neither<br />

the self-assessed Rosow-Breslau scores (mean 2.5 ± 0.8,<br />

maximum 3.0) nor the provider-generated Karnofsky scores<br />

(mean 79.7 ± 11.3, maximum 100) indicated major impairment<br />

or showed a significant correlation with patients’ actual<br />

performance measured either by activity percentage or daily<br />

steps. Conclusion: Physical activity in cirrhotic patients is highly<br />

sedentary. Self and provider assessments of capacity do not<br />

correlate with actual performance. The extent to which a potential<br />

gap between assessed capacity and performance may be<br />

favorably modified by motivational, exercise, and nutritional<br />

interventions to improve activity and ameliorate muscle wasting<br />

is unclear, and merits further study.<br />

Disclosures:<br />

The following authors have nothing to disclose: Michael A. Dunn, Deborah A.<br />

Josbeno, Amy R. Schmotzer, Doug Landsittel, Anthony Delitto<br />

1245<br />

Liver Transplantation in Older Recipients<br />

Denise M. Harnois, Kaitlyn R. Musto, Bhupendra Rawal, Justin H.<br />

Nguyen; Mayo Clinic, Jacksonville, FL<br />

Introduction: The role of liver transplantation (LT) in elderly<br />

recipients (≥70 years) is controversial. Over the last decade,<br />

longevity has improved so that individuals who reach 70 years<br />

of age can expect an additional 15 years of life (http://cdc.<br />

gov/NCHS/products/nvsr.htm#vol58/). To address this issue,<br />

we review our experience of LT at a single large volume center<br />

to determine if LT is a reasonable option for older recipients to<br />

reclaim their expected lifespan if liver failure can be overcome.<br />

Methods: Retrospective analysis of 2524 LT recipients from<br />

1998 to 2013 was completed. Recipients of primary, whole,<br />

non-DCD, liver alone transplants, non-CCCA, and recipient<br />

age ≥18 were included. Clinically available variables including<br />

patient and graft survival along with donor and recipient<br />

characteristics were used. Results: Median recipient age was<br />

56 years (Range: 18-80) with a total of 1853 patients that<br />

included 129 older (≥70 years), 1482 middle-aged (46-69<br />

years), and 242 young (18-45 years) LT recipients. 1221 (66%)<br />

recipients were male. Median raw MELD was 16 (Range: 6-51)<br />

with median of 15 (Range: 6-43) for elderly. 40% had history<br />

of HCV and 22% with HCC. Post-LT median LOS was 7 days<br />

(Range: 0-446) and 8 days (Range: 0-105) for older recipients.<br />

933 (50.4%) of all recipients were sent to ICU post-LT, 65<br />

(50.4%) of older patients. Overall patient survival post-LT was<br />

95%. The 5-year patient survival was 85.1% (graft 73.5%) in<br />

young, 79.4% (graft 74.8%) in middle-aged, and 72.7% (graft<br />

69.0%) in older recipients. The 10-year survival was 72.1%<br />

(graft 59.5%) in young, 69.1% (graft 65.1%) in middle-aged,<br />

and 51.2% (graft 48.5%) in older recipients. Conclusion: LT<br />

offers the elderly a chance to reclaim their expected lifespan as<br />

they regain their liver function. Moreover, mortality, graft loss,<br />

LOS and requirement for ICU in older recipients are similar to<br />

those of young recipients. Therefore, LT in older recipient can<br />

be successful and should be considered in selected patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Denise M. Harnois, Kaitlyn R.<br />

Musto, Bhupendra Rawal, Justin H. Nguyen


826A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1246<br />

IgG4 status in explanted livers may affect the outcome<br />

of Primary sclerosing cholangitis (PSC) after liver transplantation<br />

Mansour G. Alghanem 1 , Bandar Al-Judaibi 2 , Paul Marotta 2 , Subrata<br />

Chakrabari 3 , Chaturika Herath 3 ; 1 mansour alghanem, university<br />

of western ontario, London, ON, Canada; 2 Hepatology,<br />

Western university, London, ON, Canada; 3 Western university,<br />

London, ON, Canada<br />

BACKGROUND: The outcome of Primary Sclerosing Cholangitis<br />

(PSC) after liver transplantation can be affected by recurrent<br />

PSC (rPSC) and subsequent graft failure. IgG4 related sclerosing<br />

disease is a recent entity that has a similar morphological<br />

appearance to PSC, making the distinction difficult. However,<br />

IgG4 related sclerosing cholangitis has an excellent prognosis<br />

since, it is steroid sensitive, but the impact of IgG4 on rPSC<br />

after liver transplant is still unknown. AIM: To determine the<br />

association between IgG4 immunochemical staining in liver<br />

explants and recurrence of primary sclerosing cholangitis postliver<br />

transpantation METHODS: All adult patients who underwent<br />

liver transplantation for PSC,from 1990 to 2014, were<br />

identified. Clinical inforamtion and immunochemical staining<br />

were obtained among the patietns with PSC recurrence postliver<br />

transplantation. IgG4 immunochemical staining was perforemed<br />

on the porta-hepatis region. Immunochemical staining<br />

is considered to be positive if the score was >5 cells/HPF.<br />

RESULTS: A total of eighty patients fulfilled the criteria for the<br />

study. IgG4 staining was positive in 21/80 (26%) subjects<br />

(>5cells/HPF). Median time for follow-up in IgG4 positive<br />

group was 99.6 months compared to 152.6 months in the<br />

IgG4 negative There were more episodes of rPSC in IgG4<br />

positive group (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 827A<br />

patients had ascites, 14 had encephalopathy, and 10 had<br />

varices. The accelerometer data showed that liver transplant<br />

candidates performed remarkably little daily physical activity<br />

(Table). There was no correlation between MELD and physical<br />

activity. Additionally, physical activity was not different<br />

based on the presence of ascites, hepatic encephalopathy,<br />

esophageal varices, or etiology of liver disease. Mean moderate-vigorous<br />

physical activity was correlated with right (r=0.51,<br />

p=0.017) and left (r=0.63, p=0.002) hand-grip strength.<br />

There was no correlation between physical activity and six-minute<br />

walk test distance. CONCLUSION: Liver transplant candidates<br />

with a relatively low mean MELD score were extremely<br />

sedentary, averaging only 5 minutes/day of physical activity.<br />

Moderate-vigorous activity was associated with greater grip<br />

strength. The lack of physical activity and large amount of time<br />

spent sedentary offer a potential for intervention to improve<br />

fitness in liver transplant candidates.<br />

76%, P=0.32, HR [95% CI]: 0.85 [0.60-1.2]). However, at<br />

MELD


828A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Comparing centers performing transplants for alcoholic hepatitis<br />

to centers not performing such transplants<br />

LT: Liver transplantation; AH: Alcoholic hepatitis; NRS: Non-response<br />

to steroids<br />

Disclosures:<br />

Brendan M. McGuire - Grant/Research Support: bayer healthcare, salix<br />

The following authors have nothing to disclose: Mohsen Hasanin, Derek Dubay,<br />

Thomas D. Schiano, Ashwani Singal<br />

1251<br />

15 first-days over-immunosuppression (IS) is NOT associated<br />

with immune-outcome measures post-liver transplantation<br />

(LT)<br />

Tommaso Di Maira, Victoria Aguilera, Angel Rubin, Carmen<br />

Vinaixa, Maria Garcia, Salvador Benlloch, Eva Montalva, Angel<br />

Moya, Fernando San Juan, Rafae Lopez-Andujar, Martin Prieto,<br />

Marina Berenguer; La Fe University Hospital, Valencia, Spain<br />

Body: Background & Aims: A recent study has described a<br />

strong association between early post liver transplantation (LT)<br />

(IS) and outcome. We aimed to confirm these findings in a<br />

cohort with long-term follow up. Methods: Tac and CsA trough<br />

levels obtained during the first 15 postLT days of patients<br />

transplanted between 2006-2008 were collected. High IS<br />

was defined by median Tac or CsA higher than 10 ng/ml or<br />

250 ng/ml respectively and/or the presence of a peak of Tac<br />

or CsA greater than 20 ng/ml and 400 ng/ml, respectively.<br />

Optimal IS was defined by median Tac or CsA levels between<br />

7-10 ng/ml or 150-250 ng/ml. Exclusion criteria were: lack of<br />

available data,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 829A<br />

strategies to improve post-LT survival of patients with cirrhosis<br />

with history of pre-transplant acute kidney injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Paul S. Fitzmorris, Kirk B. Russ,<br />

Donny Kakati, Ashwani Singal<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

The following authors have nothing to disclose: Maria Aguilar, Edward W. Holt,<br />

Taft Bhuket, Benny Liu, Robert J. Wong<br />

1253<br />

Renal Function Recovery and Outcomes after Liver<br />

Transplantation Alone among Patients with Pre-Transplant<br />

Acute Kidney Injury<br />

Paul S. Fitzmorris 1 , Kirk B. Russ 1 , Donny Kakati 1 , Ashwani Singal<br />

2 ; 1 Internal medicine, University of Alabama at Birmingham,<br />

Birmingham, AL; 2 Gastroenterology and Hepatology, University of<br />

Alabama at Birmingham, Birmingham, AL<br />

Purpose: Data on recovery of renal function recovery after liver<br />

transplantation (LT) alone are controversial. Methods: From an<br />

ongoing prospective study to define urine or serum biomarkers<br />

predictive of renal function recovery after LT, high risk patients<br />

with history of acute kidney injury (AKI) while waiting for LT<br />

were prospectively followed for renal function recovery at 6 mo.<br />

and for patient survival at 1 yr. after LT. Glomerular filtration<br />

rate was estimated using the modified diet in renal disease-6<br />

(MDRD-6) equation. Results: Of 109 patients with pre-LT AKI,<br />

52 successfully received LT. After excluding 3 with simultaneous<br />

kidney and 2 with lack of pre-transplant data, 47 (38 with<br />

1, 6 with 2, and 3 with 3 episodes of AKI) receiving LT alone<br />

were reviewed. Of these, 13 were excluded from analysis for<br />

lack of MDRD-6 data at 6 months after LT (n=4), not made up to<br />

6 months after LT (n=5), and death within 6 months of LT (n=4,<br />

1 each due to operative on table, sepsis, pulmonary embolism,<br />

and exploratory laparotomy for hemorrhage). Of remaining 34<br />

patients, 11 with recovery of renal function (improved MDRD-6<br />

by 50% compared to MDRD-6 at LT) differed from 23 who<br />

without recovery of renal function for MDRD-6 at LT (28±15<br />

vs. 60±30, p=0.002) and length of stay after LT (15±6 d vs.<br />

8±3 d, p=0.0002) without differences on age at LT, sex comorbidities,<br />

pre-existent chronic kidney disease, and number of<br />

AKI episodes prior to LT. Of 25 patients with MDRD-6


830A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1255<br />

Role of Coronary Artery Calcium (CAC) Score in Identifying<br />

Patients at Risk for Cardiac Events Within the 30<br />

Day Postoperative Period Following Deceased Donor<br />

Liver Transplantation<br />

Theodore W. James 1 , Joban Vaishnav 1 , Mohammad U. Malik 2 ,<br />

Aliaksei Pustavoitau 3 , Andrew M. Cameron 4 , Stuart D. Russell 1 ,<br />

Ahmet Gurakar 1 ; 1 Medicine, Johns Hopkins Hospital, Baltimore,<br />

MD; 2 Medicine, Conemaugh Memorial Medical Center,<br />

Jonestown, PA; 3 Anesthesiology and Critical Care Medicine, Johns<br />

Hopkins Hospital, Baltimore, MD; 4 Transplant Surgery, Johns Hopkins<br />

Hospital, Baltimore, MD<br />

Objective: To determine the relationship between CAC score<br />

and cardiac events encountered within the 30 day period<br />

following deceased donor liver transplantation (LT). Methods:<br />

Patients with a history of hypertension, Diabetes Mellitus, smoking<br />

or NASH underwent CAC scoring as part of their pretransplant<br />

evaluation. Following IRB approval, the records of<br />

adult patients who underwent CAC scoring and subsequent<br />

deceased donor LT between 1/1/2012 and 3/1/2015, with<br />

a minimum of 90 day follow-up, were retrospectively reviewed.<br />

The discharge summaries and progress notes were searched<br />

for cardiac events during the index admission or 30 day postoperative<br />

period. Cardiac events were categorized as new<br />

arrhythmia, acute coronary syndrome (ACS), new heart failure<br />

symptoms, new valvular disease or other cardiac disease. Postoperative<br />

cardiac events were compared at different Agatston<br />

unit cutoffs via Receiver Operating Characteristic (ROC) analysis.<br />

Results: This study included 75 subjects; (68% males, age<br />

range 41-69, median age 57, median biological MELD at the<br />

time of transplant 20; 45 Hepatitis C, 11 Alcohol, 6 NASH,<br />

4 Autoimmune, 1 Primary Sclerosing Cholangitis, 1 Primary<br />

Biliary Cirrhosis, 7 other). Twenty recipients experienced a cardiac<br />

event within the 30 day postoperative period (70% males,<br />

median age 58, median biological MELD 20; 11 Hepatitis<br />

C, 3 Alcohol, 2 NASH, 1 Autoimmune, 3 other). Six patients<br />

experienced a new arrhythmia, 13 experienced ACS, and<br />

one patient experienced myopericarditis. ROC analysis demonstrated<br />

a maximal sensitivity and specificity for 30 day postoperative<br />

cardiac events (85% and 58% respectively) at 80<br />

Agatston units. At this cut point, the negative predictive value<br />

was 91% (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 831A<br />

were seen and evaluated for LT, 138 (59%) underwent RHC<br />

based on RVSP on echocardiography of >35 mm Hg. Patients<br />

with and without RHC did not differ for demographics, comorbidities,<br />

liver disease etiology, complications of liver disease<br />

(ascites, hydrothorax, congestive heart failure, dialysis, hepatocellular<br />

carcinoma) and MELD score. Of 123 patients with<br />

available data on pulmonary artery pressure (PAP) measurements,<br />

51 (41%) had pulmonary hypertension (PHT) with mean<br />

PAP>25 mm Hg. Patients with mean PAP>25 compared to 25 and pulmonary vascular resistance >240<br />

dynes) was identified in only 5 (4.7%) patients. Conclusion:<br />

On screening echocardiography among patients with cirrhosis<br />

who are candidates for LT, RVSP cut-off of 40 mm Hg is most<br />

optimal for further evaluation for PHT with RHC. About 5% of<br />

these patients have true PHT.<br />

Disclosures:<br />

The following authors have nothing to disclose: Natalie Mitchell, Fakhar U. Islam,<br />

Nicholas Piazza, Danisa M. Clarett, Ashwani Singal<br />

1258<br />

Pre-transplant portal vein thrombosis is an independent<br />

risk factor for graft loss due to hepatic artery thrombosis<br />

in liver transplant recipients<br />

Jonathan G. Stine 1 , Shawn Pelletier 4 , Timothy Schmitt 2 , Robert J.<br />

Porte 3 , Patrick Northup 1 ; 1 Gastroenterology & Hepatology, University<br />

of Virginia, Charlottesville, VA; 2 Surgery, University of Kansas,<br />

Kansas City, KS; 3 Surgery, University of Groningen, Groningen,<br />

Netherlands; 4 Surgery, University of Virginia, Charlottesville, VA<br />

Purpose: To examine hepatic artery thrombosis risk in liver<br />

transplant recipients. We hypothesize that recipients with portal<br />

vein thrombosis are at increased risk. Methods: Data on all<br />

transplants in the United States during the MELD era through<br />

September 2014 were obtained from UNOS. Status one,<br />

multi-visceral, living donor, re-transplants, pediatric recipients<br />

and donation after cardiac death were excluded. Recipients<br />

were sorted into two groups: those with HAT and those without<br />

and univariate comparisons were performed. Incomplete HAT<br />

data was excluded. Multivariable logistic regression models<br />

were constructed for hepatic artery thrombosis with resultant<br />

graft loss within 90 days of transplantation. Findings: 6,134<br />

recipients underwent transplantation with known hepatic artery<br />

thrombosis status. 662 recipients had early hepatic artery<br />

thrombosis; of those, 91 (13.8%) had pre-transplant portal<br />

vein thrombosis, versus 6.8% with portal vein thrombosis but<br />

no hepatic artery thrombosis (p


832A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Multivariate Model: WL Removal Death/Too Sick<br />

Variables not significant: BMI, blood type, sex, dialysis, PVT,<br />

prior abdominal surgery<br />

Disclosures:<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

The following authors have nothing to disclose: Jeanne-Marie Giard<br />

1260<br />

Rotational Thromboelastometry (ROTEM) versus Conventional<br />

Coagulation Tests during Orthotopic Liver Transplantation:<br />

Comparison of Intra-operative Blood Loss,<br />

Transfusion Requirements, and Cost<br />

Laura Smart 1 , Danielle T. Scharpf 2 , Nicole O’Bleness Gray 1 ,<br />

Daniel Traetow 2 , Sylvester Black 3 , Anthony Michaels 1 , Elmahdi<br />

Elkhammas 3 , Robert B. Kirkpatrick 1 , Khalid Mumtaz 1 , A. James<br />

Hanje 1 ; 1 Gastroenterology,Hepatology, and Nutrition, The Ohio<br />

State University Medical Center, Columbus, OH; 2 Anesthesiology,<br />

The Ohio State University Medical Center, Columbus, OH; 3 Transplant<br />

Surgery, The Ohio State University Medical Center, Columbus,<br />

OH<br />

Purpose: Orthotopic liver transplantation (OLT) can be associated<br />

with significant bleeding requiring multiple blood product<br />

transfusions, especially in patients with severe liver dysfunction.<br />

Rotational thromboelastometry (ROTEM) is a point-of-care<br />

device that has been used successfully to monitor coagulation<br />

on whole blood samples during OLT. Whether it reduces<br />

blood loss and transfusions during OLT remains controversial.<br />

Methods: ROTEM or conventional coagulation tests (activated<br />

partial thromboplastin time (aPTT), prothrombin time (PT), international<br />

normalized ratio (INR), platelet count, and fibrinogen)<br />

were used to guide transfusion of platelets, cryoprecipitate,<br />

and fresh frozen plasma (FFP) during OLT. 68 consecutive<br />

patients were included in this non-randomized retrospective<br />

study. 34 historic controls had transfusions guided by conventional<br />

labs and 34 patients had transfusions guided by<br />

ROTEM during OLT over 3 years. Patient characteristics as well<br />

as pre- and post- transplant laboratory data were collected.<br />

Intra-operative blood loss, type and amount of blood products<br />

transfused, and cost were compared between the two groups.<br />

Results: The ROTEM group had significantly less intra-operative<br />

blood loss compared to the conventional group (2.0L vs 3.0L,<br />

p=0.04). The total amount of blood products transfused was<br />

less in the ROTEM group but did not reach statistical significance<br />

(14.5 units vs. 17 units, p=0.11). Patients in the ROTEM<br />

group received significantly less FFP (4 units vs. 6.5 units,<br />

p=0.02) but more cryoprecipitate (2 units vs. 1 units, p=0.04).<br />

Patient characteristics and pre-transplant laboratory results<br />

were not statistically different between groups. Post-transplant,<br />

both the INR and platelet count were significantly higher in the<br />

ROTEM group (2.0 vs. 1.7, p=0.01 and 98,000 vs. 63,000,<br />

p=0.002, respectively). The direct cost of blood products was<br />

also less in the ROTEM group compared to the conventional<br />

group ($103,786.09 vs $123,067.01). Conclusion: Implementation<br />

of a ROTEM-guided transfusion algorithm resulted<br />

in a reduction in intra-operative blood loss and a trend toward<br />

decreased total blood product usage with resultant decrease in<br />

cost during orthotopic liver transplantation.<br />

Disclosures:<br />

Anthony Michaels - Advisory Committees or Review Panels: Gilead, BMS, Janssen;<br />

Speaking and Teaching: Gilead<br />

A. James Hanje - Consulting: Salix pharmaceutical<br />

The following authors have nothing to disclose: Laura Smart, Danielle T. Scharpf,<br />

Nicole O’Bleness Gray, Daniel Traetow, Sylvester Black, Elmahdi Elkhammas,<br />

Robert B. Kirkpatrick, Khalid Mumtaz<br />

1261<br />

Systolic Dysfunction Post-Liver Transplantation is associated<br />

with Increased All-Cause Mortality<br />

Michael Cheung 2 , She-Yan Wong 1 , Arun Mathew 3 , Dina Halegoua-De<br />

Marzio 1 ; 1 Medicine, Division of Gastroenterology and<br />

Hepatology, Thomas Jefferson University, Philadelphia, PA; 2 Medicine,<br />

Division of Cardiology, Lankenau Medical Center, Philadelphia,<br />

PA; 3 Medicine, Thomas Jefferson University Hospital,<br />

Philadelphia, PA<br />

Introduction Left ventricular systolic dysfunction has been<br />

described after liver transplantation (LT), and is associated with<br />

a critically ill status, high pre-transplant Model of End-Stage<br />

Liver Disease (MELD) score and malnutrition. Previous <strong>studies</strong> of<br />

this condition have reported a high likelihood of ejection fraction<br />

recovery and survival at 1 year. The aim of this study was<br />

to evaluate all-cause mortality of systolic dysfunction post-LT at<br />

1 year and associated risk factors. Methods A retrospective<br />

chart review from a single liver transplant center was completed<br />

on all adult patients who received a liver transplant<br />

between 1/2002 and 1/2015. Patients with systolic dysfunction,<br />

defined as left ventricular ejection fraction (EF)


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 833A<br />

Disclosures:<br />

The following authors have nothing to disclose: Michael Cheung, She-Yan Wong,<br />

Arun Mathew, Dina Halegoua-De Marzio<br />

1262<br />

Downstaging or Bridging Therapy of Hepatocellular<br />

Cancer Does Not Increase Incidence of Biliary Complications<br />

after Liver Transplantation<br />

Meera Ramanathan 2 , Mark Pedersen 2 , Myunghan Choi 3 , Anil B.<br />

Seetharam 1 ; 1 Banner Transplant and Advanced Liver Disease Center,<br />

University of Arizona College of Medicine-Phoenix, Phoenix,<br />

AZ; 2 Internal Medicine, University of Arizona College of Medicine-Phoenix,<br />

Phoenix, AZ; 3 Arizona State University College of<br />

Nursing and Health Innovation, Phoenix, AZ<br />

Background: Donor and vascular risk factors predisposing to<br />

bile duct ischemia are causes of biliary complications after liver<br />

transplant (LT). Manipulation of the hepatic artery prior to LT as<br />

often occurs in the management of hepatocellular cancer (HCC)<br />

may represent a novel risk factor for non anastomotic biliary<br />

complications. Aim: Evaluate effect of pre-LT hepatic artery<br />

based locoregional therapy on incidence of non anastomotic<br />

biliary complications after LT. Methods: Retrospective cohort<br />

study of consecutive LT recipients between 2009 and 2013.<br />

Incidence of non anastomotic biliary complications (stricture<br />

or leak) tabulated by review of radiologic or endoscopic cholangiogram.<br />

Univariate regression evaluated donor, recipient,<br />

surgical, and locoregional treatment variables associated with<br />

increased risk of biliary complication: cytomegalovirus (CMV)<br />

donor status, organ source (brain or cardiac death), donor<br />

age, native model for end stage liver disease (MELD) and body<br />

mass index (BMI) at LT, etiology of cirrhosis (viral vs. non-viral),<br />

warm ischemic time, cold ischemic time, need for take back surgery<br />

to control bleed in the 1 st month post LT, number of lesions<br />

at the time of HCC diagnosis, modality of arterial locoregional<br />

treatment (chemo vs radio-embolization) and number of treatment<br />

sessions. Variables of interest (p


834A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1264<br />

Severity and Predictive Factors of Chronic Renal Dysfunction<br />

after Liver Transplantation<br />

Yi Huang 1,2 , Leon A. Adams 1,2 , Oscar Arauz 2 , Peter W. Angus 3 ,<br />

Marie Sinclair 3 , Graeme A. Macdonald 4 , Uthayanan Chelvaratnam<br />

4 , Alan J. Wigg 6 , Sze Pheh Yeap 6 , Nicholas A. Shackel 5 ,<br />

Linda I. Lin 5 , Spiro C. Raftopoulos 2 , Geoff McCaughan 5 , Gary P.<br />

Jeffrey 2,1 ; 1 The University of Western Australia, Perth, WA, Australia;<br />

2 Sir Charles Gairdner Hospital, Perth, WA, Australia; 3 The<br />

Austin Hospital, Melbourne, VIC, Australia; 4 Princess Alexandria<br />

Hospital, Brisbane, Brisbane, QLD, Australia; 5 Royal Prince Alfred<br />

Hospital, Sydney, NSW, Australia; 6 Flinders Medical Centre, Adelaide,<br />

SA, Australia<br />

Background and aims: Renal dysfunction is one of the common<br />

complications after orthotropic liver transplantation (OLT). The<br />

aim of this study was to evaluate the prevalence and the predictive<br />

factors of chronic renal dysfunction and the requirement<br />

of dialysis post-OLT. Methods: Patients who underwent OLT<br />

in Australia in 2003-2009 were included. Exclusion criteria<br />

included: multi-organ transplantation, second OLT and a follow<br />

up time less than 6 months after OLT. Estimated glomerular<br />

filtration rate (eGFR), creatinine and dialysis status were<br />

recorded at the end of the follow up. Chronic renal dysfunction<br />

was defined as an eGFR less than 30 ml/min per 1.73 m 2 or<br />

the requirement of dialysis. Results: 474 patients were included<br />

in the final analysis: 338 (71%) were male and 136 (29%)<br />

were female, mean age was 50, one (0.2%) had dialysis pre-<br />

OLT, 71 (15%) had diabetes, 41 (9%) had hypertension. The<br />

mean follow up was six years (range: 0.5 to 10.3). 56 patients<br />

died during follow up: 23 patients died from liver related<br />

causes, 16 from non-HCC malignancy, 6 from infection, 2 from<br />

cardiovascular disease and 9 from other causes. Nine (2%)<br />

patients required a dialysis. Among patients without dialysis,<br />

the mean increase in creatinine post-OLT compared to pre-OLT<br />

was 22.8 umol/L. 114 (24%) patients had a creatinine reduction,<br />

308 (65%) patients had no change in creatinine level, 36<br />

(8%) patients had 1-fold creatinine increase, 7 (1.5%) patients<br />

had 2-fold increase and 8 (1.7%) patients had 3-fold increase.<br />

339 patients had eGFR measured at the end of follow up with<br />

a mean of 66 ml/min per 1.73 m 2 . 15 (4%) patients had an<br />

eGFR < 30 ml/min per 1.73 m 2 , 63(19%) patients had an<br />

eGFR < 50 ml/min per 1.73 m 2 and 261 (77%) patients had<br />

an eGFR ≥ 50 ml/min per 1.73 m 2 . 24 (7%) patients had<br />

chronic renal dysfunction. Univariate analysis found that diabetes,<br />

pre-OLT creatinine, hypertension, cholesterol were significantly<br />

correlated with chronic renal dysfunction with odds<br />

ratio of 3.3, 1.01, 3.5 and 0.7 respectively. Diabetes, pre-OLT<br />

creatinine, hypertension and days in ICU were significantly correlated<br />

with the requirement of dialysis post-OLT with odds ratio<br />

of 4.7, 1.01, 5.5 and 1.06 respectively. Multivariate analysis<br />

found that diabetes and pre-OLT creatinine were independent<br />

predictors for chronic renal dysfunction. Diabetes, pre-OLT creatinine<br />

and days in ICU were independent predictors for the<br />

requirement of dialysis post-OLT. Conclusion: This study showed<br />

a lower than expected rate of chronic renal dysfunction and<br />

dialysis post-OLT in Australia. Diabetes and pre-OLT creatinine<br />

were independent predictors for chronic renal dysfunction.<br />

Disclosures:<br />

Leon A. Adams - Patent Held/Filed: Quest diagnostics<br />

Peter W. Angus - Advisory Committees or Review Panels: Gilead Sciences, BMS;<br />

Grant/Research Support: Gilead sciences<br />

The following authors have nothing to disclose: Yi Huang, Oscar Arauz, Marie<br />

Sinclair, Graeme A. Macdonald, Uthayanan Chelvaratnam, Alan J. Wigg,<br />

Sze Pheh Yeap, Nicholas A. Shackel, Linda I. Lin, Spiro C. Raftopoulos, Geoff<br />

McCaughan, Gary P. Jeffrey<br />

1265<br />

Ohio Solid Organ Transplantation Consortium (OSOTC)<br />

criteria for liver transplantation in patients with alcoholic<br />

liver disease<br />

Kaveh Hajifathalian 1 , Zubin Arora 2 , Annette Humberson 2 , David<br />

S. Barnes 2 , Arthur J. McCullough 2 , Nizar N. Zein 2 , Bijan Eghtesad<br />

2 , Ibrahim A. Hanouneh 2 ; 1 Internal Medicine, Cleveland Clinic<br />

Foundation, Cleveland, OH; 2 Cleveland Clinic Foundation, Cleveland,<br />

OH<br />

Background A minimum of six to 12 months of abstinence and<br />

three months of participation in an alcohol rehabilitation program<br />

are typically required before patients with alcoholic liver<br />

disease are eligible for transplantation. Some patients are too<br />

ill to participate in a rehab program. The state of Ohio Solid<br />

Organ Transplantation Consortium (OSOTC) provides mechanisms<br />

for medically urgent patients to be listed for transplant<br />

after only one to three months of abstinence. Methods From<br />

our center’s liver transplant program, we selected patients with<br />

alcoholic liver disease who were deemed eligible for transplant<br />

based on OSOTC exception criteria. These are patients with<br />

low to medium risk of recidivism, who are willing to sign a contract<br />

to commit to a recovery program. Survival was compared<br />

between these patients and matched patients with alcohol liver<br />

disease from Organ Procurement and Transplant Network<br />

(OPTN) data records. Results 19 transplant patients with alcoholic<br />

liver disease who were listed based on OSOTC exception<br />

criteria were included in the analysis. Patients had a mean (SD)<br />

MELD score of 26 (8.8) before transplant. Median follow-up<br />

after transplant was 3.3 years (IQR 2.4-4.8). These patients<br />

were randomly matched based on exact MELD score to 38<br />

transplant patients from OPTN data records with a median<br />

follow up of 3.9 years months (IQR 1.6-6.5). At one year,<br />

cumulative survival rates (±SE) were 90±7% and 92±4% in<br />

OSOTC exception criteria and general liver transplant populations,<br />

respectively. The cumulative survival rates were 81±10%<br />

and 80±7% at three years. There was no difference in overall<br />

survival between two groups (p for log rank test=0.907). Only<br />

one patient from OSOTC group resumed drinking by the last<br />

follow-up visit. Conclusions Compared to traditional criteria<br />

for assessment of risk of recidivism in patient with alcohol liver<br />

disease, a careful selection process with more flexibility to evaluate<br />

eligibility on a case to case basis can lead to similar<br />

survival rates after transplant.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kaveh Hajifathalian, Zubin<br />

Arora, Annette Humberson, David S. Barnes, Arthur J. McCullough, Nizar N.<br />

Zein, Bijan Eghtesad, Ibrahim A. Hanouneh<br />

1266<br />

Recipient Risk Factors for Acute Cellular Rejection After<br />

Orthotopic Liver Transplant in Patients with Average<br />

MELD More Than 25<br />

Mengyuan Liu 1 , Kelly Galen 1 , Grace Guzman 2 , Hoonbae Jeon 3 ,<br />

Costica Aloman 1 ; 1 Division of Gastroenterology/Liver Diseases<br />

(MC716), University of Illinois at Chicago, Chicago, IL; 2 Pathology,<br />

University of Illinois at Chicago, Chicago, IL; 3 Transplant<br />

Surgery, University of Illinois at Chicago, Chicago, IL<br />

Background & Aims: The use of MELD score for liver organ<br />

allocation has resulted in transplanting sicker patients. Risk<br />

factors of acute cellular rejections (ACR) in liver transplant were<br />

reviewed in other previous <strong>studies</strong> before MELD era and in<br />

patients with MELD score less than 20. It is unclear whether<br />

the timing, severity, and risk factors of ACR have changed as<br />

a result of progressive increasingly MELD score at transplantation.<br />

Our aim was to assess ACR characteristics in our center,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 835A<br />

where the average MELD score at transplantation is higher than<br />

previously published <strong>studies</strong>. Methods: This is a single center<br />

retrospective study designed to asses risk factors associated<br />

with ACR status post adult orthotopic liver transplant (OLT).<br />

Recipient demographics, preoperative clinical and laboratory<br />

data and cytomegalovirus (CMV) immune status were recorded<br />

for each transplant. Diagnosis of ACR was confirmed by biopsy<br />

using the Banff schema for grading liver allograft rejection. Univariate<br />

and multivariate regressions were preformed to look for<br />

variables that are significant predictors of ACR. Results: This<br />

study included 178 consecutive OLT patients transplanted in<br />

our center from January 2008 to June 2013 at the University<br />

of Illinois Hospital. Only 49.4% (88/178) of patients were<br />

Caucasians. In this study population, the average MELD at<br />

transplantation was 27.6, which is much higher than previously<br />

published <strong>studies</strong> that looked into risk factors of ACR.<br />

The average time from transplant to ACR diagnosis was 283.9<br />

days. Majority of ACR episodes were mild/moderate and only<br />

8.5% were graded as severe based on Banff classification.<br />

Of the twenty-two characteristics analyzed, creatinine was a<br />

significant predicator for ACR in both univariate and multivariate<br />

models. Patients with creatinine elevation greater than 2.7<br />

mg/dL have lower rates of ACR. Underlying liver etiology was<br />

also studied and patients with either primary biliary cirrhosis<br />

(PBC) or primary sclerosing cholangitis (PSC) have significantly<br />

higher rates of ACR. Conclusion: The study confirmed a change<br />

regarding the timing and severity of ACR in MELD era. Recipient<br />

characteristics such as elevated creatinine and PBC/PSC<br />

background may affect the risk of ACR and should be taken<br />

into consideration when managing immunosuppression.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mengyuan Liu, Kelly Galen,<br />

Grace Guzman, Hoonbae Jeon, Costica Aloman<br />

1267<br />

Pre-Transplant Sarcopenia Does Not Reverse After Successful<br />

Liver Transplantation<br />

Trushar Patel, Valery Vilchez, Roberto Gedaly, Rashmi T. Nair,<br />

Anna Christina Dela Cruz, Terrence Barrett; University of Kentucky,<br />

Lexington, KY<br />

Objective: To study the evolution of pre-transplant Sarcopenia<br />

after successful Liver Transplantation and compare survival of<br />

sarcopenic and non-sarcopenic patients Methods: Patients who<br />

underwent liver transplantation between 2008 and 2013 at<br />

our center were analyzed. We used CT scan at third lumbar<br />

vertebrae to analyze with the SliceOmetic 5.0 software (Tomovision,<br />

Canada) which enables specific tissue demarcation using<br />

previously reported Hounsfield unit (HU) thresholds. Cutoffs for<br />

sarcopenia were based on a CT-based sarcopenia study for<br />

patients with malignancies (Skeletal Muscle Index (SMI)


836A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

significant CMV infection seen mostly within 1 st month in seropositive<br />

population. Post-Tx CMV infection does not correlate<br />

with pre-Tx CMV immunity.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ekta Gupta, Viniyendra Pamecha,<br />

Nadeem Hasnain, Yogita Verma, Ajeet S. Bhadoria, Niteen Kumar, Shiv<br />

K. Sarin<br />

1269<br />

Good results of liver transplantation for decompensated<br />

cirrhosis with multi-organ failure in ICU: a retrospective<br />

study in two French centers<br />

Florent Artru 1 , Alexandre Louvet 1 , Isaac Ruiz 2 , Julien Labreuche 3 ,<br />

Eric Levesque 2 , Philippe Ichai 2 , Sebastien Dharancy 1 , Guillaume<br />

Lassailly 1 , Emmanuel Boleslawski 4 , Gilles Lebuffe 5 , Audrey Coilly 2 ,<br />

Eric Vibert 6 , Philippe Mathurin 1 , Didier Samuel 2 , Faouzi Saliba 2 ;<br />

1 Hepatology, CHRU Lille, Lille, France; 2 Hepatology, Hopital Paul<br />

Brousse, Villejuif, France; 3 Biostatistics Unit, CHRU Lille, Lille,<br />

France; 4 Surgery Transplant Unit, CHRU Lille, Lille, France; 5 Anaesthesiology<br />

Department, CHRU Lille, Lille, France; 6 Surgery Transplant<br />

Unit, Hopital Paul Brousse, Villejuif, France<br />

Resuscitation and transfer in intensive care unit (ICU) is often<br />

debated in patients with end-stage liver disease (ESLD), a condition<br />

associated with a very poor outcome. Few data are<br />

available on the results of liver transplantation (LT) in cirrhotic<br />

patients with multiorgan failure (MOF) and ESLD, admitted to<br />

the ICU. Aims: 1) to describe outcome of patients with ESLD<br />

and MOF transplanted while in ICU 2) to compare the outcome<br />

of these patients to those who were not hospitalized in<br />

ICU at time of LT and to those admitted to ICU but who did not<br />

benefit from LT. Methods: Patients were retrospectively selected<br />

from two French centers if they underwent LT for ESLD while in<br />

ICU with at least 2 non-hematological organ failures defined<br />

by SOFA score or if they had only liver failure but required<br />

mechanical ventilation at time of LT. We excluded combined<br />

liver-kidney transplantation. In order to evaluate the impact<br />

of MOF at time of LT, we performed a case-control study in<br />

which each patient with MOF was matched A) to 4 control<br />

patients transplanted outside ICU (matching on age and gender)<br />

B) to 1 (or 2 when possible) patients hospitalized in ICU<br />

with MOF, not transplanted (matching on age, gender, SOFA<br />

score). Results: From 2008 to 2014, 41 transplant recipients<br />

with a median age of 53.4 years (19.5% female) met the<br />

inclusion criteria. At time of LT, MELD score was 40 (95%CI:<br />

33-40), SAPS 2 score was 51 (49-63) and SOFA score was<br />

14 (13-15). Median duration of ICU stay before LT was 9<br />

days (5-12). Prior to LT, 39% had vasopressive therapy, 51%<br />

had mechanical ventilation and 51% had renal replacement<br />

therapy. Duration of stay after LT was 12 days in ICU and 43<br />

days in total. No PNF was observed. 6 patients underwent a<br />

surgical procedure during the first month. 32 patients (78%)<br />

developed bacterial infection post-LT, 12 (29%) CMV reactivation<br />

and 7(17%) fungal infection. The 1-year patient survival<br />

of the cohort was 85.04±5.6%. Based on the case-control<br />

study, this good survival at 1 year was not different from that<br />

of matched controls transplanted outside ICU (n=164): 85.04%<br />

vs. 87.04%, p=0.8. As expected, patients transplanted with<br />

MOF had a higher MELD score at time of LT as compared<br />

to patients transplanted outside ICU: 40 vs. 16 (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 837A<br />

MS was diagnosed at 3, 6 and 12 months post-LT respectively<br />

in 6/26 (23%), 7/23 (30%) e 7/16 (44%) patients. After 12<br />

months post-LT 11/16 patients (66%) had an HOMA2 test<br />

suggestive for IR, whereas at 6 months post-LT that finding<br />

was present in 10/23 (43%). No cardiovascular events were<br />

recorded in the study cohort. In conclusion, these data show<br />

that post-LT MS affects nearly half of LT recipients, starting early<br />

after LT. Lifestyle modifications, should be recommended to<br />

transplanted recipients, starting in the early post-LT period. This<br />

would facilitate prevention of body weight gain and the associated<br />

abnormalities, thus reducing incidence of post-LT MS and<br />

the related cardio-vascular events.<br />

Disclosures:<br />

The following authors have nothing to disclose: Veronica Pepe, Giacomo Germani,<br />

Alberto Ferrarese, Alberto Zanetto, Elena Nadal, Ilaria Bortoluzzi, Francesco<br />

P. Russo, Marco Senzolo, Umberto Cillo, Patrizia Burra<br />

Intraoperative Factors<br />

1271<br />

How intra-operative factors affect kidney recovery after<br />

simultaneous liver-kidney transplantation (SLK)<br />

Mario Spaggiari 1 , Mohamed Shaaban 2 , Galal El-Gazzaz 1 , Lisa<br />

Louwers 1 , Emmanouil Palaios 1 , Cristiano Quintini 1 , Federico N.<br />

Aucejo 1 , Koji Hashimoto 1 , Teresa Diago 1 , Masato Fujiki 1 , Bijan<br />

Eghtesad 1 , Charles M. Miller 1 , Mauricio Perilla 2 , Dympna Kelly 1 ;<br />

1 Cleveland Clinic, Cleveland, OH; 2 Department of Anesthesia,<br />

Cleveland Clinic Foundation, Cleveland, OH<br />

Introduction: The present study by analyzing our SLK population<br />

wants to identify the modifiable risk factors that affect<br />

kidney recovery after transplant with particular focus on the<br />

intraoperative events. Material and Methods: This a retrospective<br />

analysis performed between 2004 and 2014. The<br />

patient/graft survival of SLK were compared with the group of<br />

liver transplant alone (OLT) and kidney transplant alone (KT)<br />

in the same era. DGF was defined as the need of dialysis in<br />

the first week post-transplant. Results: 64 SLK were performed<br />

from 2004 to 2014.21.8% of patients (14/64) didn’t require<br />

dialysis at the time of the transplant. When compared to OLT<br />

(n=936) and KT (n=632) there was no significant differences<br />

in patient and graft survival. The incidence of DGF was higher<br />

in the SLK group (27%) compared to 20% in the KT (P= 0.3).<br />

Pre-operative factors associated with DGF were higher recipient<br />

BMI (p


838A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1273<br />

Morbid Obesity and Diabetes (DM) are Associated with<br />

Increased Risk of Death on the Liver Transplant (LT)<br />

Waiting List<br />

Ani A. Kardashian 1 , Jennifer L. Dodge 2 , John P. Roberts 2 , Danielle<br />

Brandman 1 ; 1 Medicine, University of California, San Francisco,<br />

San Francisco, CA; 2 Surgery, University of California, San Francisco,<br />

San Francisco, CA<br />

Purpose: Obesity is a growing problem in LT candidates, paralleling<br />

the US obesity epidemic and the increase in LT for<br />

nonalcoholic steatohepatitis (NASH). While post-LT survival<br />

appears to be similar in obese and non-obese patients, data<br />

is scarce regarding risk of waitlist dropout in morbidly obese<br />

patients. We aimed to determine the impact of obesity on<br />

waitlist outcomes and evaluate predictors of dropout in those<br />

with morbid obesity or NASH. Methods: We retrospectively<br />

evaluated LT candidates listed between 3/02-12/13. We<br />

performed a competing risk analysis to evaluate predictors of<br />

removal or death on the waitlist. Variables with p-value40. Compared to those<br />

with BMI 25-29.9, patients with BMI≥40 were more likely to<br />

be female (46% vs 28%), diabetic (25% vs 18%), have NASH<br />

(35% vs 13%), and have shorter median waitlist times (161 vs<br />

208 days); all p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 839A<br />

Disclosures:<br />

The following authors have nothing to disclose: Kazuhiro Takahashi, David<br />

A. Bruno, Marwan Kazimi, Atsushi Yoshida, Marwan S. Abouljoud, Gabriel<br />

Schnickel<br />

1275<br />

Post-Transplant Lymphoproliferative Disease is more<br />

common in primary sclerosing cholangitis patients after<br />

liver transplant<br />

Ahmed Abu Shanab 1,2 , Yasser Gayed 1 , Naeem Ullah 1 , Diarmaid<br />

D. Houlihan 1 , Aiden McCormick 1 ; 1 Liver Unit, St-Vincent University<br />

Hospital, Dublin, Ireland; 2 Internal Medicine, Menofiya University,<br />

Shebin Alkom, Egypt<br />

Background: Post transplant lymphoproliferative disorder (PTLD)<br />

is a well recognized complication of therapeutic immunosuppression<br />

in solid organs transplant recipients. It represents a<br />

broad spectrum of abnormalities ranging from a benign infectious<br />

mononucleosis like illness to malignant lymphoma and<br />

is associated with high mortality. Methods: A retrospective<br />

study was performed by collecting data of liver transplant (LT)<br />

patients in a single institution from December 1993 to December<br />

2014. Data was analyzed to identify PTLD patients and<br />

determine their demographic details, the indication for liver<br />

transplant, presenting symptoms, immunosuppression regimens<br />

and patient survival and PTLD outcome. Results: From a total of<br />

705 liver transplants recipients, 20 patients (2.8%) were diagnosed<br />

with PTLD. There were more males (n=17, 85%) than<br />

females (n=3, 15%), with median time from LT to diagnosis of<br />

PTLD was 83 months. The primary indication for LT of the PTLD<br />

cohort was PSC (n=8, 40%) of overall 67 PSC post LT patients.<br />

The rate of occurrence of PTLD in PSC was significantly higher<br />

than the overall LT patients with p value


840A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The mean body mass index (BMI) at transplant was 27±4.2kg/<br />

m 2 with mean albumin 2.96±0.8g/dL. The mean TPA was<br />

927.1±280mm 2 , which is much lower than the expected normal<br />

mean of 1211.9±242.3mm 2 . During the study time period,<br />

13 (54%) patients died at a mean of 1.2±2 years. Conclusion:<br />

Although LVSD is an uncommon complication post-LT, patients<br />

with high MELD and malnutrition maybe at highest risk. Sarcopenia,<br />

measured by TPA, maybe a better marker for malnutrition<br />

than traditional markers. BMI can be falsely elevated by<br />

volume overload, which can explain the normal BMI seen in<br />

our patient group. Sarcopenia should be studied further as a<br />

possible predictor for LVSD. Pre-operative nutritional optimization<br />

and close post-LT echocardiogram monitoring is advisable<br />

in this population at risk for increased mortality post-LT.<br />

Disclosures:<br />

The following authors have nothing to disclose: Arun Mathew, She-Yan Wong,<br />

Michael Cheung, Flavius Guglielmo, Cataldo Doria, David A. Sass, Dina Halegoua-De<br />

Marzio<br />

1277<br />

Risk Stratification for Optimal Timing of Liver Transplantation<br />

Following Repeated Locoregional Therapies in<br />

Patients With Hepatitis B-Related Hepatocellular Carcinoma<br />

Hwi Young Kim 1 , Won Kim 1 , Yong Jin Jung 1 , Hyeyoung Kim 2 ,<br />

Nam-Joon Yi 2 , Kwang-Woong Lee 2 , Hae Won Lee 3 , Kyung-Suk<br />

Suh 2 ; 1 Department of Internal Medicine, Seoul National University<br />

Boramae Medical Center, Seoul, Korea (the Republic of); 2 Department<br />

of Surgery, Seoul National University Hospital, Seoul, Korea<br />

(the Republic of); 3 Department of Surgery, Seoul National University<br />

Boramae Medical Center, Seoul, Korea (the Republic of)<br />

BACKGROUND: Decision of timing for liver transplantation (LT)<br />

in patients with HCC often depends on the responses to pre-LT<br />

locoregional therapies especially in cases of organ shortage.<br />

However, prediction of favorable tumor biology before LT<br />

is difficult, and supporting data for decision of LT timing is<br />

scarce. The aims were to prognosticate HCC patients who<br />

underwent LT according to their responses to pre-LT locoregional<br />

therapies, and to shed light on the decision of optimal<br />

timing for LT. METHODS: Between January 2005 and<br />

December 2011, a total of 264 patients with hepatitis B-related<br />

HCC underwent LT. Excluding 52 patients with their<br />

tumors beyond Milan criteria, 212 patients were included.<br />

Responses to pre-LT locoregional therapies were assessed using<br />

the modified RECIST criteria. Pre-LT risk scores were generated<br />

as follows: [(No. of recurrence following radiofrequency<br />

ablation or ethanol injection + 2x(No. of progressive disease<br />

following transarterial chemoembolization(TACE)) + (No. of<br />

recurrence following previous response to TACE)]. Cox proportional<br />

hazard model was used to investigate prognostic<br />

factors for recurrence-free survival. RESULTS: Median age at<br />

the time of HCC diagnosis was 51 years, and male patients<br />

were 81.1%. Median time to LT was 6.5 months(interquartile<br />

range(IQR), 1.3-27.6). Median follow-up duration after LT<br />

was 55.3 months(IQR, 39.6-81.4). Median maximal diameter<br />

of the largest tumor was 2.0cm(range, 1.1-5). Hepatitis<br />

B viral DNA was 1875 IU/mL(median) at baseline. Baseline<br />

serum alpha-fetoprotein was 19 ng/mL(range, 1.8-23200).<br />

Baseline Child-Pugh classes were A in 101(47.6%) and B in<br />

61(28.8%), respectively. Before LT, 68 patients received at<br />

least one session of radiofrequency ablation or ethanol injection,<br />

and 100 patients received at least one session of TACE,<br />

respectively. Overall, 127 patients(59.9%) received pre-LT<br />

locoregional therapies. Pre-LT risk scores were as follows: 0,<br />

n=126(59.4%); 1-2, n=36(17.0%); 3-5, n=30(14.2%); ≥6,<br />

n=20(9.4%). From multivariable Cox analysis, pre-LT risk<br />

scores and maximal tumor diameter were independently significant<br />

prognostic factors for recurrence-free survival: risk score<br />

0vs.1-2, hazard ratio(HR)=1.422(95% confidence interval(CI),<br />

0.523-3.867), P=0.490); 0vs.3-5, HR=2.424(95%CI, 0.922-<br />

6.369; P=0.072); 0vs.≥6, HR=4.805(95%CI, 1.782-12.957;<br />

P=0.002); maximal diameter, HR=1.514(95%CI, 1.097-<br />

2.090; P=0.012). CONCLUSION: Increasing pre-LT risk score<br />

may represent a warning signal toward prompt LT rather than<br />

repeated locoregional therapies, especially in case of score≥6.<br />

Pre-LT risk score could be used as a surrogate marker of tumor<br />

biology for the decision of LT timing.<br />

Disclosures:<br />

Kwang-Woong Lee - Grant/Research Support: ChongGeunDang, Astellas,<br />

GreenCross<br />

The following authors have nothing to disclose: Hwi Young Kim, Won Kim, Yong<br />

Jin Jung, Hyeyoung Kim, Nam-Joon Yi, Hae Won Lee, Kyung-Suk Suh<br />

1278<br />

Synergistic anticancer effect of metformin in combination<br />

with immunosuppressant on hepatocellular carcinoma<br />

cell lines<br />

Suk-Won Suh 1 , Kwang-Woong Lee 2 , Yoo-Shin Choi 1 ; 1 Department<br />

of Surgery, Chung-Ang University College of Medicine, Seoul,<br />

Korea (the Republic of); 2 Seoul National University College of<br />

Medicine, Seoul, Korea (the Republic of)<br />

After liver transplantation (LT), immunosuppression is needed<br />

to avoid rejection and graft loss, however, it can stimulate<br />

hepatocellular carcinoma (HCC) recurrence and progression.<br />

Previous <strong>studies</strong> have shown that metformin had an anticancer<br />

effect on several cancers, including HCC. The aim of this<br />

study was to evaluate the interactions between metformin and<br />

immunosuppressive agents including sirolimus, tacrolimus and<br />

mycophenolate mofetil (MMF) for antitumor activity. Three cell<br />

lines (Huh7, HepG2 and Hep3b) were tested. Cell viability<br />

was determined using a MTT assay and Western blot analysis<br />

for mammalian target of rapamycin (mTOR) pathway related<br />

proteins were performed to reveal their mechanism. Metformin<br />

and sirolimus had synergistic antiproliferative effect and sirolimus<br />

plus metformin supplemented with MMF also showed<br />

synergistic antiproliferative effect in specific HCC cells. Synergistic<br />

effect of metformin and sirolimus through the inhibition<br />

of mTOR and its down-stream, p70S6K, and p-4EBP1 were<br />

demonstrated. Metformin and sirolimus also showed synergistic<br />

effect for the down-regulation of Livin and Survivin expressions<br />

in HepG2 and Hep3b cells. In conclusion, metformin had synergistic<br />

interactions with sirolimus in terms of anticancer effects<br />

for HCC cells and the mechanism explaining this synergistic<br />

inhibition might be related with mTOR pathway. These results<br />

may provide a foundation for further <strong>studies</strong> for patients with<br />

HCC who underwent LT in clinical era.<br />

Disclosures:<br />

Kwang-Woong Lee - Grant/Research Support: ChongGeunDang, Astellas,<br />

GreenCross<br />

The following authors have nothing to disclose: Suk-Won Suh, Yoo-Shin Choi


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 841A<br />

1279<br />

Evaluation of risk indices to determine 1 year graft survival<br />

after liver transplantation using the UNOS database<br />

Shahzad Ahmed, Vinay Sundaram; Cedars-Sinai Medical Center,<br />

Los Angeles, CA<br />

Aim: Given the shortage of viable organs available for liver<br />

transplantation (LT), relative to the supply, it is critical to ensure<br />

optimal post-transplant graft survival in recipients of LT. The<br />

Model for End-Stage Liver Disease (MELD) scoring systems was<br />

developed to determine 3-month survival prior to LT. Additional<br />

scoring systems have been developed to determine post-transplant<br />

outcomes including Donor Risk Index (DRI), Organ Patient<br />

Index (OPI) and Balance of Risk (BAR) score. The aim of this<br />

study was to evaluate the accuracy of these risk indices in evaluating<br />

post-transplant graft survival at 1 year. Methods: We<br />

analyzed the UNOS database from 2006-2009. We excluded<br />

patients under age 18 or who underwent multi-organ transplantation<br />

Results. A total of 24,388 patients were studied. Among<br />

these patients, 35% had hepatitis C virus, 15% had alcoholic<br />

liver disease and 12% had cholestatic liver disease (primary<br />

biliary cirrhosis and primary sclerosing cholangitis). Regarding<br />

1 year graft surivival, our findings indicated that all indices performed<br />

with modest accuracy as follows: MELD (>30/=16/


842A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

with adequate control of dyslipidemia in both groups. Blood<br />

glucose was higher in CNI (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 843A<br />

apeutic potential for the management of liver IRI in transplant<br />

recipients.<br />

Disclosures:<br />

Ronald W. Busuttil - Grant/Research Support: Bayer, Fujisawa, Novartis, Roche/<br />

Genentech, NIH<br />

The following authors have nothing to disclose: Shi Yue, Changyong Li, Xingliang<br />

Zhou, Qilong Ying, Jerzy Kupiec-Weglinski, Bibo Ke<br />

1284<br />

The long noncoding RNA VL30-1 is a key regulator of<br />

liver injury via enhancement of HIF1a mediated inflammation<br />

Xinshou Ouyang, Sheng-Na Han, Irma Garcia-Martinez, Luke<br />

Cai, Richard A. Flavell, Wajahat Z. Mehal; Yale University, New<br />

Haven, CT<br />

Introduction: Inflammation is the key feature of liver injury in<br />

response to many insults. The inflammatory gene expression<br />

program is comprised of several specific transcription factors<br />

involving long noncoding RNAs (lncRNAs). HIF1a is well<br />

known regulator of inflammatory gene transcription, and is<br />

required for persistent liver inflammation. LncRNAs play important<br />

roles in diverse biological processes. However, the role of<br />

lncRNAs in regulating hepatic inflammation is not known. Aim:<br />

To identify lncRNA (s) that regulates liver inflammation and<br />

injury. Methods: Primary mouse macrophages were treated<br />

with bacterial lipopolysaccharide (LPS), and whole-transcriptome<br />

analysis (RNA-seq) was conducted. Numerous lncRNAs<br />

were differentially expressed including the short form of the<br />

endogenous retrotransposon VL30-1. Further analysis determined<br />

the functional role and molecular mechanisms of VL30-1<br />

in regulation of immune genes. This was done by siRNA knockdown,,<br />

transcriptome microarray, promoter luciferase assay,<br />

ELISA, RT-PCR, confocal imaging and RNA immunoprecipitation<br />

PCR (RIP)-PCR. The in vivo role of VL30-1 in liver injury<br />

was demonstrated by knockdown in the LPS/D galactosamine<br />

(LPS/D GlN) model. Results: VL30-1 is strongly induced by LPS<br />

in mouse macrophages (fold > 90 vs control). Stable depletion<br />

of VL30-1 in Raw264.7 cells by shRNA significantly down-regulated<br />

LPS induced Il1b mRNA and protein levels. IL-1b secretion<br />

was significantly reduced by VL30-1 depletion in mouse<br />

peritoneal macrophages in response to typical inflammasome<br />

activation (808+/-45 pg/ml in control vs 214+/-125 pg/ml<br />

in VL30-1 siRNA). VL30-1 depletion also markedly reduced<br />

mIL-1b luciferase activity under forced expression of HIF1a<br />

and HIF2a. Conversely, forced VL30-1 expression dose-dependently<br />

trans-activated mIL-1b reporter luciferase activity,<br />

confirming that VL30-1 augments HIF1a mediated IL-1b<br />

transcription. RIP-PCR assay showed a direct VL30-1 binding<br />

with HIF1a. VL30-1 co-localized with HIF1a by confocal<br />

imaging analysis. Depletion of VL30-1 in mice significantly<br />

decreased LPS/ D-GalN induced serum IL-1b protein concentration<br />

(188+/-76 pg/ml in control vs 23+/- 26 pg/ml) as<br />

well as reduced liver damage as judged by morphology and<br />

serum ALT level (418+/-297 in control vs 151+/- 76 in VL30-1<br />

siRNA). Conclusions: These <strong>studies</strong> reveal a critical role of the<br />

lncRNA VL30-1 as a broad-acting regulatory component in<br />

the control of HIF1a induced inflammatory pathway activation<br />

in mouse macrophages and liver injury. This occurs via direct<br />

interaction of VL30-1 and HIF-1a, and up-regulation of HIF-1a<br />

mediated inflammatory transcripts.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xinshou Ouyang, Sheng-Na Han,<br />

Irma Garcia-Martinez, Luke Cai, Richard A. Flavell, Wajahat Z. Mehal<br />

1285<br />

Galectin-3 mediates NLRP3 inflammasome signaling<br />

and contributes to pathogenesis of primary biliary cirrhosis<br />

Jijing Tian 1,2 , Guo-Xiang Yang 1 , Fu-Tong Liu 1 , M. Eric Gershwin<br />

1 , Natalie J. Torok 1 , Joy Jiang 1 ; 1 UCDMC, Sacramento, CA;<br />

2 Research Center for Eco-environmental Sciences, Chinese Academy<br />

of Sciences, Beijing, China, Beijing, China<br />

Macrophages play a significant role in early inflammatory<br />

changes in primary biliary cirrhosis (PBC). However, their<br />

role in modulating fibrogenic responses has not been evaluated.<br />

We and others have shown that macrophages are<br />

a major source of galectin-3, a profibrogenic lectin. As the<br />

pathogenesis of PBC is not well understood, we hypothesized<br />

that galectin-3 is required to activate inflammasome signaling<br />

contributing to inflammation, fibrosis and progression in PBC.<br />

Methods: Liver tissues from PBC patients and healthy controls;<br />

and from the dnTGFβRII transgenic mouse model of PBC and<br />

wt controls were collected for RT-PCR and Western blot to analyze<br />

the expression of galectin-3, NLRP3, ASC and IL-1β. The<br />

cleavage of caspase-1 and IL-1β in PBC patients was examined<br />

by Western blots. Primary hepatic macrophages were<br />

isolated from wt and the galectin3 -/- mice, and treated with<br />

Deoxycholic Acid (DCA) with/without recombinant galectin-3.<br />

RT-PCR was done to analyze the activation of inflammasome-related<br />

transcripts; immunofluorescence (IF) to detect<br />

the activation of inflammasomes; and immunoprecipitation (IP)<br />

to detect the association of galectin-3 and NLRP3. DnTGFβRII/<br />

gal3 -/- mice were generated by crossing the dnTGFβRII mice<br />

with galectin3 -/- mice and the liver tissues were analyzed for<br />

inflammasome activation and fibrosis. Results: The transcripts<br />

of galectin-3, NLRP3, ASC and IL-1β significantly increased in<br />

the livers of PBC patients (N=4, p


844A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1286<br />

Kupffer cells signature in alcoholic steatohepatitis in<br />

mice encompasses loss of signalling/ chaperone and<br />

enrichment of cytoskeleton proteins in the lipid rafts<br />

Angela Dolganiuc 1 , Tracie Lo 2 ; 1 Medicine/GI, University of Florida,<br />

Gainesville, FL; 2 New York Medical College, New York, NY<br />

The mechanisms of alcoholic liver disease (ALD) are complex<br />

and largely related to dysfunctional Kupffer cells (KCs). Here<br />

we set to define the role of lipid rafts (LR) in alcohol exposed<br />

KCs. Mice were fed alcohol-containing or pair-fed Lieber-De-<br />

Carli Diet for 4 weeks. Plasma and liver tissue were analyzed<br />

for Tg and ALT by enzymatic methods; RNA by PCR. KCs were<br />

isolated by tissue digestion, gradient centrifugation and adherence<br />

to plastic. KCs and their model cell line RAW 264.7<br />

were further stimulated in vitro with LPS; cell migration was<br />

assayed with RTCA DP method; detergent-resistant membranes<br />

fractions, representative of lipid rafts (LR), were separated by<br />

gradient centrifugation and analyzed with LC-MS/MS and<br />

MALDI-TOF mass spectrometry. Alcohol consumption, unlike<br />

pair-fed diet, was associated with elevated serum ALT and liver<br />

Tg, suggestive of ALD steatohepatitis. KCs of alcohol-fed mice<br />

produced higher TNFa, suggestive of imbalanced signaling,<br />

and migrated slower on plastic, suggestive rigidity of cytoskeleton,<br />

compared to alcohol naive counterparts. The analysis<br />

of Log2 relative expression of proteins in the lipid raft protein<br />

profile resulted from mass spectrometry revealed that 93% of<br />

all detected proteins were similar in resting alcohol naïve and<br />

alcohol-treated RAW 264.7 cells. Among proteins enriched<br />

in the LRs of alcohol-treated cells compared to alcohol naïve<br />

counterparts were cytoskeleton-related proteins which are<br />

known to anchor to the cytoplasmic side of the LRs, including<br />

tubulin, beta actin, moesin, talins and synaptic protein VAT-<br />

1, alongside with those exposed on the extracellular side of<br />

the LRs including Fermitin family homolog pleckstrin homology<br />

domain-containing family C member 1. In contrast, we<br />

observed a significant loss of a cohort of signaling proteins in<br />

LPS-activated alcohol-treated RAW 264.7 cells compared to<br />

alcohol naïve counterparts; among most prominently depleted<br />

were iron-sulfur binding proteins, PARK chaperones, nontransforming<br />

monomeric GTP-binding proteins and sorting nexins.<br />

Only few signaling proteins were upregulated in alcohol<br />

exposed LPS-activated cells however these did not reach statistical<br />

significance. The RNA levels and the total cellular content of<br />

all proteins above were comparable in KCs of alcohol exposed<br />

mice and their pair-fed control counterparts. In conclusion, our<br />

novel results suggest impaired recruitment of signaling proteins<br />

and chaperones likely leading to functional impairment of lipid<br />

rafts in activated alcohol-exposed KCs, possibly secondary to<br />

increased LR rigidity secondary to enrichment of cytoskeleton<br />

proteins. These results point to novel therapeutic targets in ALD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Angela Dolganiuc, Tracie Lo<br />

1287<br />

IL28B genetic variants determine the extent of monocyte-induced<br />

activation of NK cells in hepatitis C<br />

Benjamin Krämer 1 , Beatriz Sastre Turrión 2 , Claudia Finnemann 1 ,<br />

Philipp Lutz 1 , Andreas Glässner 1 , Franziska Wolter 1 , Pavlos<br />

Kokordelis 1 , Dominik J. Kaczmarek 1 , Felix Goeser 1 , Christian P.<br />

Strassburg 1 , Ulrich Spengler 1 , Jacob Nattermann 1 ; 1 Department of<br />

Internal medicine I, Univeritiy of Bonn, Bonn, Germany; 2 Infectious<br />

Diseases Department, Hospital Ramón y Cajal - IRYCIS, Madrid,<br />

Spain<br />

Background: Genetic variants in IL28B (IFNλ3) in combination<br />

with the natural killer (NK) cell-associated killer cell Ig-like<br />

receptors (KIR) gene KIR2DS3 synergistically increased the<br />

risk of chronic infection over either factor alone. A proposed<br />

direct functional link between NK cells and λ-IFNs could not<br />

be confirmed since in line with the known limited expression of<br />

IFN-λR1, the natural λ-IFN receptor, on epthelial cells. NK cells<br />

do not express IFN-λR1 and do not respond to λ-IFNs. Thus, the<br />

synergism between IL28B genotype and NK cell activity must<br />

be mediated by indirect mechanisms. Here, we studied differential<br />

interactions between monocytes and NK cells which can<br />

contribute to the observed synergism between polymorphisms<br />

in the KIR and IFNλ genes. Methods: A total of 74 HCV(+)<br />

patients and 80 healthy controls were enrolled into this study.<br />

IL28B (rs 12979860) genotypes were determined by rtPCR.<br />

Peripheral NK cells and monocytes, respectively, were isolated<br />

using MACS technology. Total PBMC, isolated monocytes, and<br />

NK cells were stimulated with IL28B, the TLR7/8 agonist R848,<br />

or a combination of both. NK cell IFN-g response was analysed<br />

by FACS. IL12 and IL-18 secretion of monocytes was studied by<br />

ELISA. In blocking experiments anti-IL12 and/or anti-IL18 were<br />

used. Results: Following stimulation of total PBMCs with R848<br />

we found CD56Bright NK cell IFN-g responses to be associated<br />

with the IL28B genotype, with carriers of a T/T genotype<br />

displaying the lowest frequency of IFN-g(+) CD56 Bright NK<br />

cells (C/C vs. T/T, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 845A<br />

1288<br />

High baseline expression of type 1 interferon (IFN)-inducible<br />

genes involved in cell-autonomous immunity<br />

predicts mortality in patients with decompensated alcoholic<br />

cirrhosis<br />

Marc Pineton de Chambrun 1 , Emmanuel Weiss 1 , Pierre-Emmanuel<br />

Rautou 2 , Magali Fasseu 1 , Mikhael Giabicani 1 , Rakhi Maiwall<br />

3 , Dominique Valla 2,1 , Didier Lebrec 2,1 , Francois Durand 2,1 ,<br />

Pierre de la Grange 4 , Sophie Lotersztajn 1,2 , Richard Moreau 1,2 ;<br />

1 UMR S1149, Center for Research on Inflammation (CRI), Inserm<br />

and Paris Diderot University, Clichy, France; 2 DHU Unity, Service<br />

d’Hépatologie, Hôpital Beaujon, Clichy, France; 3 Hepatology<br />

Department, Institute of Liver and Biliary Science, New Delhi,<br />

India; 4 Genosplice, Paris, France<br />

Background & Aims: Although systemic inflammation is<br />

believed to be a major driver of mortality in patients with<br />

decompensated alcoholic cirrhosis, its mechanisms are unclear.<br />

We hypothesized that baseline expression levels of genes<br />

involved in cell-autonomous immunity (most of which are type<br />

1 IFN-inducible) and/or of genes encoding secreted inflammatory<br />

cytokines or chemokines by circulating mononuclear cells<br />

in patients with cirrhosis may be related to patient outcome.<br />

Thus, we measured the baseline gene expression in peripheral<br />

blood mononuclear cells (PBMCs) from patients with decompensated<br />

alcoholic cirrhosis and investigated their relationship<br />

with the risk of death. Methods: PBMCs were isolated in 98<br />

non infected patients (MELD score: 19 (15-24)) and 50 healthy<br />

subjects. Using RT-qPCR, we monitored 52 type 1 IFN-inducible<br />

genes and 10 IFN-non-inducible genes encoding ‘traditional’<br />

secreted inflammatory cytokines (e.g., IL6) or chemokines (e.g.,<br />

CXCL1). We also used a prespecified IFN score (a composite<br />

of type 1 IFN-inducible genes: OAS2, MX2, DDX58, GBP4,<br />

IFIH1, TRIM22, IFIT1, CXCL10). Gene expression in ‘cirrhotic<br />

cells’ was normalized using expression in ‘healthy’ cells.<br />

Patients were then followed-up for 10±12 months, until death,<br />

transplantation or the end of the study. Results: In univariate<br />

analysis, a higher MELD score, higher baseline expression of<br />

9 IFN-inducible genes as well as higher IFN score values were<br />

significant predictors of death. In contrast, cytokine or chemokine<br />

gene expressions were not significantly related to death. In<br />

bivariate analysis including the IFN and MELD scores, only the<br />

former significantly predicted death (RR=3.58; 95% CI, 1.18-<br />

10.91; P=0.02). These results were mainly due to the elevated<br />

intrinsic prognostic value of OAS2 (RR=2.49; P=0.04) and<br />

MX2 (RR=1.35; P=0.01). It is interesting to note that IFI35 and<br />

IFI44, two genes not included in the IFN score were significant<br />

predictors of death independent of the MELD score. Additional<br />

experiments performed in blood from patients with early septic<br />

shock with (n=7) or without (n= 7) cirrhosis showed that the<br />

IFN score was 2.3 higher in patients with than in those without<br />

cirrhosis (due to higher values of OAS2, IFIT1, GBP4). Conclusions:<br />

In PBMCs from patients with decompensated alcoholic<br />

cirrhosis, higher baseline levels of type 1 IFN-inducible genes<br />

involved in cell-autonomous immunity, but not expression of<br />

genes encoding secreted inflammatory cytokines/chemokines,<br />

are highly predictive of the risk of death. Deregulation of type 1<br />

IFN-inducible gene expression in circulating immune cells may<br />

play a role in the mechanisms resulting in death from cirrhosis.<br />

Disclosures:<br />

Pierre-Emmanuel Rautou - Speaking and Teaching: Gilead<br />

Dominique Valla - Advisory Committees or Review Panels: Sequana medical;<br />

Consulting: IRIS<br />

Francois Durand - Advisory Committees or Review Panels: Astellas, Novartis,<br />

BMS; Speaking and Teaching: Gilead<br />

Pierre de la Grange - Management Position: GenoSplice<br />

The following authors have nothing to disclose: Marc Pineton de Chambrun,<br />

Emmanuel Weiss, Magali Fasseu, Mikhael Giabicani, Rakhi Maiwall, Didier<br />

Lebrec, Sophie Lotersztajn, Richard Moreau<br />

1289<br />

Proportion of Intrahepatic CD56 Bright Natural Killer Cells<br />

Correlates with Well-preserved Liver Function<br />

Erin H. Doyle 1 , Adeeb Rahman 2 , Arielle L. Klepper 1 , Sang Kim 3 ,<br />

Brandy M. Haydel 4 , Sander S. Florman 5 , M. Isabel Fiel 6 , Thomas<br />

D. Schiano 5 , Andrea D. Branch 1 ; 1 Department of Liver Diseases,<br />

Icahn School of Medicine at Mount Sinai, New York, NY; 2 Human<br />

Immune Monitoring Core, Icahn School of Medicine at Mount<br />

Sinai, New York, NY; 3 Department of Anesthesiology, The Mount<br />

Sinai Hospital, New York, NY; 4 Center for Translational Transplant<br />

Research, Icahn School of Medicine at Mount Sinai, New York,<br />

NY; 5 Recanati Miller Transplantation Institute, The Mount Sinai<br />

Hospital, New York, NY; 6 Department of Pathology, The Mount<br />

Sinai Hospital, New York, NY<br />

Background/Aim: HCV-related liver damage is largely<br />

immune-mediated, but immunosuppression increases liver damage.<br />

We investigated this paradox, seeking innate immune<br />

cells whose prevalence correlates with well-preserved liver<br />

function. Methods: Immune cells were isolated from 5-25 g<br />

of tissue from 20 HCV-infected liver transplant patients using<br />

mechanical disruption and enzymatic digestion. Peripheral<br />

blood (obtained prior to surgery) and liver mononuclear<br />

leukocytes were isolated on ficoll gradients and analyzed<br />

by FACS and microarray. 8 subsets of innate immune cells<br />

were enumerated. Pearson’s correlation coefficient was used<br />

to find subsets whose prevalence correlated with liver status.<br />

Microarrays were performed on RNA from cells with or without<br />

exposure to TLR ligands. Results: Intrahepatic CD56 Bright /<br />

CD16 - NK cells were strongly correlated with well-preserved<br />

liver function and inversely related to markers of liver dysfunction:<br />

MELD score (R 2 =0.41, p=0.007), total bilirubin (R 2 =0.74,<br />

p


846A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

These findings have direct implications for our understanding<br />

of the roles of IL-22 and IL-22BP in regulating liver immunity.<br />

Disclosures:<br />

The following authors have nothing to disclose: Wagdi Almishri, Julie Deans,<br />

Mark Swain<br />

Disclosures:<br />

Thomas D. Schiano - Advisory Committees or Review Panels: salix, merck, gilead,<br />

pfizer; Grant/Research Support: galectin, massbiologics, biotest<br />

Andrea D. Branch - Grant/Research Support: Gilead, Janssen<br />

The following authors have nothing to disclose: Erin H. Doyle, Adeeb Rahman,<br />

Arielle L. Klepper, Sang Kim, Brandy M. Haydel, Sander S. Florman, M. Isabel<br />

Fiel<br />

1290<br />

Activation of hepatic innate immunity markedly alters<br />

the IL-22/IL-22R1/IL-22BP axis within the liver: Implications<br />

for hepatic immunity.<br />

Wagdi Almishri, Julie Deans, Mark Swain; Medicine, University of<br />

Calgary, Calgary, AB, Canada<br />

IL-22 is an innate cytokine that has been broadly implicated<br />

in the regulation of hepatic immunity, showning both pro- and<br />

anti-inflammatory effects in animal models of liver injury. IL-22<br />

exerts its’ biological functions by interacting with its’ receptor<br />

IL-22R1 which is widely expressed on epithelial cells and<br />

hepatocytes. Importantly, IL-22 activity is regulated by an<br />

endogenous inhibitor, IL-22BP, which binds IL-22 and prevents<br />

its’ biological function. Therefore, we performed a series of<br />

experiments to delineate alterations that occur in the IL-22/<br />

IL-22BP/IL-22R1 axis in the liver, using a mouse model of<br />

hepatic innate immune activation due to NKT cell activation.<br />

In addition, we determined changes in IL-22BP and IL-22R1<br />

expression in liver biopsy samples obtained from patients with<br />

hepatitis B and C, PBC, PSC and autoimmune hepatitis (AIH).<br />

Results: We found (by flow cytometry) the rapid and striking<br />

recruitment of IL-22 producing innate lymphoid cells (ILC3’s)<br />

into the liver, followed later by other IL-22 producing immune<br />

cells, including Th22 cells. In addition, using IL-22 knockout<br />

mice we demonstrate a proinflammatory role of IL-22 in the<br />

liver during innate immune activation. The recruitment of IL-22 +<br />

immune cells was paralleled by recruitment of IL-22BP + immune<br />

cells into the liver, and the marked up-regulation of IL-22BP<br />

expression in hepatocytes. Moreover, despite our findings of<br />

a robust influx of IL-22 + immune cells into the liver after NKT<br />

cell activation, bioavailable hepatic IL-22 levels actually significantly<br />

decreased in the 1-2 days after activation of hepatic<br />

innate immunity. Within the liver IL-22BP expression in recruited<br />

immune cells and in hepatocytes was largely dependent upon<br />

activation of the inflammasome. Viral and autoimmune liver diseases<br />

in patients have been linked to the activation of NKT cells<br />

and an enhancement of hepatic innate immunity. Therefore, we<br />

examined IL-22BP and IL-22R1 expression in liver biopsy specimens<br />

obtained from patients with liver disease, and found a<br />

striking increase in the hepatic expression of both IL-22BP and<br />

IL-22R1 in both viral and autoimmune liver diseases (ie. PBC,<br />

PSC, AIH, hepatitis B and C). Conclusion: These observations<br />

highlight a tight co-regulation of the IL-22/IL-22BP/IL-22R1<br />

axis during hepatic innate immune responses, and suggest<br />

that IL-22BP plays an important role in regulating IL-22 bioactivity<br />

within the liver in the context of innate immune activation.<br />

1291<br />

Impaired TRAF6 methylation and TLR responses in liver<br />

tissue and circulating monocytes from patients with<br />

spontaneous bacterial peritonitis<br />

Irina Tikhanovich, Jody C. Olson, Ryan Taylor, Brian Bridges, Kenneth<br />

Dorko, Benjamin R. Roberts, Steven A. Weinman; University<br />

of Kansas Medical Center, Kanas City, KS<br />

Patients with cirrhosis have a number of abnormalities in their<br />

response to infection including failure to clear bacteria from<br />

the peritoneal space or blood and an exaggerated systemic<br />

inflammatory response. Cirrhotic patients with a history of<br />

spontaneous bacterial peritonitis (SBP) have particularly severe<br />

defect in bacterial clearance. Our previous work has shown<br />

that the arginine methyltransferase PRMT1 and the demethylation<br />

enzyme JMJD6 regulate innate immune signaling<br />

pathways by controlling the arginine methylation of TRAF6,<br />

a critical ubiquitin ligase for TLR responses. We found that a<br />

decrease in the PRMT1/JMJD6 ratio results in a loss of TRAF6<br />

methylation causing preactivation of TLR pathways but reduced<br />

responses to bacterial antigens. The AIM of this study was to<br />

determine whether cell type specific defects in TRAF6 arginine<br />

methylation occur in cirrhosis and can contribute to susceptibility<br />

to infections. METHODS: Liver tissue sections and peripheral<br />

blood monocytes from patients without cirrhosis and cirrhotic<br />

patients with or without a history of SBP were analyzed for<br />

the levels of PRMT1, TRAF6 methylation and cytokine production<br />

in response to LPS using IHC staining, western blotting<br />

and proximity ligation assay. RESULTS: PRMT1 to JMJD6 protein<br />

ratios was lower in total liver protein from cirrhosis than<br />

normal donors (2.2±1.1 vs 0.68±0.43) and lower in patients<br />

with history of SBP compared to those with no history of SBP<br />

(0.92±0.48 vs 0.43±0.16). Tissue staining revealed that in<br />

patients with SBP compared to cirrhotics without an SBP history,<br />

PRMT1 was lower in non-parenchymal cells (p-value 0.029),<br />

and JMJD6 was higher both in hepatocytes and non-parenchymal<br />

cells (p-values 0.0009 and 0.025 respectively). PRMT1/<br />

JMJD6 ratio was significantly lower both in hepatocytes and<br />

non-parenchymal cells of patients with history of SBP (p-values<br />

0.044 and 0.005 respectively). TRAF6 methylation levels<br />

were directly analyzed by PLA and were similarly lower in<br />

livers of patients with history of SBP. Similar to the situation<br />

in liver, blood monocytes from patients with a history of SBP<br />

showed a similar defect in PRMT1 levels compared to monocytes<br />

from patients with cirrhosis without SBP (p-value 0.032)<br />

and the associated decrease in PRMT1/JMJD6 ratio correlated<br />

with decreased TRAF6 methylation and cytokine production<br />

in response to LPS. CONCLUSION: PRMT1 and JMJD6 levels<br />

control TRAF6 methylation and antibacterial innate immune<br />

responses. TRAF6 methylation is decreased in multiple cell<br />

types of patients with a history of SBP, which may contribute to<br />

the abnormal infection responses seen in those patients.<br />

Disclosures:<br />

Jody C. Olson - Advisory Committees or Review Panels: Baxter<br />

Steven A. Weinman - Consulting: Cardax, Inc.<br />

The following authors have nothing to disclose: Irina Tikhanovich, Ryan Taylor,<br />

Brian Bridges, Kenneth Dorko, Benjamin R. Roberts


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 847A<br />

1292<br />

Implants of matrix-embedded endothelial cells rescue<br />

ischemic tissue and liver engraftment in hepatectomized<br />

mice<br />

Pedro Melgar-Lesmes 1 , Mercedes Balcells 2,1 , Elazer R. Edelman 1,3 ;<br />

1 Edelman Lab, Institute for Medical Engineering and Science,<br />

Massachusetts Institute of Technology, Cambridge, MA, USA,<br />

Cambridge,, MA; 2 Bioengineering department, Institut Químic de<br />

Sarrià, Ramon Llull Univ, Barcelona, Spain; 3 Cardiovascular Division,<br />

Brigham and Women’s Hospital, Harvard Medical School,<br />

Boston, MA<br />

Background and aims: Ischemia during liver transplantation<br />

promotes a cascade of cellular injury responses that lead to<br />

inflammation, cell death, and organ dysfunction. Matrix-embedded<br />

endothelial cells (MEECs) have immunomodulatory<br />

effects in vitro and in vivo. Therefore we investigated the benefits<br />

of implants of MEECs as modulators of inflammation and<br />

regeneration after liver engraftment in mice. Methods: Partial<br />

hepatectomy (70%) was performed by excising the left lobe<br />

and half the median lobe of the mouse liver. Four groups of<br />

10 animals were distributed as follows: 1) acellular matrices<br />

in the interface between the remaining median lobe and an<br />

autograft from the left lobe; 2) MEECs between the remaining<br />

median lobe and an autograft from the left lobe; 3) acellular<br />

matrices between the remaining median lobe and an allograft<br />

from the left lobe of another mouse; 4) MEECs between the<br />

remaining median lobe and an allograft from the left lobe of<br />

another mouse. Vascular architecture was analysed using angiography<br />

with FITC dextran 2000 kDa by fluorescent multiphoton<br />

microscopy 7 days post-op. Hepatic DNA fragmentation<br />

and apoptosis were quantified in the median lobe and grafts<br />

by TUNEL assay and Western Blot of activated caspase 3,<br />

respectively. Serum markers of liver damage Alanine Aminotransferase<br />

(ALT) and Aspartate Aminotransferase (AST) were<br />

quantified in all animal groups. The phenotype of lymphocyte<br />

subsets in the liver after implantation of allografts was quantified<br />

by gene expression of Th1 (interferon gamma: INFγ; interleukin<br />

2: IL-2) and Th2 (interleukin 4: IL-4; interleukin 10: IL-10)<br />

genes by Real-Time PCR. Results: Implants with MEECs created<br />

a functional vascular splice between the ischemic median lobe<br />

and autografts or allografts improving the rate of liver donor<br />

regeneration by 15% compared to acellular controls. MEECs<br />

prevented apoptosis by 85% in donor liver lobes, by 85%<br />

in autologous grafts and by 79% in allogeneic engraftments.<br />

As a result, serum levels of ALT and AST were significantly<br />

reduced after autologous or allogeneic engraftments. The beneficial<br />

immunomodulatory effects of MEECs promoted allograft<br />

immunotolerance expressed as a significant reduction of Th1<br />

(INFγ and IL-2) and increase of Th2 (IL-4 and IL-10) cytokine<br />

expression. Conclusions: Implants of MEECs are endothelial<br />

cell-based scaffolds with potential to be used to improve current<br />

surgical procedures in liver transplantation and to reduce<br />

ischemic and inflammatory disorders. Their application may<br />

contribute obtaining functional liver grafts of adequate size for<br />

the recipient without subjecting the donor to undue risk.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pedro Melgar-Lesmes, Mercedes<br />

Balcells, Elazer R. Edelman<br />

1293<br />

Identification of liver monocytic myeloid-derived suppressor<br />

cells and elucidation of their roles in non-alcoholic<br />

fatty liver disease<br />

Liying Yao, Masanori Abe, Teruki Miyake, Yoshiko Nakamura,<br />

Yusuke Imai, Yohei Koizumi, Takao Watanabe, Osamu Yoshida,<br />

Masashi Hirooka, Yoshio Tokumoto, Bunzo Matsuura, Yoichi<br />

Hiasa; Department of Gastroenterology and Metabology, Ehime<br />

Universiy Graduate School of Medicine, Ehime, Japan<br />

Background/Aim: Myeloid-derived suppressor cells (MDSCs)<br />

comprise a heterogeneous population of myeloid cells and are<br />

recognized as suppressors of T cell functions. In mice, these<br />

cells identified by the co-expression of CD11b and Gr-1 surface<br />

markers. Previous <strong>studies</strong> revealed that liver MDSC exhibit<br />

phenotypic and functional diversity; however, information<br />

regarding their characteristics and mechanisms of suppression,<br />

reported in these <strong>studies</strong>, was conflicting in nature. Immune<br />

responses may contribute to the pathogenesis of non-alcoholic<br />

fatty liver disease (NAFLD); however, little is known regarding<br />

the role of myeloid regulatory cells. In this study, we characterized<br />

the phenotype of liver MDSCs in a murine model of<br />

NAFLD and explored the mechanism underlying their immunosuppression.<br />

Methods: C57BL/6 mice were fed a normal<br />

diet (ND) or a high-fat diet (HFD) for 12 months. Different subtypes<br />

of CD11b + Gr-1 + cells were sorted from liver non-parenchymal<br />

cells by FACS. CD11b + Gr-1 + cells were co-cultured<br />

with T cells in the presence of anti-CD3 beads or allogeneic<br />

dendritic cells, and suppressive functions were investigated<br />

using a carboxyfluorescein succinimidyl ester assay. Nitric<br />

oxide (NO) concentrations in culture supernatants were measured<br />

using Griess reagent. In some experiments, L-NIL, an<br />

inducible NO synthase (iNOS) inhibitor, was added at the<br />

start of the culture. Results: In the livers of HFD-fed mice, the<br />

frequency of CD11b + Gr1 dim cells (monocytic myeloid cells),<br />

but not CD11b + Gr1 hi cells, was higher than that in the livers<br />

of ND-fed mice over time. CD11b + Gr1 dim cells were divided<br />

into SSC high and SSC low populations. The frequency of SSC low-<br />

CD11b + Gr1 dim cells increased in the livers of HFD-fed mice<br />

to a greater extent than that observed in the livers of ND-fed<br />

mice. In addition, SSC low CD11b + Gr1 dim cells suppressed T cell<br />

proliferation in a dose-dependent manner; however, this suppression<br />

was not observed in SSC low CD11b + Gr1 high cells. We<br />

also found that SSC low CD11b + Gr1 dim cells expressed iNOS<br />

after co-culture with T cells, and the supernatants obtained from<br />

the co-culture of SSC low CD11b + Gr1 dim cells with T cells had<br />

higher concentrations of NO. By adding iNOS inhibitor L-NIL<br />

to the SSC low CD11b + Gr1 dim cells and T cells co-culture, T cell<br />

proliferation was restored. Conclusion: SSC low CD11b + Gr1 dim<br />

cells represent authentic monocytic MDSCs in the liver, and<br />

their numbers were increased in the livers of NAFLD mice. The<br />

suppressive function of monocytic MDSCs was dependent on<br />

NO production by iNOS. These cells might have a role in the<br />

pathogenesis of NAFLD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Liying Yao, Masanori Abe, Teruki<br />

Miyake, Yoshiko Nakamura, Yusuke Imai, Yohei Koizumi, Takao Watanabe,<br />

Osamu Yoshida, Masashi Hirooka, Yoshio Tokumoto, Bunzo Matsuura, Yoichi<br />

Hiasa


848A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1294<br />

Chronic alcohol consumption skews hepatic dendritic<br />

cell homeostasis and reveals a novel potential cellular<br />

source of TNFα<br />

Holger Fey, Costica Aloman; Division of Gastroenterology/Liver<br />

Diseases (MC716), University of Illinois at Chicago, Chicago, IL<br />

Background/Aim: Alcohol abuse is one of the leading causes<br />

of chronic liver disease and liver-related mortality. Tumor necrosis<br />

factor alpha (TNFα) is considered a central mediator in the<br />

pathogenesis of alcoholic liver damage. Dendritic cells (DC)<br />

are mononuclear phagocytes and critical regulators of innate<br />

and adaptive immunity. Several populations of DC coexist in<br />

mouse and humans: classic DC (cDC) and plasmacytoid DC<br />

(pDC). In this study we investigated the effects of chronic alcohol<br />

exposure on hepatic DC homeostasis and their contribution<br />

to the hepatic cytokine environment. Methods: Alcohol (EtOH)<br />

was delivered in the drinking water to C57BL/6 mice at 20%<br />

v/v (Meadows-Cook diet). Hepatic DC sets and bone marrow<br />

(BM) progenitors of the myeloid lineage were analyzed by flow<br />

cytometry. Circulating growth factors involved in DC development<br />

(Flt3L, GM-CSF, M-CSF) were measured by ELISA. TNFα<br />

production was assessed by intracellular cytokine staining of<br />

hepatic leukocytes isolated after in vivo toll-like receptor (TLR)<br />

stimulation with TLR4 and TLR9 agonists. Results: C57BL/6<br />

mice receiving alcohol in drinking water for 12 weeks have a<br />

70% increase (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 849A<br />

recruitment under conditions of shear stress identified inflammatory<br />

cells that migrated into HSEC and subsequently exhibited<br />

novel cell-to-cell crawling. This process was dependent on<br />

IFN-γ and occurred predominantly in HSEC when compared<br />

to vascular endothelium. Expression of chemokine transcripts<br />

and transcellular permeability were similar in both endothelia,<br />

whereas distinct differences were identified in junctional molecule<br />

expression, with HSEC lacking occludin and expressing<br />

lower levels of JAM-A than vascular endothelium. In support<br />

of these findings immunofluorescent analysis of human liver<br />

tissue demonstrated CD3+/CD4+ cells within sinusoidal endothelial<br />

cells. Conclusion - We demonstrate a novel behaviour<br />

of lymphocyte migration which occurs during interactions with<br />

primary human HSEC. This process appears to be mediated by<br />

the unique junctional expression pattern of HSEC and microenvironmental<br />

stimuli such as IFN-γ. We believe this is a novel<br />

step in the adhesion cascade that could have implications for<br />

treating chronic inflammatory liver disease and modulating<br />

intrahepatic immunity.<br />

Disclosures:<br />

David H. Adams - Advisory Committees or Review Panels: GSK, Proximagen;<br />

Grant/Research Support: Takeda, Biotie Therapies, Novimmune, ChemoCentryx,<br />

Novimmune, Biotie Therapies; Patent Held/Filed: biotie therapies, Biotie Therapies,<br />

Biotie Therapies, Biotie Therapies, Biotie Therapies, Biotie Therapies, Biotie<br />

Therapies, Biotie Therapies<br />

The following authors have nothing to disclose: Shishir Shetty, Daniel A. Patten,<br />

Chris J. Weston<br />

1297<br />

Inhibition of the FGL2:FcγRIIB Immunosuppressive Pathway<br />

Restores Antiviral T and B Cell Responses During<br />

Chronic Viral Infection<br />

Ramzi Khattar, Olga Luft, Kaveh Farrokhi, Vanessa Rojas Luengas,<br />

Hassan Sadozai, Gary Levy, Nazia Selzner; University Health<br />

Network, Toronto, ON, Canada<br />

Background and Aims: HBV and HCV utilize host immunosuppressive<br />

pathways to limit antiviral T cell responses and promote<br />

the generation of dysfunctional T cells in a process termed<br />

T cell exhaustion. Recent reports have implicated the upregulation<br />

of immunosuppressive cytokines and involvement of Tregs<br />

as causative agents of persistent HBV and HCV infection. FGL2<br />

is a potent immunosuppressive effector molecule of CD4 + C-<br />

D25 + TIGIT + FOXP3 + Treg. The role for FGL2 in establishing T<br />

cell exhaustion in chronic viral infection caused by LCMV Cl-13<br />

was assessed. Methods: Fgl2 +/+ and fgl2 -/- mice were infected<br />

with 2x10 6 PFU of LCMV Cl-13. FGL2 was measured in plasma<br />

by ELISA. Frequencies of LCMV specific T cells were determined<br />

following stimulation with viral peptide. Neutralizing<br />

antibody titers were evaluated by plaque reduction assay. Viral<br />

titers in plasma and liver were assessed by an LCMV focus<br />

forming assay. Fgl2 +/+ were treated with neutralizing antibodies<br />

toward FGL2 and FCγRIIB and subsequently infected with<br />

2x10 6 PFU of LCMV Cl-13. LCMV antiviral immune responses<br />

were measured. Results: Following infection with 2x10 6 PFU of<br />

LCMV Cl-13, plasma levels of FGL2 in Fgl2 +/+ mice increased<br />

(51.6±2.4 ng/mL) reaching maximum levels on day 7 pi<br />

(135.6±7.0 ng/mL; P


850A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

αEβ7 suggest homing to the peri-biliary region where LI MAIT<br />

cells are activated in response to bacteria-exposed BEC in an<br />

MR1 dependent manner, leading to liver injury through their<br />

release of cytolytic enzymes and inflammatory cytokines.<br />

Disclosures:<br />

David H. Adams - Advisory Committees or Review Panels: GSK, Proximagen;<br />

Grant/Research Support: Takeda, Biotie Therapies, Novimmune, ChemoCentryx,<br />

Novimmune, Biotie Therapies; Patent Held/Filed: biotie therapies, Biotie Therapies,<br />

Biotie Therapies, Biotie Therapies, Biotie Therapies, Biotie Therapies, Biotie<br />

Therapies, Biotie Therapies<br />

The following authors have nothing to disclose: Hannah C. Jeffery, Bonnie Van<br />

Wilgenburg, Kathryn Stirling, Krishan Parekh, Sheree Roberts, Ayako Kurioka,<br />

Stuart Hunter, Paul Klenerman, Ye H. Oo<br />

1299<br />

Alterations in liver dendritic cell subtypes during non-alcoholic<br />

steatohepatitis<br />

Eva-Carina Heier, Hannes Borchardt, Henrike Julich, Christoph<br />

Eckert, Niklas Klein, Thomas Tschernig, Veronika Lukacs-Kornek;<br />

Department of Medicine II, Saarland University, Homburg, Germany<br />

BACKGROUND/AIMS: Dendritic cells (DCs) represent a heterogeneous<br />

subpopulation that is primarily identified as<br />

CD11c + MHCII + cells in the liver. However, these markers are<br />

not exclusively identifying DCs rather they are relevant to a<br />

mixture of myeloid/lymphoid cells. Nevertheless, CD11c + cells<br />

limited inflammation by reducing the destructive effect of innate<br />

cells and promoting the clearance of cellular debris in non-alcoholic<br />

steatohepatitis (NASH). The various DC subsets have different<br />

roles in liver homeostasis and divergent factors determine<br />

their functional operation during inflammation (Lukacs-Kornek<br />

V. J Hepatol. 2013 Nov;59(5):1124-6). In order to clarify the<br />

nature and subset of DCs involved in NASH, the study aimed<br />

to characterize the CD11c + myeloid infiltrates and dissect the<br />

role of the various DC subtypes during disease progression.<br />

METHODS: CD11c + cell infiltrate was analyzed in methionine<br />

choline deficient (MCD) diet induced liver steatohepatitis in the<br />

progression and regression phase. Using multicolor FACS analyses,<br />

the myeloid compartment was mapped during disease<br />

progression and the individual DC subtypes, and precursor<br />

DCs were identified and distinguished from monocytes, macrophage<br />

and neutrophils. RESULTS: During disease progression,<br />

CD11c + cells were increased and among these cells, the<br />

CD11c + CD11b + myeloid DC subtype increased an average of<br />

2.1 fold in 1wk and 2.2 fold in 5wks MCD treated animals.<br />

Not only the myeloid but also the lymphoid DC subtypes followed<br />

similar changes. Accordingly, the frequency of lymphoid<br />

DCs (CD103 + CD11c + CD11b - ) increased during disease progression<br />

(2.2 and 2.3 fold increase in 1wk and 5wks MCD<br />

treated animals). During regression of steatohepatitis, these<br />

changes were reduced and after 1wk of regression period<br />

DC parameters were similar to control animals. Functionally,<br />

both DC subtypes produced high amount of inflammatory cytokines<br />

(e.g. TNFα), (30%-40% of myeloid DCs and 10% of lymphoid<br />

DCs). Besides, the precursor DC population was more<br />

abundant in 1wk MCD treatment and further elevated in 5wks<br />

MCD treatment. CONCLUSIONS: During MCD diet induced<br />

steatohepatitis both DC subtypes show increased abundance<br />

and exhibited similar fold changes suggesting the importance<br />

of both DC subtypes in the pathomechanism of NASH. DC<br />

changes are paralleled by alterations observed in the precursor<br />

DC population suggesting the relevance of local DC differentiation<br />

during steatohepatitis. Studies were supported by<br />

the Alexander von Humboldt Foundation – Sofja Kovalevskaja<br />

Award 2012 to VLK.<br />

Disclosures:<br />

The following authors have nothing to disclose: Eva-Carina Heier, Hannes<br />

Borchardt, Henrike Julich, Christoph Eckert, Niklas Klein, Thomas Tschernig,<br />

Veronika Lukacs-Kornek<br />

1300<br />

Soluble CD163 plasma concentration is related with<br />

necroinflammatory activity and liver function impairment<br />

in compensated cirrhotic patients and is not<br />

related with bacterial translocation<br />

Francesco Figorilli 2,1 , Karen Louise Thomsen 2 , Jane Mac-<br />

Naughtan 2 , I.Jane Cox 3 , Alba Moratalla 4 , Isidora Ranchal 2 ,<br />

Holger J. Møller 5 , Rajeshwar Mookerjee 2 , Nathan Davies 2 , Rajiv<br />

Jalan 2 ; 1 Universita’ di Cagliari, Cagliari, Italy; 2 Liver Failure<br />

Group, UC, Institute for Liver and Digestive Health, UCL Medical<br />

School, Royal Free Campus, London, United Kingdom; 3 nstitute of<br />

Hepatology, London, Foundation for Liver Research, 69-75 Chenies<br />

Mews, London, United Kingdom; 4 Liver unit, Hospital General<br />

Universitario de Alicante, CIBER de Enfermedades Hepáticas y<br />

Digestivas (CIBERehd), Alicante, Spain; 5 Department of Clinical<br />

Biochemistry, Aarhus University Hospital, Aarhus, Denmark<br />

Background: CD163 is a scavenger receptor expressed on<br />

monocytes and macrophages e.g. Kupffer cells in the liver.<br />

Plasma levels of soluble CD163 (sCD163, normal range 0.89-<br />

3.95 mg/L) are associated with macrophage activation and<br />

grade of portal hypertension and provide prognostic information<br />

in cirrhosis. The mechanism underlying this increase<br />

in sCD163 is not clear. In this study we aimed to determine<br />

whether there was a relationship between the level of sCD163<br />

and gut permeability and markers of bacterial translocation<br />

or severity of liver disease. Methods: 66-clinically-stable liver<br />

cirrhosis out-patients were included (Male 69.7%; Age 57<br />

±9.15; Alcoholic liver disease 57.6%). Exclusion criteria were<br />

a Child Pugh score >10 and clinical or microbiological evidence<br />

of infection 7 days prior to enrolment. sCD163 plasma<br />

level was measured by an ELISA test, gut permeability using<br />

the lactulose/rhamnose test, bacterial DNA by PCR reaction<br />

and LPS concentration was obtained by HEK-Blue LPS Detection<br />

Kit 2s. Urine was collected for proton magnetic resonance<br />

spectroscopy ( 1 HMRS). Patients were divided accordingly to<br />

their sCD163 level (3.95). For the statistical analysis<br />

we used Chi-square, Student T-test, Man Whitney test and<br />

Spearman correlation. Results: The median value of sCD163<br />

was 4.33mg/L with a range of 1.44-15.43. We did not find<br />

any significant correlation between the levels of sCD163, gut<br />

permeability (0.028, range 0.002-0.112), or LPS concentration<br />

(0.02ng/ml, range 0-14.1). There was no difference in<br />

the prevalence of bacterial DNA positivity in patients with an<br />

abnormal value of sCD163 (11.1% vs. 2.6%). 39/66 (59.1%)<br />

patients with a sCD163>3.95mg/L showed a significant<br />

higher MELD value (9.0 vs. 7.5) and higher levels of AST (43<br />

vs. 30 U/L), ALP (102 vs. 70U/L), GGT (98 vs. 48 U/L) and<br />

lower level of platelets (107 vs. 178 x 10 9 ). Child-Pugh B was<br />

more common in patients with high sCD163 levels (17,9 vs.<br />

0%)). No significant differences were observed between the<br />

two groups regarding aetiology, alcohol consumption, drug<br />

use, comorbidity, previous episode of liver decompensation or<br />

infection in the previous month. At 1 HMRS analysis, patients<br />

with sCD163>3.95 showed no significant difference in urinary<br />

level of hippuric acid, formate, citrate or lactate. Conclusion:<br />

Our results suggest that Kupffer cell activation is an early<br />

event in the pathogenesis of chronic liver disease, which is<br />

not dependent of gut permeability or bacterial translocation.<br />

sCD163 concentration might reflect the grade of necroinflammation<br />

and liver function impairment in compensated cirrhotic<br />

patients.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 851A<br />

Disclosures:<br />

Holger J. Møller - Grant/Research Support: Danish Council for Strategic<br />

Research; Independent Contractor: IQ-Products, NL; Patent Held/Filed: Aarhus<br />

University; Stock Shareholder: Affinicon Aps<br />

Rajiv Jalan - Consulting: Ocera Therapeutics, Conatus; Grant/Research Support:<br />

Grifols, Gambro; Patent Held/Filed: Yaqrit, Cyberliver<br />

The following authors have nothing to disclose: Francesco Figorilli, Karen Louise<br />

Thomsen, Jane MacNaughtan, I.Jane Cox, Alba Moratalla, Isidora Ranchal,<br />

Rajeshwar Mookerjee, Nathan Davies<br />

1301<br />

Hepatitis B X-interacting protein (HBXIP) negatively<br />

regulates IFN-β production and antiviral response by<br />

promoting proteasomal degradation of TANK-binding<br />

kinase 1<br />

Hang Zhang; Basic medical, Hangzhou Normal University, Hangzhou,<br />

China<br />

TANK-binding kinase 1 (TBK1) plays an essential role in Tolllike<br />

receptor (TLR)– and retinoic acid–inducible gene I (RIG-I)–<br />

mediated induction of type I interferon (IFN; IFN-α/β) and host<br />

antiviral responses. How TBK1 activity is negatively regulated<br />

remains largely unknown. We report that Hepatitis B virus X-interacting<br />

protein (HBXIP) promotes proteasomal degradation of<br />

TBK1 and inhibits RIG-I-induced IFN-β signaling. HBXIP knockdown<br />

resulted in augmented activation of IFN regulatory factor<br />

3 (IRF3) and enhanced expression of IFN-β in RIG-I-activated<br />

primary hepatocyte cells, whereas overexpression of HBXIP<br />

had opposite effects. Consistently, HBXIP impaired vesicular<br />

stomatitis virus (VSV) infection–induced IRF3 activation and<br />

IFN-β production and promoted VSV replication. HBXIP negatively<br />

regulated the cellular levels of TBK1 by directly binding<br />

to and promoting K48-linked polyubiquitination of TBK1. Therefore,<br />

we identified HBXIP as a negative regulator in RIG-I–triggered<br />

antiviral responses and suggested HBXIP as a potential<br />

target for the intervention of diseases with uncontrolled IFN-β<br />

production.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hang Zhang<br />

1302<br />

Toll-like receptor 9 and B-cell activating factor signals<br />

induce impaired immunological memory in cirrhosis<br />

Hiroyoshi Doi 1 , David E. Kaplan 3 , Eiichi Hayashi 1 , Jun Arai 1 , Risa<br />

Omori 1 , Manabu Uchikoshi 1 , Kenichi Morikawa 4 , Junichi Eguchi 2 ,<br />

Takayoshi Ito 2 , Hitoshi Yoshida 1 ; 1 Gastroenterology, Showa University,<br />

Shinagawa, Japan; 2 Digestive Disease Center, Showa University<br />

Koto Toyosu Hospital, Toyosu, Japan; 3 Gastroenterology,<br />

University of Pennsylvania, Philadelphia, PA; 4 Gastroenterology<br />

and Hepatology, Hokkaido University, Sapporo, Japan<br />

[Background] Cirrhosis is an immunocompromised state and<br />

their immune cells are thought to be impaired. We previously<br />

reported peripheral CD27+ memory B-cells in cirrhotic patients<br />

are decreased and dysfunctional compared to those in even<br />

non-cirrhotic chronic liver disease patients. We also found Tolllike<br />

receptor (TLR) 4 and TLR 9 signals play important roles for<br />

B-cell activation and survival. (Doi et al. Hepatology 2012).<br />

Clinically, we confirmed elevated immunoglobulin levels in cirrhosis<br />

regardless of the etiology. The abnormal humoral immunity<br />

could be caused by antigen non-specific way such as TLR<br />

signals and results in polyclonal B-cell activation. The purpose<br />

of this study is to investigate the impact of antigen non-specific<br />

signals on immunological change and immunodeficiency<br />

in cirrhosis. [Methods] We isolated peripheral blood mononuclear<br />

cells (PBMC) and serum from whole blood. Then we<br />

measured memory B-cell and memory T-cell defined by CCR7<br />

and CD45RA expressions using PBMC by flow cytometry. As<br />

we previously reported, B-cell stimulants such as TLR ligands<br />

increased in cirrhotic plasma probably due to bacterial translocation<br />

and we investigated TLR 4 and TLR 9 expressions on<br />

B-cell. B-cell activating factors (BAFF and APRIL) in sera were<br />

measured by ELISA. To analyze B-cell signaling, we negatively<br />

isolated CD19+ B-cell from healthy donor PBMC and cultured<br />

the B-cell for 72 hours with LPS (TLR4 agonist), CpG-ODN<br />

(TLR9 agonist), anti-IgG/A/M (B-cell receptor agonist) and<br />

combinations. Then we measured activation marker expressions<br />

and apoptosis by flow cytometry. [Results] In accordance<br />

with our previous study, cirrhotic patients (CIR) had decreased<br />

peripheral memory B-cell compared to non-cirrhotic chronic<br />

liver disease patients (CLD). (CLD vs CIR; p=0.021) They also<br />

showed lower frequency of both CD4+ and CD8+ effector<br />

memory T-cells. (p= 0.003, 0.027, respectively) TLR9 but not<br />

TLR4 expression on memory B-cell was increased in cirrhotic<br />

patients. (p=0.043) We also found serum BAFF but not APRIL<br />

level was elevated as well. (p=0.003) Memory B-cell frequency<br />

tended to be associated with the TLR 9 expression and BAFF<br />

levels. (p=0.016, 0.025, respectively) These findings are more<br />

apparent in advanced cirrhosis. In vitro, B-cell stimulants shown<br />

above didn’t make much impact on CD27 expression but<br />

CD27- B-cells, which relatively dominant in cirrhosis, tended<br />

to be apoptotic by TLR 4 and TLR 9 stimulations. [Conclusion]<br />

Cirrhotic patients especially in advanced stage have elevated<br />

BAFF and TLR9 expression on memory B-cell. Those antigen<br />

non-specific signals have an impact on dysfunctional immunological<br />

memory observed in cirrhosis.<br />

Disclosures:<br />

David E. Kaplan - Grant/Research Support: Bayer Pharmaceuticals, Inovio Pharmaceuticals<br />

The following authors have nothing to disclose: Hiroyoshi Doi, Eiichi Hayashi,<br />

Jun Arai, Risa Omori, Manabu Uchikoshi, Kenichi Morikawa, Junichi Eguchi,<br />

Takayoshi Ito, Hitoshi Yoshida<br />

1303<br />

Low Dose Zinc Sulfate (220mg) Supplementation for<br />

Three Months Normalizes Zinc Levels, Endotoxeima,<br />

Pro-Inflammatory/Fibrotic Biomarkers & Improves Clinical<br />

Parameters in Alcoholic Cirrhosis– A Double-Blind<br />

Placebo Controlled – (ZAC) Clinical Trial<br />

Mohammad K. Mohammad 1,2 , Keith C. Falkner 1 , Ming Song 1 ,<br />

Craig J. McClain 1,2 , Matthew C. Cave 1,2 ; 1 Hepatology, gastroenterology<br />

and nutrition, university of louisville, Louisville, KY; 2 Hepatology,<br />

Robely Rex Veteran Hospital, Louisville, KY<br />

Purpose: Zinc deficiency occurs in human subjects with alcoholic<br />

cirrhosis (AC), and zinc supplementation attenuates liver<br />

injury/inflammation in murine models of alcoholic liver disease.<br />

The aim of this NIH-funded clinical trial was to determine<br />

if zinc sulfate therapy improves serologic biomarkers of liver<br />

injury/inflammation & clinical status in AC Methods: 22 consented<br />

subjects with Child-Pugh class A-B alcoholic cirrhosis<br />

were randomized to placebo (n=11) or zinc sulfate 220 mg<br />

daily (n=11) in the single center, IRB approved, double-blind,<br />

placebo-controlled clinical trial. 10 non-drinking, healthy volunteers<br />

were recruited as controls (HC). Serum cytokines (IL-1β,<br />

IL-2, IL-4, IL-6, IL-8, IL-10, IL-17, IL-18, IFN-γ, PAI-1, MCP-1 &<br />

TNF-α) and whole blood ex-vivo unstimulated and lipopolysacharide-stimulated<br />

(LPS) cytokine production were measured<br />

by Luminex, while TGF-β & LPS were measured by ELISA. Differences<br />

between means were evaluated by t-test, p < 0.05<br />

was considered significant Results: Of the 22 AC subjects,<br />

20 subjects completed 3 months treatment: placebo (n=10);


852A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

zinc (n=10). Demographic variables were similar between<br />

groups. At baseline, means for Zinc, AST, ALT,Albumin, CTP<br />

& MELD were 56.9±14; 42.3±15.88; 17.5±10.11; 3.5±0.7;<br />

6.9±1.22; 9.4±2.11 in the Zinc group, and 62.58±22.52;<br />

51.7±24.42; 27.3±17; 3.9±0.4, 5.6±0.9; 8.5±0.85 for<br />

placebo. There were more subjects with active EtOH drinking,<br />

varices & ascites (7/9/4) in the zinc arm vs. (2/6/1)<br />

in the placebo At baseline, when compared to HC, AC had<br />

increased mean serum LPS, TNF-α, IL-6, IL-8 while decreased<br />

mean Zinc. AC patients had increased mean ex-vivo unstimulated<br />

& LPS-stimulated production (Priming) of IL-6, IL-8, IL-18<br />

and TNF-α. No differences were observed for IL-1β, IL-2, IL-4,<br />

IL-10, IL-17 and IFN-γ in AC vs. HC with & without LPS stimulation.<br />

At 3 months, mean Zinc, LPS, TGF-β, TNF-α, IL-18 & Child-<br />

Pugh significantly improved in the Zinc group, while only IL-18<br />

improved in the placebo arm. In the zinc group, ex-vivo unstimulated<br />

& LPS stimulated production of TNFα & IL-6 normalized<br />

at 3 months together with unstimulated IL-18 & LPS stimulated<br />

production of IL-1β, no difference was observed in placebo.<br />

There were no side effects for zinc treatment Conclusions: Subjects<br />

with alcoholic cirrhosis had increased biomarkers of liver<br />

injury, and inflammation compared to healthy controls which<br />

improved with zinc supplementation. To our knowledge, this<br />

was first human study to demonstrate the safety & efficacy of 3<br />

months Zinc supplementation in improving zinc levels, endotoxemia,<br />

pro-inflammatory biomarkers (TNF-α, TGF-β, IL-1β, IL-6<br />

and IL-18) & clinical status in alcoholic cirrhosis<br />

Disclosures:<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

Matthew C. Cave - Advisory Committees or Review Panels: Intercept, Abbvie;<br />

Consulting: Abbvie, Diapharma; Grant/Research Support: Merck, Gilead, Intercept,<br />

Conatus, Lumena, Cepheid, Tobira, Galectin, Bayer; Speaking and Teaching:<br />

BMS, Abbvie, Gilead, Janssen, Genentech<br />

The following authors have nothing to disclose: Mohammad K. Mohammad,<br />

Keith C. Falkner, Ming Song<br />

1304<br />

Acrolein, a lipid-derived aldehyde, is a critical pathogenic<br />

mediator and potential therapeutic target for alcoholic<br />

liver disease<br />

Wei-Yang Chen 1 , Jingwen Zhang 1 , Shirish Barve 1 , Craig J.<br />

McClain 1,2 , Swati Joshi-Barve 1 ; 1 Department of Medicine/GI, University<br />

of Louisville, Louisville, KY; 2 Robley Rex VAMC, Louisville,<br />

KY<br />

Purpose: Alcoholic liver disease (ALD) remains a major cause<br />

of morbidity and mortality, with no FDA approved therapy.<br />

Chronic alcohol consumption causes a pro-oxidant environment<br />

in the liver and increases hepatic lipid peroxidation. Acrolein<br />

is the most reactive and toxic aldehyde generated by lipid<br />

peroxidation, and is known to form covalent protein adducts.<br />

We have shown that acrolein exposure in hepatocytes triggers<br />

endoplasmic reticulum (ER) stress, which is an important etiologic<br />

factor in ALD. This study (i) investigates the pathogenic<br />

role of acrolein in alcohol-induced hepatic ER stress, steatosis,<br />

and liver injury in experimental ALD, and (ii) tests acrolein<br />

elimination/scavenging (using hydralazine) as a potential<br />

therapeutic strategy against ALD. Methods: We used in vitro<br />

(rat hepatic H4IIEC cells) and in vivo (chronic+binge (NIAAA)<br />

alcohol feeding in C57Bl/6 mice) models to study the role of<br />

acrolein in ALD. Also, the potential protective effects of the<br />

acrolein scavenger, hydralazine, were examined in vitro and<br />

in vivo. Alcohol-induced acrolein accumulation was detected<br />

by immunostaining for acrolein-protein adducts. The effects of<br />

alcohol-induced acrolein were assessed on (i) hepatic steatosis<br />

(H&E and Oil Red O); (ii) ER stress (ATF3, ATF4, GRP78,<br />

GRP94 mRNA and protein); (iii) pro-apoptotic signaling (activation<br />

of JNK and caspase-12 and pro-apoptotic CHOP); (iv)<br />

cell death (MTT) and hepatocyte apoptosis (TUNEL); (v) liver<br />

injury (serum ALT and AST). Results: We demonstrate that alcohol<br />

exposure resulted in substantial hepatic accumulation of<br />

acrolein-protein adducts, which was dependent on alcohol<br />

metabolism via cytochrome P4502E1 and alcohol dehydrogenase.<br />

Alcohol-induced acrolein accumulation led to hepatic<br />

ER stress and upregulation of ATF3 and ATF4, with minimal<br />

induction of the protective chaperones GRP78 and GRP94.<br />

A concomitant increase was seen in proapoptotic signaling<br />

with activation of JNK and caspase12, and CHOP, resulting<br />

in significant hepatic steatosis, hepatocyte cell death and liver<br />

injury. Acrolein was able to mimic the in vivo adverse effects<br />

of alcohol in cultured H4IIEC cells. Importantly, hydralazine, a<br />

known acrolein scavenger, protected against alcohol-induced<br />

ER stress, steatosis, apoptosis and liver injury, both in vitro and<br />

in vivo. Conclusions: Our study demonstrates that acrolein is a<br />

major mediator of alcohol-induced hepatic ER stress, cell death<br />

and liver injury. Notably, our study also shows that removal/<br />

clearance of acrolein by scavengers may have therapeutic<br />

potential in attenuating the adverse effects of alcohol consumption,<br />

and preventing ALD.<br />

Disclosures:<br />

Shirish Barve - Speaking and Teaching: Abbott<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Wei-Yang Chen, Jingwen Zhang,<br />

Swati Joshi-Barve<br />

1305<br />

Intestinal HIF-1α Deletion Exacerbates Alcohol-induced<br />

Hepatic Steatosis Mediated by Intestinal Dysbiosis and<br />

Barrier Dysfunction<br />

Tuo Shao 1 , Liming Liu 1 , Fengyuan Li 1 , Lihua Zhang 1 , Craig J.<br />

McClain 1,2 , Wenke Feng 1 ; 1 Department of Medicine/GI, University<br />

of Louisville, Louisville, KY; 2 Robley Rex VAMC, Louisville, KY<br />

Background: Alcoholic liver disease (ALD) is characterized by<br />

increased gut permeability and bacterial translocation. Hypoxia<br />

Induced Factor 1α (HIF-1α) has been implicated in transcriptional<br />

regulation of intestinal barrier integrity and inflammation.<br />

However, the link between intestinal HIF-1α regulation and ALD<br />

is unclear. We tested our hypothesis that HIF-1α plays a critical<br />

role in gut microbiota homeostasis and the maintenance of<br />

intestinal barrier integrity in a mouse model of ALD. Methods:<br />

Wide type (WT) and intestinal-specific HIF-1α knockout mice<br />

(HIF-1α -/- ) were pair-fed modified Lieber-DeCarli liquid diets<br />

containing 5% (w/v) alcohol (AF) or isocaloric maltose dextrin<br />

(PF) for 24 days. Blood samples were collected for determination<br />

of ALT and AST, endotoxin level and bacteria DNA.<br />

The liver and intestine tissues were used for histology staining,<br />

gene expression and hepatic triglyceride (TG) content measurements.<br />

Ex vivo ileal permeability was determined. Intestinal<br />

microbiota were determined in cecel samples Results: Alcohol<br />

feeding significantly increased serum levels of ALT, AST and<br />

LPS in the WT mice, indicating liver injury and increased endotoxemia<br />

by alcohol exposure. These elevations were more pronounced<br />

in HIF1α -/- mice. Histology examination revealed that<br />

hepatic lipid accumulation was increased by alcohol exposure,<br />

and it was exaggerated in HIF1α -/- mice. Liver triglyceride<br />

concentrations were increased by alcohol exposure in the WT<br />

mice and further elevated in the KO mice, and these were associated<br />

with up-regulation of hepatic SREBP-1c and down-regulation<br />

of PPAR-α. Intestinal HIF-1α deletion resulted in a marked


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 853A<br />

increase in hepatic IL-6 and IL-1β expression and neutrophil<br />

infiltration as showed by CAE staining, which were associated<br />

with TLR4 up-regulation. In the HIF-1α -/- mice, alcohol exposure<br />

disrupted intestinal structure integrity as shown by H&E staining<br />

of ileal sections. 16s RNA analysis of the cecal and fecal<br />

samples from HIF1α -/- mice exposed to alcohol showed an elevated<br />

dysbiosis. PCR analysis revealed that alcohol exposure<br />

decreased the expression of tight junction gene, occludin, and<br />

CRAMP, P-gp, and ITF, which are HIF-1α transcription target<br />

genes and important for gut microbiota homeostasis and intestinal<br />

barrier integrity. Ex vivo assay demonstrated that HIF1α<br />

-/-<br />

increased alcohol-induced ileal permeability. Conclusions:<br />

Our results demonstrated that intestinal HIF-1α is required for<br />

the adaptation response to alcohol exposure-induced changes<br />

in intestinal microbiota and barrier function leading to elevated<br />

endotoxemia and hepatic steatosis and injury.<br />

Disclosures:<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Tuo Shao, Liming Liu, Fengyuan<br />

Li, Lihua Zhang, Wenke Feng<br />

1306<br />

Fibroblast growth factor 21 is critically involved in alcohol-induced<br />

hepatic lipid metabolism and adipose tissue<br />

lipolysis in mice<br />

Cuiqing Zhao 1 , Liming Liu 1 , Ying Yang 1 , Craig J. McClain 1,2 ,<br />

Wenke Feng 1 ; 1 Department of Medicine/GI, University of Louisville,<br />

Louisville, KY; 2 Robley Rex VAMC, Louisville, KY<br />

Purpose: Fibroblast growth factor 21 (FGF21) is a novel<br />

metabolic regulator produced primarily in the liver that regulates<br />

hepatic fat metabolism in various models of metabolic<br />

syndromes. Alcoholic liver disease (ALD) is characterized by<br />

hepatic lipid accumulation. Here, we sought to determine<br />

the role of FGF21 in mice exposed to alcohol and to discern<br />

underlying mechanisms. Methods: Male FGF21 knockout (KO)<br />

and control (WT) mice were fed either the Lieber DeCarli diet<br />

containing 5% alcohol (AF) or an isocaloric diet (PF). For the<br />

chronic alcohol exposure, mice were fed for 4 weeks. For the<br />

chronic-binge alcohol exposure, mice were fed for 12 days,<br />

followed by a single dose of alcohol (5g/kg) or isocaloric<br />

maltose dextrin treatment by gavage. Mice were sacrificed<br />

6 hours later. One group of AF mice were administered with<br />

recombinant human FGF21 (rhFGF21) for the last 5 days.<br />

Liver steatosis and inflammation and epididymal white adipose<br />

tissue (eWAT) lipolysis were investigated. Results: Alcohol<br />

exposure induced plasma FGF21 elevation. FGF21 knockout<br />

exacerbated chronic alcohol-induced hepatic steatosis and<br />

liver injury, which was associated with increased activation<br />

of genes involved in lipogenesis mediated by SREBP1c and<br />

decreased expression of genes involved in fatty acid (FA)<br />

β-oxidation mediated by PGC1α. Chronic alcohol exposure<br />

induced a moderate adipose tissue lipolysis in eWAT. rhFGF21<br />

administration reduced alcohol-induced hepatic steatosis and<br />

inflammation in WT mice. Surprisingly, in the chronic-binge<br />

model, alcohol consumption-induced fatty liver is blunted in<br />

the KO mice. Chronic-binge alcohol exposure-induced in situ<br />

hepatic lipogenesis and FA β-oxidation were comparable to<br />

that in chronic alcohol-exposed mice. However, FGF21 deficiency<br />

markedly reduced chronic-binge alcohol exposure-induced<br />

eWAT lipolysis, which may contribute to hepatic FA<br />

uptake. Further analysis showed that in the KO mice alcohol<br />

induced significant elevation in systemic catecholamine, which<br />

is known to promote adipose lipolysis. rhFGF21 administration<br />

increased alcohol- induced fat loss in parallel with an increase<br />

of circulating norepinephrine concentration in KO mice. Conclusion:<br />

Our results demonstrated that the role of FGF21 in ALD<br />

is alcohol exposure model dependent. Lack of FGF21 exacerbates<br />

chronic alcohol exposure-induced fatty liver, and this<br />

exaggeration is overridden in chronic-binge alcohol exposed<br />

mice through a reduction of adipose lipolysis mediated by<br />

systemic catecholamine release involving sympathetic nervous<br />

system activation. Pharmacologic strategies targeting hepatic<br />

FGF21 elevation and reduction in adipose tissue lipolysis are<br />

warranted.<br />

Disclosures:<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Cuiqing Zhao, Liming Liu, Ying<br />

Yang, Wenke Feng<br />

1307<br />

Early prediction of response in patients with severe<br />

acute alcoholic hepatitis by using Lille model on day 4<br />

Mauricio Garcia-Saenz de Sicilia 1 , Chitharanjan Duvoor 2 , Jose<br />

Altamirano 4,5 , Veronica Prado 3,4 , Juan Caballeria 3,4 , Andres<br />

Duarte-Rojo 1 ; 1 Gastroenterology and Hepatology, University of<br />

Arkansas for Medical Sciences, Little Rock, AR; 2 Internal Medicine,<br />

University of Arkansas for Medical Sciences, Little Rock, AR; 3 Liver<br />

Unit, Hospital Clinic, Barcelona, Spain; 4 Institut d’Investigacions<br />

Biomediques August Pi I Sunyer (IDIBAPS), Barcelona, Spain; 5 Vall<br />

d’Hebron Institut de Recerca, Barcelona, Spain<br />

BACKGROUND: Corticosteroids are the mainstay of treatment<br />

for severe alcoholic hepatitis (SAH) associated with a lower<br />

28-day mortality. Lille model (LM) can identify non-responders<br />

after 7 days of treatment, however this requires longer<br />

exposure and increases the risk of infection. We aimed to<br />

evaluate LM at day 4(LM4) of treatment with corticosteroids<br />

in comparison to day 7(LM7). METHODS: We performed a<br />

retrospective analysis of a multinational cohort with SAH based<br />

on a discriminant function (DF)>32. Response to corticosteroids<br />

was assessed with LM7 and LM4 according to the validated<br />

cutoff value (CUV>0.45). ROC-curves were constructed to<br />

determine an ideal CUV for LM4, and to compare accuracy<br />

of LM4, LM7, MELD, and ABIC. Logistic regression models<br />

were constructed to predict 90-day mortality. RESULTS: A total<br />

of 132 (83.9%) out of 162 patients with SAH received corticosteroids.<br />

Mean age was 47.87±10.58, 98 (60.49%) were<br />

males, and BMI was 27.58±6 kg/m 2 . On admission the mean<br />

bilirubin was 17±10 mg/dL, creatinine 0.97 (0.7-1.8) mg/<br />

dL, INR 1.92±0.63. Severity parameters on admission were<br />

DF 58.34±33.38, MELD 24.98±8.79, and ABIC 8.18±1.56.<br />

Although mean LM4 and LM7 were statistically different<br />

(0.56±0.03 and 0.52±0.03, p=0.005), the proportion of<br />

responders was similar 45%(n=53) and 49%(n=56) p=0.08,<br />

and there was agreement in 90% of cases (k=0.80). Three<br />

patients became responders between days 4 and 7 (all died).<br />

Accuracy of SAH scores to predict 90-day mortality (AUROC)<br />

was: MELD 0.687 (0.595-0.778), ABIC 0.729 (0.648-0.809),<br />

LM4 0.686 (0.593-0.775), LM7 0.723 (0.633-0.813). In univariate<br />

analysis LM4, SIRS, infection, hepatic encephalopathy,<br />

and gastrointestinal bleeding predicted mortality, however, in<br />

the multivariate analysis only LM4 (OR=4.67 [1.40-15.54])<br />

and GIB (OR=4.94 [1.40-15.25]) remained significant, with a<br />

trend for HE (OR=1.86 [0.96-4.55]). The operational characteristics<br />

of LM4 and LM7 as predictors of 90-day mortality are<br />

shown in table-1. CONCLUSION: LM4 is good as LM7 as predictor<br />

of 90-day mortality or response to corticosteroids. Using<br />

LM4 instead of LM7 would help decrease exposure to corticosteroids,<br />

infection risk, and health care costs. Although there


854A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

was a 4% difference in corticosteroid response between LM4<br />

and LM7, the rate of responders that survived was unchanged<br />

(36%). Whether earlier responders at day 4 have a favorable<br />

prognosis need to be investigated.<br />

Table-1 Operational characteristics of Lille Model on day 4 and<br />

day 7<br />

a known inhibitor of autophagy, and AMPK, that activates<br />

autophagy due to dephosphorylation of these counter regulatory<br />

signaling molecules. We also observed that the activity of<br />

protein phosphatase 2A, a critical dephosphorylating enzyme,<br />

increased while the activity of PI3Kγ, an upstream inhibitor of<br />

PP2A, was reduced by alcohol.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gangarao Davuluri, Samjhana<br />

Thapaliya, Avinash Kumar, Megan R. McMullen, Rebecca L. McCullough, Laura<br />

E. Nagy, Sathyamangla V. Naga Prasad, Srinivasan Dasarathy<br />

Disclosures:<br />

Andres Duarte-Rojo - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Vital Therapies; Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Mauricio Garcia-Saenz de<br />

Sicilia, Chitharanjan Duvoor, Jose Altamirano, Veronica Prado, Juan Caballeria<br />

1308<br />

Ethanol-impairs mTOR activity by a novel non-canonical<br />

PI3Kγ-PP2A axis activating skeletal muscle autophagy<br />

that underlies sarcopenia.<br />

Gangarao Davuluri 2 , Samjhana Thapaliya 2 , Avinash Kumar 2 ,<br />

Megan R. McMullen 2 , Rebecca L. McCullough 2 , Laura E. Nagy 2 ,<br />

Sathyamangla V. Naga Prasad 3 , Srinivasan Dasarathy 1 ; 1 Department<br />

Of Gastroenterology and Hepatology, Cleveland Clinic,<br />

Cleveland, OH; 2 Pathobiology, Cleveland Clinic, Cleveland, OH;<br />

3 Molecular Cardiology, Cleveland Clinic, Cleveland, OH<br />

Background. Sarcopenia or loss of skeletal muscle mass in<br />

alcoholic cirrhosis is frequent and contributes to morbidity and<br />

mortality. No therapeutic interventions to reverse sarcopenia<br />

in alcoholic cirrhosis exist because the mechanisms are poorly<br />

understood. We have previously observed an alcohol mediated,<br />

AMPK independent, PI3Kγ dependent inhibition of mTOR<br />

signaling that activates skeletal muscle autophagy. In the present<br />

<strong>studies</strong> we show that ethanol-induced mitochondrial reactive<br />

oxygen species impairs mTOR activity by a novel non-canonical<br />

PI3Kγ-protein phosphatase 2 A (PP2A) axis that is responsible<br />

for increased skeletal muscle autophagy that underlies<br />

sarcopenia. Methods. Skeletal muscle mass in patients and<br />

controls and myotube diameter were quantified by image analysis.<br />

Phosphorylation of AMPK, mTOR and its activity was measured<br />

by Immunoblots of protein from C2C12 murine myotubes<br />

exposed to 100 mM ethanol for 6 hr and skeletal muscle from<br />

mice fed ethanol or pair fed for 25 days. Thin layer chromatography<br />

blots for PI3Kγ activity was quantified in C2C12 myotubes<br />

exposed to ethanol and gastrocnemius muscle from mice<br />

fed ethanol or pair fed controls. Results. Here we report that<br />

ethanol increases skeletal muscle autophagy in humans, mouse<br />

models and murine myotubes with unaltered or reduced proteasome<br />

activity and contributes to sarcopenia. The two major<br />

regulators of autophagy are mTOR that inhibits autophagy and<br />

AMPK that activates autophagy. Surprisingly, our data showed<br />

that both mTOR signaling (autophagy inhibitor) and AMPK<br />

activation (autophagy activation) were decreased in ethanol<br />

treated myotubes and in mice exposed to ethanol. The surprising<br />

signaling response via 2 pathways that converge on autophagy,<br />

suggested that dephosphorylation may be a potential<br />

mechanism for reduced phosphorylation of both mTOR and<br />

AMPK. Consistently, we observed that skeletal muscle PP2A<br />

activity was increased and was accompanied by a decreased<br />

PI3Kγ activity, an inhibitor of PP2A. These data showed that<br />

ethanol activates PP2A dependent dephosphorylation of mTOR<br />

and AMPK. We confirmed this by using fostriecin, a PP2A<br />

inhibitor, that reversed ethanol-induced inactivation of mTOR<br />

and AMPK. Conclusions. Alcohol inhibits activation of mTOR,<br />

1309<br />

Differential Changes In Pro-Apoptotic Plasma Glycochenodeoxycholate<br />

And Fecal Glycodeoxycholate Characterize<br />

Bile Acid Metabolome In Alcoholic Liver Disease<br />

Puneet Puri 2 , Andrew R. Joyce 1 , Amon Asgharpour 2 , Mohammad<br />

S. Siddiqui 2 , Richard K. Sterling 2 , R. Todd Stravitz 2 , Velimir A.<br />

Luketic 2 , Scott C. Matherly 2 , Kalyani Daita 2 , Susan A. Walker 2 ,<br />

Stephanie Taylor 2 , Christian E. Ammons 2 , Faridoddin Mirshahi 2 ,<br />

Arun J. Sanyal 2 ; 1 Venebio, Richmond, VA; 2 Virginia Commonwealth<br />

University, Richmond, VA<br />

BACKGROUND: Bile acid (BA) derangements are implicated<br />

in the pathophysiology of alcoholic (ALD) and nonalcoholic<br />

fatty liver disease (NAFLD). Two NAFLD phenotypes, fatty liver<br />

(NAFL) and steatohepatitis (NASH) relate to disease severity.<br />

Recently, a BA analogue showed encouraging therapeutic<br />

potential for NASH. We hypothesized that ALD and NAFLD<br />

are associated with dysregulated BA metabolism. AIMS: 1) To<br />

characterize plasma and fecal BA metabolome compared to<br />

controls (CTL), and 2) To evaluate BA metabolites as potential<br />

biomarkers of ALD, NAFL and NASH. METHODS: Controls,<br />

biopsy confirmed NAFL and NASH, and ALD subjects were<br />

enrolled. Pertinent clinical data, plasma and stool samples<br />

were collected. High throughput mass spectrometry was used<br />

for metabolomics. Data were analyzed using pairwise and multiple<br />

group comparisons by ANOVA, regularized Hotelling’s<br />

T 2 , and partial least square discriminant analysis with variable<br />

importance projection (VIP) scores. A VIP score >1 is significant.<br />

Higher VIP score indicates importance of that variable contributing<br />

to the change. RESULTS: Four groups with mean(±SD)<br />

age (yrs): CTL (n=24, 39±12), NAFL (n=22, 53±10), NASH<br />

(n=37, 58±9) and ALD (n=18, 50±11) were studied. Plasma<br />

BA metabolome: ALD subjects had markedly increased total<br />

primary (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 855A<br />

also offer potential for novel biomarkers and therapeutic targets.<br />

Disclosures:<br />

Puneet Puri - Advisory Committees or Review Panels: Health Diagnostic Laboratory<br />

Inc.; Consulting: NPS Pharmaceuticals Inc.<br />

Andrew R. Joyce - Independent Contractor: Venebio Group, LLC, algorithmRx,<br />

LLC; Management Position: Venebio Group, LLC, algorithmRx, LLC<br />

Richard K. Sterling - Advisory Committees or Review Panels: Merck, Vertex, Salix,<br />

Bayer, BMS, Abbott, Gilead, Baxter, Jansen; Grant/Research Support: Merck,<br />

Roche/Genentech, Pfizer, Gilead, Boehringer Ingelheim, Bayer, BMS, Abbott<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: Amon Asgharpour, Mohammad<br />

S. Siddiqui, R. Todd Stravitz, Velimir A. Luketic, Scott C. Matherly, Kalyani Daita,<br />

Susan A. Walker, Stephanie Taylor, Christian E. Ammons, Faridoddin Mirshahi<br />

1310<br />

Ethanol Exposure Inhibits Hepatocellular Lipophagy by<br />

Inactivating the Small Regulatory GTPase Rab7<br />

Ryan J. Schulze 1 , Shaun Weller 1 , Micah B. Schott 1 , Karuna<br />

Rasineni 2 , Barbara Schroeder 1 , Carol A. Casey 2 , Mark A.<br />

McNiven 1 ; 1 Biochemistry and Molecular Biology, Mayo Clinic,<br />

Rochester, MN; 2 University of Nebraska Medical Center, Omaha,<br />

NE<br />

Alcohol consumption is a well-established risk factor for the<br />

onset and progression of fatty liver disease, in which an estimated<br />

90% of heavy drinkers are thought to develop significant<br />

liver steatosis. For these reasons, an increased understanding<br />

of the molecular basis for alcohol-induced hepatic steatosis<br />

is urgently required. Recently, it has become clear that autophagy,<br />

a catabolic process of intracellular degradation and<br />

recycling, plays a key role in hepatic lipid metabolism. We<br />

have previously shown that the small GTPase Rab7, known<br />

to regulate membrane trafficking and fusion events in the late<br />

endocytic pathway, also acts as a central orchestrator of hepatocellular<br />

lipophagy, a selective form of autophagy in which<br />

lipid droplets (LDs) are specifically targeted for turnover by<br />

the autophagic machinery (Schroeder et al., 2015 and article<br />

highlight by Carmona-Gutierrez et al., 2015). Nutrient starvation<br />

resulted in Rab7 activation on the surface of the LD as<br />

well as on degradative compartments such as the lysosome.<br />

This activation resulted in the mobilization of triglycerides<br />

stored within the LDs for energy production. The GOAL of<br />

this study was to test whether the steatotic effects of alcohol<br />

exposure are a result of perturbations in the Rab7-mediated<br />

lipophagic pathway. Rats fed a 6-week Lieber-DeCarli ethanol-containing<br />

diet accumulated a significantly higher amount<br />

of fat in their hepatocytes. Interestingly, primary hepatocytes<br />

isolated from either these ethanol-fed rats or cultured VA-13<br />

hepatocytes (HepG2 cells expressing alcohol dehydrogenase)<br />

contained enlarged and juxtanuclear-localized multivesicular<br />

bodies (MVBs) and lysosomes that exhibited impaired motility<br />

and fewer productive interactions with LDs. Importantly, similar<br />

morphological changes in these organelles are observed<br />

in hepatocytes subjected to an siRNA-mediated knockdown<br />

of Rab7. Consistent with these defects in the MVB and lysosomal<br />

compartments, we observed a marked 80% reduction<br />

in Rab7 activity in cultured hepatocytes as well as a complete<br />

block in starvation-induced Rab7 activation in primary hepatocytes<br />

isolated from chronic ethanol-fed animals. In summary,<br />

these findings support a direct mechanism whereby acute or<br />

chronic ethanol exposure inhibits Rab7 activity, resulting in<br />

the impaired transport, targeting, and fusion of the autophagic<br />

machinery with LDs, leading to an accumulation of hepatocellular<br />

lipids and hepatic steatosis. This study was generously supported<br />

by funding from the NIDDK [5R37DK044650 (MAM)],<br />

the NIAAA [5R01AA020735 (MAM and CAC)], the Mayo<br />

Clinic Center for Cell Signaling in Gastroenterology, and the<br />

Robert and Arlene Kogod Center on Aging.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ryan J. Schulze, Shaun Weller,<br />

Micah B. Schott, Karuna Rasineni, Barbara Schroeder, Carol A. Casey, Mark<br />

A. McNiven<br />

1311<br />

Alcohol modifies the composition of the intrahepatic<br />

macrophage pool through FOXO3-dependent and cell<br />

type specific apoptosis<br />

Zhuan Li, Jie Zhao, Steven A. Weinman; Department of Internal<br />

Medicine, University of Kansas Medical Center, Kansas City, KS<br />

Hepatic macrophages are a heterogeneous population consisting<br />

of both resident and infiltrating macrophages that carry<br />

out diverse pro- and anti-inflammatory functions necessary<br />

for homeostasis, disease progression and injury repair. The<br />

mechanisms that regulate macrophage subsets in alcoholic<br />

liver disease are not yet understand. We recently showed that<br />

FOXO3 protects the liver from alcohol-induced inflammation.<br />

It also induces S574 phosphorylation of FOXO3 which selectively<br />

enhances expression of pro-apoptotic genes and induces<br />

apoptosis. FOXO3-dependent apoptosis was induced in macrophages<br />

by LPS. The AIMS of this study were to determine<br />

whether FOXO3-dependent macrophage apoptosis protects<br />

the liver from alcohol through changes in intrahepatic macrophage<br />

populations. METHODS: ChIP assays assessed promoter<br />

binding. qPCR was used to measure gene expression. Mouse<br />

peritoneal macrophages were treated with LPS for 24 hours or<br />

differentiated with INF-γ or IL-4 before LPS treatment. Mice were<br />

gavaged once with alcohol or fed a Lieber-DiCarli alcohol diet<br />

for 3 wks. RESULTS: LPS treatment of THP-1 monocytes induced<br />

FOXO3 S574 phosphorylation, enhanced expression of TRAIL,<br />

suppressed Bcl-2 and induced apoptosis. A similar result was<br />

obtained with peritoneal macrophages from wild-type (wt)<br />

mice but none of these effects occurred in FOXO3 -/- macrophages.<br />

To determine if apoptosis was selective for macrophage<br />

type, we separately measured LPS-induced apoptosis in<br />

INF-γ differentiated macrophages (M1) and IL-4 differentiated<br />

macrophages (M2). Only M1 macrophages underwent apoptosis<br />

and this was FOXO3 dependent. Acute alcohol gavage<br />

resulted in nearly a 50% decrease of hepatic macrophage<br />

number by 9 hours in wt but not FOXO3 -/- mice and this loss<br />

of macrophages was selective for cells with M1 markers.<br />

Unlike acute gavage, chronic ethanol feeding had no effect<br />

on hepatic macrophage number or M1/M2 ratio in wt mice<br />

but increased the M1/M2 ratio by 50% in FOXO3 -/- mice.<br />

In the FOXO3 -/- mice there was a strong positive correlation<br />

between M1 pro-inframammary macrophage number and ALT<br />

(P


856A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Zhuan Li, Jie Zhao<br />

1312<br />

PDE4 inhibition attenuates alcohol induced hepatic oxidative<br />

stress by increasing antioxidant enzyme expression<br />

Diana Avila 1 , Jingwen Zhang 1 , Craig J. McClain 1,2 , Shirish<br />

Barve 1 , Leila Gobejishvili 1 ; 1 Department of Medicine/GI, University<br />

of Louisville, Louisville, KY; 2 Robley Rex VAMC, Louisville, KY<br />

Alcohol metabolism leads to generation of free radicals and<br />

oxidative stress with a resultant formation of lipid peroxidation<br />

products, which contribute to the development of alcoholic<br />

liver disease (ALD). Additionally, alcohol decreases antioxidant<br />

capacity of liver, specifically nuclear factor erythroid 2–<br />

related factor 2 (Nrf2) dependent enzymes. cAMP increasing<br />

agents have been shown to effectively mitigate oxidative stress<br />

both in vivo and in vitro. Our previous work has shown that<br />

alcohol increases the expression of PDE4, cAMP degrading<br />

enzyme, in macrophages and hepatic Kupffer cells. Hence,<br />

we hypothesized that compromised cAMP levels due to alcohol<br />

might contribute to the decreased antioxidant capacity<br />

of liver, and restoring cAMP signaling could prevent alcohol<br />

induced oxidative stress. To test this hypothesis, we employed<br />

two approaches to block degradation of cAMP by phosphodiesterases<br />

(PDEs): genetic (by using PDE4B knockout mice)<br />

and pharmacological inhibition of PDE4 (Rolipram) in a mouse<br />

model of ALD. C57BL/6 and Pde4b knockout (pde4b -/- ) male<br />

mice were pair-fed Lieber-DeCarli liquid diet containing either<br />

alcohol (AF) or isocaloric maltose dextrin (PF) for 4 weeks.<br />

Additional groups of mice were treated with Rolipram at 5<br />

mg/kg, 3 times a week for 4 weeks. Mice were sacrificed<br />

at 1, 2 and 4 weeks after starting alcohol. Liver cAMP levels<br />

were measured by cAMP ELISA kit; Hepatic inflammation and<br />

oxidative stress was examined by immunohistochemical staining<br />

of liver tissue with 4HNE (4-Hydroxynonenal) and F4/80<br />

(monocyte/macrophage marker) antibodies. Effect of alcohol<br />

on activation of Nrf2 was evaluated by examining nuclear<br />

Nrf2 levels. Further, expression of Nrf2-dependent hepatic antioxidant<br />

enzymes, superoxide dismutase 1 and 2 (SOD1/2),<br />

Catalase and glutathione peroxidase (GPx) was analyzed by<br />

Western blot. Alcohol feeding resulted in a significant increase<br />

in hepatic PDE4 expression and a decrease in cAMP levels.<br />

Liver 4HNE and F4/80 staining was significantly increased in<br />

alcohol fed WT mice demonstrating increased oxidative stress,<br />

but was significantly attenuated by PDE4 inhibition. Alcohol<br />

reduced Nrf2 activation in wild type mice but not in Rolipram<br />

treated and PDE4B knockout mice. Catalase and SOD2 levels<br />

did not seem to be affected by alcohol, but SOD1 and GPx1/2<br />

were decreased in alcohol fed wild type mice. Notably, in comparison<br />

with wild type, Pde4b knockout and Rolipram treated<br />

alcohol fed mice had higher levels of SOD1/2, and GPx1/2.<br />

In summary, these data indicate that alcohol effect on hepatic<br />

cAMP levels contribute to compromised antioxidant capacity of<br />

the liver, which could be restored by PDE4 inhibition.<br />

Disclosures:<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

Shirish Barve - Speaking and Teaching: Abbott<br />

The following authors have nothing to disclose: Diana Avila, Jingwen Zhang,<br />

Leila Gobejishvili<br />

1313<br />

Hepatocyte-specific deletion of the essential circadian<br />

clock transcription factor BMAL1 enhances chronic ethanol-induced<br />

depletion of hepatic glycogen in mice<br />

Uduak S. Udoh, Telisha Swain, Shannon Bailey; Pathology, University<br />

of Alabama at Birmingham, Birmingham, AL<br />

The molecular circadian clock plays a central role in regulating<br />

fundamental physiological and metabolic processes and allows<br />

organisms/organs/cells to anticipate and adapt to changes in<br />

their 24 hr environment. Previously, we found that chronic alcohol<br />

consumption disrupts the liver molecular clock mechanism<br />

and daily rhythms in liver glycogen metabolism. Herein, our<br />

goal was to determine the effect chronic alcohol consumption<br />

has on hepatic glycogen metabolism using a model of hepatocyte<br />

clock dysfunction; i.e., the hepatocyte-specific BMAL1<br />

knockout (HBK) mouse. For these <strong>studies</strong>, single-housed male<br />

HBK and wild type (WT) littermates were kept under a 12:12<br />

hr light-dark cycle and fed either the Lieber-DeCarli control<br />

diet or the ethanol-containing (3% w/v) diet using a pair-fed<br />

isocaloric study design for 5 weeks. Livers were collected every<br />

4 hr for a 24 hr period at Zeitgeber time (ZT) 3, 7, 11, 15,<br />

19, and 23 where ZT 0 = lights on (beginning of less active<br />

phase) and ZT 12 = lights off (beginning of more active phase),<br />

and used for glycogen determination and gene expression.<br />

Chronic alcohol consumption significantly decreased glycogen<br />

content in livers of WT mice at all times of the day. The diurnal<br />

rhythm in glycogen was also altered in livers of WT ethanol-fed<br />

mice as compared to the rhythm measured in livers of WT control-fed<br />

mice [Mesor/Mean: WT-Control (40.30) vs. WT-Ethanol<br />

(24.87), p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 857A<br />

inhibition of two ceramide synthetic pathways on hepatic Plin2<br />

expression, steatosis, and glucose and lipid homeostasis in<br />

vivo. In vitro, we investigated the individual effects of de novo<br />

synthetic enzymes on Plin2 regulation. METHODS Mice were<br />

fed a Lieber-DeCarli control or ethanol diet for 4 weeks and<br />

pharmacologically treated i.p. with imipramine (an inhibitor<br />

of sphingomyelinase (SMASE)); myriocin (an inhibitor of serine<br />

palmitoyl transferase (SPT), the rate limiting enzyme of de<br />

novo ceramide synthesis); or saline. Various metabolic parameters,<br />

serum chemistries and glucose tolerance were measured.<br />

Liver samples were taken for immunoblotting, mRNA analysis<br />

and ceramide measurement by mass spectrometry. In vitro, we<br />

incubated VL-17A cells (ethanol-metabolizing human hepatoma<br />

cells) with control or ethanol media and inhibited key enzymes<br />

in the de novo ceramide synthetic pathway with myriocin,<br />

fumonisin B1 (FB-1, an inhibitor of ceramide synthase), or<br />

GT-11 (an inhibitor of dihydroceramide desaturase). Plin2<br />

mRNA and protein, lipids, and markers of translation were<br />

quantified. RESULTS SPT and SMASE inhibition in alcohol-fed<br />

mice reduced hepatic triglycerides (TG) by approximately 37%<br />

(p


858A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

BMI 29, 44% coffee users, 13% black tea users, 4% green tea<br />

users, mean ALT 26, mean AST 27, mean bilirubin 0.5 mg/<br />

dL, mean albumin 3.8 mg/dL) were recruited. Apart from liver<br />

tests, cases and controls differed for age (P=0.046), education<br />

level (P=0.03), coffee use (P=0.0002). On a logistic regression<br />

analysis model after controlling for relevant co-variates,<br />

white race was associated with AH with OR (95% CI) of 5.5<br />

(1.8-16.5). In contrast, odds of AH were lower by about 80%<br />

among coffee users; 0.21 (0.10-0.45). Recidivism was self-reported<br />

by 41 (16 cases) and 14 (6 cases) at 6 and 12 months<br />

respectively. On a logistic regression analysis model, low (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 859A<br />

intestinal crypts that are enriched in Paneth cells resulted in a<br />

dose- and time-dependent increase in IL-17 protein production.<br />

Conclusions: Our results highlight substantial hepatic immune<br />

cell changes due to alcohol and indicate that increased number<br />

of Paneth cells in the small intestine is an early event in alcohol-induced<br />

intestinal pathology. For the first time, we describe<br />

that IL-17 production by Paneth cells is induced by alcohol in<br />

small intestinal crypts and represents an early intestinal inflammatory<br />

marker in ALD. These observations identify a novel role<br />

of Paneth cells and intestinal IL-17 in ALD and highlight the<br />

importance of the gut-liver axis in alcoholic liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Benedek Gyongyosi, Patrick<br />

Lowe, Arvin Iracheta-Vellve, Abhishek Satishchandran, Aditya Ambade, Banishree<br />

Saha, Gyongyi Szabo<br />

1319<br />

Microbiota protects mice against acute alcohol-induced<br />

liver injury<br />

Peng Chen, Yukiko Miyamoto, Magdalena Mazagova, Lars Eckmann,<br />

Bernd Schnabl; University of California San Diego, La Jolla,<br />

CA<br />

Background: Chronic alcohol abuse is associated with intestinal<br />

bacterial overgrowth, increased intestinal permeability,<br />

and translocation of microbial products from the intestine to<br />

the portal circulation and liver. There is strong evidence that<br />

experimental alcoholic liver disease depends on translocated<br />

microbial products. The aim of our study was to investigate<br />

the physiological relevance of the intestinal microbiota in<br />

alcohol-induced liver injury using conventional and germ-free<br />

C57BL/6 mice. Methods: We employed germ-free mice in a<br />

model of acute alcohol exposure that mimics binge drinking.<br />

Acute alcoholic liver injury was induced by a single gavage of<br />

300 μl of 30% (vol/vol) ethanol. Control mice were gavaged<br />

with an isocaloric amount of dextrose (300 μl, 0.42 g/ml).<br />

Mice were harvested 9 hours after gavage. Results: Germ-free<br />

mice showed significantly greater liver injury as determined<br />

by plasma ALT levels and H&E staining of liver sections after<br />

oral gavage of ethanol compared with conventional mice.<br />

This was accompanied by increased hepatic expression of<br />

macrophage and neutrophil markers. Several proinflammatory<br />

cytokines and chemokines including TNFα, Ccl2, Ccl4, Ccl5,<br />

Cxcl1, Cxcl2 and Cxcl5 were increased in alcohol-binged<br />

germ-free relative to conventional mice. In parallel, germ-free<br />

mice exhibited increased hepatic steatosis and upregulated<br />

expression of genes involved in fatty acid and triglyceride synthesis<br />

compared with conventional mice after acute ethanol<br />

administration. The absence of microbiota was also associated<br />

with increased hepatic expression of ethanol metabolizing<br />

enzymes, which led to faster ethanol elimination from the<br />

blood and lower plasma ethanol concentrations. In particular,<br />

hepatic Cyp2e1 gene and protein expression were significantly<br />

higher in germ-free mice before and after an alcohol<br />

binge. Furthermore, expression of aldehyde dehydrogenase 2<br />

(Aldh2), which catalyzes the transformation of acetaldehyde<br />

to acetic acid, was significantly higher in control and alcohol<br />

challenged germ-free mice. Intestinal levels of ethanol metabolizing<br />

genes showed regional expression differences, and<br />

were overall higher in germ-free relative to conventional mice.<br />

In conclusion, our findings indicate that absence of the intestinal<br />

microbiota increases hepatic ethanol metabolism and the<br />

susceptibility to binge-like alcohol drinking. The commensal<br />

microbiota has therefore a beneficial role in protection against<br />

acute alcoholic liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Peng Chen, Yukiko Miyamoto,<br />

Magdalena Mazagova, Lars Eckmann, Bernd Schnabl<br />

1320<br />

Hepatocyte-specific StARD1 deletion attenuates alcoholic<br />

steatohepatitis<br />

Vicent Ribas 1,2 , Susana Nuñez 2,1 , Carmen Garcia-Ruiz 2,1 , Jose<br />

Fernandez-Checa 2,1 ; 1 Liver Unit, Hospital Clínic i Provincial de<br />

Barcelona, IDIBAPS, Barcelona, Spain; 2 Instituto de Investigaciones<br />

Biomédicas de Barcelona (IIBB-CSIC)), Barcelona, Spain<br />

Previous <strong>studies</strong> have shown that chronic alcohol feeding<br />

results in the depletion of mitochondrial GSH due to the accumulation<br />

of cholesterol in mitochondria. Mitochondrial GSH<br />

depletion stimulates reactive oxygen species and oxidative<br />

stress, and sensitizes hepatocytes to hypoxia and inflammatory<br />

cytokines-mediated apoptosis. Steroidogenic acute regulatory<br />

domain 1 (StARD1) protein was identified in steroidogenic<br />

tissues as a key protein responsible for cholesterol trafficking<br />

to the mitochondrial inner membrane for the synthesis of steroids.<br />

Mice with global StARD1 deletion die within 10 days<br />

due to adrenal lipoid hyperplasia. To circumvent this limitation,<br />

we generated mice with hepatocyte-specific deletion of<br />

StARD1 to study its role in alcoholic steatohepatitis induced<br />

by a novel model in which the Lieber DeCarli ethanol liquid<br />

diet was supplemented with 0.2% cholesterol. METHODS: We<br />

generated mice with floxP sites flanking exons 2-5 of StARD1<br />

(floxed mice) and were crossed with Albumin-Cre transgenic<br />

mice. StARD1 liver-specific knockout animals (SLKO) and their<br />

corresponding StARD1 floxed control littermates were fed a<br />

Lieber DeCarli ethanol liquid diet supplemented with 0.2% cholesterol<br />

(L-C+Ch) for four weeks. At the end of this period, we<br />

determined parameters of liver damage, mitochondrial cholesterol<br />

and GSH homeostasis, markers of endoplasmic reticulum<br />

(ER) stress, inflammation, fibrosis and expression of lipogenic<br />

genes. RESULTS: In agreement with previous <strong>studies</strong>, StARD1<br />

mRNA was induced by L-C+Ch feeding in StARD1 floxed<br />

animals (2-4 fold) while StARD1 expression was effectively<br />

ablated in SLKO liver. In parallel, L-C+Ch diet increased serum<br />

ALT (200-250U/L) in StARD1 floxed animals, an event that<br />

was significantly reduced by 50-60% in SLKO animals. Interestingly,<br />

total liver cholesterol was increased by L-C+Ch treatment<br />

in both genotypes, but mitochondrial cholesterol loading<br />

was increased only in StARD1 floxed animals. In line with<br />

this, SLKO liver mitochondria exhibited increased mGSH levels<br />

after L-C+Ch consumption, when compared to StARD1 floxed<br />

mice. Additionally, L-C+Ch feeding increased expression of ER<br />

stress markers (Chop, Atf6, Pdi), genes involved in lipogenesis<br />

(Srebp1, Srebp2, Dgat2, Fasn), fibrosis (Col1a1, Acta2) and<br />

inflammation (Ccl2, Tnfa, IL-6 and IL-1b) in StARD1 floxed mice<br />

(3-6 fold), whereas the rise in these markers was markedly<br />

attenuated (40-70%) in SLKO mice. CONCLUSIONS: These<br />

results suggest a key role of StARD1-mediated mitochondrial<br />

cholesterol trafficking in the pathogenesis of alcoholic steatohepatitis,<br />

indicating that StARD1 may be a novel target for the<br />

treatment of alcohol induced liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Vicent Ribas, Susana Nuñez,<br />

Carmen Garcia-Ruiz, Jose Fernandez-Checa


860A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1321<br />

Infection in alcoholic hepatitis: a major prognostic indicator<br />

and potential therapeutic target<br />

Richard Parker 1,4 , Fiona Jones 2 , Daniel J. Wheatley 3 , Ian A.<br />

Rowe 1 , Christopher Corbett 3 , Andrew Holt 4 , Stephen Stewart 2 ;<br />

1 Centre for Liver Research, University of Birmingham, Birmingham,<br />

United Kingdom; 2 Mater Misericordiae University Hospital, Dublin,<br />

Ireland; 3 New Cross Hospital, Wolverhampton, United Kingdom;<br />

4 University Hospitals Birmingham NHS Foundation Trust, Birmingham,<br />

United Kingdom<br />

Introduction We sought to detail the incidence, consequences<br />

and causes of infection in a cohort of patients with Alcoholic<br />

Hepatitis (AH). Methods Three centres in the UK and Ireland<br />

gathered data from patients with a clinical diagnosis of alcoholic<br />

hepatitis. Clinical, biochemical and microbiological<br />

records were reviewed retrospectively. Infection was defined<br />

by evidence of consolidation on chest x-ray, positive urinalysis,<br />

ascitic fluid analysis showing 250 neutrophils/mm 3 or<br />

positive microbiological culture. Survival was analysed with<br />

Kaplan-Meier (KM) curves and Cox proportional hazard analysis<br />

to control for confounding factors. Characteristics of patients<br />

with or without evidence of infection were compared with twotailed<br />

student’s t-test. Results In total 215 patients were included<br />

with an median follow-up period of 30 months. Average discriminant<br />

function (DF) was 72.7 and 85% of patients had<br />

severe disease (DF >32). Evidence of infection was seen in 97<br />

patients (45%). Patients with infection had significantly higher<br />

serum urea (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 861A<br />

1323<br />

Ethanol-Induced Alterations In The Intestinal Methionine<br />

Metabolic Pathway Promote Tight Junction Disruption<br />

Dan Feng 1,2 , Paul G. Thomes 3 , Natalia A. Osna 1,2 , Dean J.<br />

Tuma 1,2 , Kusum K. Kharbanda 1,2 ; 1 Research Service, Veterans<br />

Affairs Nebraska-Western Iowa Health Care System, Omaha, NE;<br />

2 Department of Internal Medicine, University of Nebraska Medical<br />

Center, Omaha, NE; 3 Liver Pathobiology Laboratory, Cannon<br />

Research Center, Carolinas Healthcare System, Charlotte, NC<br />

The gut-liver interaction has emerged as a critical component<br />

in the pathogenesis of alcoholic liver disease (ALD). The central<br />

mediator of ALD progression is the gut luminal antigens, especially<br />

endotoxins that translocate to the liver via compromised<br />

gut epithelial tight junctions (TJ). We have previously shown<br />

that ethanol consumption can alter the liver methionine metabolic<br />

pathway to impair several methylation reactions leading<br />

to the generation of many hallmark features of alcoholic liver<br />

injury. This study was undertaken to examine whether alcohol<br />

exposure alters the intestinal methionine metabolic pathway<br />

and to explore whether these alterations play a causal role in<br />

disrupting the gut barrier integrity. Adult male Wistar rats were<br />

pair-fed the Lieber DeCarli control or ethanol diet for 8 weeks.<br />

At the end of the feeding regimen, intestines were removed and<br />

analyzed. We observed the level of methionine synthase (MS,<br />

an enzyme of the methionine metabolic pathway that prevents<br />

buildup of a harmful metabolite, S-adenosylhomocysteine,<br />

SAH) was decreased by 50% in the ileum of ethanol-fed rats<br />

compared to controls. This decrease was accompanied by a<br />

compensatory increase in betaine homocysteine methyltransferase<br />

(BHMT), another enzyme that maintains the level of a vital<br />

methylating agent, S-adenosylmethionine (SAM), and prevents<br />

SAH buildup. However, while the adaptive increase in BHMT<br />

maintained SAM levels in the ileum of the ethanol-fed rat, there<br />

was a significant elevation in SAH levels ultimately lowering<br />

the SAM:SAH ratio compared to controls. All these changes<br />

were similar to those previously reported in the livers of ethanol-fed<br />

rats. Concomitant with the changes in the crucial components<br />

of the methionine metabolic pathway that controls the<br />

cellular methylation potential, we saw the disassembly of TJs<br />

in the ileum of the ethanol-fed rats. There was a disorganized<br />

localization of the TJ proteins, occludin and claudin -1, which<br />

normally seal the intercellular spaces to prevent the paracellular<br />

translocation of luminal antigens into portal circulation. Further<br />

<strong>studies</strong> using an inhibitor of lysine methyltransferase identified<br />

that the methylation reactions catalyzed by this enzyme<br />

were important for maintaining the barrier integrity. We further<br />

observed that occludin is methylated and this post-translational<br />

modification occurred on the lysine and not arginine residue(s).<br />

Taken together, our results indicate that alcohol-induced alterations<br />

in the intestinal methionine metabolic pathway and the<br />

resulting impairments in methylation reactions contribute to the<br />

disruption of the TJs, thereby promoting gut leakiness and progressive<br />

liver injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Dan Feng, Paul G. Thomes,<br />

Natalia A. Osna, Dean J. Tuma, Kusum K. Kharbanda<br />

1324<br />

Role of Pre-mRNA Splicing Regulator SLU7 in Mediating<br />

Detrimental Action of Ethanol in the Liver<br />

Jiayou Wang 1,2 , Xudong Hu 1,3 , Alvin Jogasuria 1 , Kwangwon Lee 1 ,<br />

Jiashin Wu 1 , Min You 1 ; 1 Pharmaceutical Sciences, Northeast Ohio<br />

Medical University (NEOMED), Rootstown, OH; 2 Department of<br />

Anatomy, Guangzhou University of Chinese Medicine, Guangzhou,<br />

China; 3 Department of Biology, Shanghai University of Traditional<br />

Chinese Medicine, Shanghai, China<br />

Precursor mRNA (pre-mRNA) splicing is a critical step in gene<br />

expression that removes intronic sequences from immature<br />

mRNA, producing mature mRNA that can be translated into<br />

protein. Errors in splicing contribute to at least 15% of human<br />

genetic disorders. Aberrant pre-mRNA splicing plays a significant<br />

role in liver diseases, but very little is known about its<br />

role in alcoholic steatosis/steatohepatits. Here, we identified<br />

a new target of ethanol action in the liver, namely, pre-mRNA<br />

splicing regulator SLU7. The effects of ethanol on splicing factor<br />

SLU7 were assessed in cultured hepatocytes and in animal<br />

liver. Feeding mice a modified Lieber-DeCarli ethanol-containing<br />

liquid diet for 4-wks resulted in significant decreases in the<br />

mRNA and protein expression levels of hepatic SLU7. More<br />

importantly, ethanol exposure robustly reduced the amount of<br />

splicing factor SLU7 in the nucleus in cultured hepatocytes and<br />

in mouse liver. Slu7 governs the splicing and/or expression of<br />

multiple genes essential for hepatic lipid metabolism, including<br />

serine/arginine-rich splicing factor 3 (SRSF3) and hepatocyte<br />

nuclear factor 4α (HNF4α). SRSF3 pre-mRNA alternative splicing<br />

generates the full-length isoform lacking exon 4 (Iso1) and<br />

an alternative isoform including exon 4 (Iso2). Ethanol-mediated<br />

disruption of splicing factor SLU7 triggered altered SRSF3<br />

pre-mRNA splicing by inhibiting the mRNAs of total SRSF3,<br />

SRSF3_Iso 1, or SRSF3_ Iso2 and HNF-4α in the livers of ethanol-fed<br />

mice. Remarkably, the alteration of hepatic SLU7-SRSF3<br />

axis by ethanol exposure was closely associated with the development<br />

of excess accumulation of hepatic triglycerides in mice.<br />

Taken together, our findings suggest that splicing regulator<br />

SLU7 may play an essential role in regulating the effects of ethanol<br />

on hepatic lipid metabolism and development of alcoholic<br />

fatty liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jiayou Wang, Xudong Hu, Alvin<br />

Jogasuria, Kwangwon Lee, Jiashin Wu, Min You<br />

1325<br />

Impaired Proteasome Function is Associated with<br />

Alcohol-Induced Hepatic Receptor Interacting Protein<br />

3-Mediated Necroptosis, Steatosis and Liver Injury<br />

Wen-Xing Ding, Shaogui Wang, Hong-Min Ni; Pharmacology,<br />

Toxicology and Therapeutics, The University of Kansas Medical<br />

Center, Kansas City, KS<br />

Both hepatocyte apoptosis and necrosis are hallmarks of alcohol-induced<br />

liver injury. Recent evidence suggests that receptor-interacting<br />

protein kinase (RIP) 3-mediated necroptosis is<br />

important in chronic alcohol feeding-induced liver injury in<br />

mice. However, the mechanisms by which alcohol activates<br />

RIP3-mediated necroptosis are unclear. Using the recently established<br />

chronic alcohol feeding plus binge (Gao-binge) model,<br />

we demonstrated that Gao-binge alcohol treatment increased<br />

RIP3 but not RIP1 protein levels in mouse livers based on both<br />

western blot analysis and immunohistochemical staining for<br />

RIP1 and RIP3. 7-OCl-Nec 1, a recently developed specific<br />

RIP1 inhibitor, failed to protect against Gao-binge alcohol-induced<br />

steatosis and liver injury, suggesting alcohol-induced


862A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

necroptosis is RIP1-independent. In contrast, RIP3 knockout<br />

mice had decreased neutrophil infiltration, serum alanine<br />

amino transferase (ALT) activity and steatosis compared to wild<br />

type mice after Gao-binge alcohol treatment, suggesting that<br />

RIP3 but not RIP1 contributes to alcohol-induced necroptosis<br />

and liver injury. Gao-binge alcohol treatment did not change<br />

hepatic mRNA levels of RIP3 but decreased mRNA levels of<br />

RIP1, suggesting the increase in hepatic RIP3 protein level in<br />

alcohol-treated mouse livers was likely due to post-translational<br />

regulation. In primary cultured mouse and human hepatocytes,<br />

we found that alcohol treatment decreased proteasome activity<br />

and increased the protein level of RIP3. Similarly, Bortezomib,<br />

a proteasome inhibitor, also increased the protein level of RIP3<br />

in cultured hepatocytes. Consistent with these in vitro findings,<br />

we also found that Gao-binge alcohol treatment decreased<br />

the levels of proteasome subunit alpha type-5 (PSMA5) and<br />

proteasome subunit beta type-5 (PSMB5), two key proteasome<br />

subunits that are important for proteasome function. As a result,<br />

Gao-binge treatment also decreased proteasome activity in<br />

the mouse livers. More importantly, we also found decreased<br />

expression of PSMA5 and PSMB5 and increased protein levels<br />

of RIP3 in human alcoholic liver biopsy samples compared<br />

to healthy human livers. In conclusion, results from this study<br />

suggest that impaired hepatic proteasome function by alcohol<br />

exposure may contribute to alcohol-induced steatosis and liver<br />

injury by inducing necroptosis through blocking proteasomal<br />

degradation of RIP3.<br />

Disclosures:<br />

The following authors have nothing to disclose: Wen-Xing Ding, Shaogui Wang,<br />

Hong-Min Ni<br />

1326<br />

Characteristic features of microbiomes in alcoholic<br />

liver disease analyzed by 16S ribosomal RNA gene<br />

pyrosequencing and transcriptional fragment length<br />

polymorphism methods using fecal samples: the impact<br />

of abstinence<br />

Makiko Taniai, Etsuko Hashimoto, Kuniko Yamamoto, Yuichi<br />

Ikarashi, Kazuhisa Kodama, Tomomi Kogiso, Nobuyuki Torii, Katsutoshi<br />

Tokushige; Internal Medicine, Institute of Gastroenterology,<br />

Tokyo Women’s Medical University, Tokyo, Japan<br />

Intestinal bacterial overgrowth and dysbiosis<br />

induced by ethanol ingestion appears to play an important<br />

role in the pathogenesis of alcoholic liver disease (ALD).<br />

Recently, determination of microbiomes by 16S ribosomal RNA<br />

(rRNA) gene pyrosequencing and terminal restriction fragment<br />

length polymorphism (TRFLP) using stool samples has been<br />

widely used in place of conventional culture methods for the<br />

assessment of the diversity of complex bacterial communities<br />

and rapid comparison of the community structure. In this study,<br />

we evaluated the characteristics of intestinal microbiomes in<br />

ALD patients and the impact of abstinence on it using TRFLP<br />

analysis. We investigated samples from 20 patients<br />

clinicopathologically diagnosed as ALD (75% males, 21 to<br />

76 years, including 7 patients with cirrhosis) and 10 healthy<br />

volunteers as controls matched with age and sex. Five mg of<br />

feces were collected for analysis. Principle component analysis<br />

(PCA) and phylogenetic cluster analysis were employed to<br />

assess the comparison about the component of microbiomes<br />

in two groups. In 6 cases who successfully stopped drinking,<br />

we rechecked the stool samples at the point from 3 to 6 weeks<br />

after abstinence. Large differences at microbial phylum,<br />

family, and genus levels were noted between the patients<br />

with ALD and control subjects. The most striking feature about<br />

microbiomes in ALD patients was the variability of microbiomes<br />

were significantly decreased and very simplified compared to<br />

that of control subjects. In ALD patients, the increase of Bacteroides<br />

fragilis group and Clostridium coccoides group and the<br />

decrease of Lactobacillus group were prominent. Interestingly,<br />

in all cases who stopped drinking, the variability of microbiomes<br />

dramatically improved and the components approached<br />

to those of controls. Intestinal microbiomes<br />

in ALD patients showed distinct composition and significant<br />

decrease in the variability of microbiomes compared to that of<br />

control subjects. Those dysbiosis improved after relatively shortterm<br />

abstinence. This findings could lead to the elucidation of<br />

pathogenesis and new therapeutic approach for ALD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Makiko Taniai, Etsuko Hashimoto,<br />

Kuniko Yamamoto, Yuichi Ikarashi, Kazuhisa Kodama, Tomomi Kogiso,<br />

Nobuyuki Torii, Katsutoshi Tokushige<br />

1327<br />

LPS-TLR4 pathway mediates ductular reaction expansion<br />

in alcoholic hepatitis<br />

Gemma Odena 1 , Raluca Dumitru 2 , Jiegen Chen 1 , Jose Altamirano<br />

3,4 , Veronica L. Massey 1 , Hiroshi Matsushita 5 , Daniel Rodrigo-Torres<br />

3 , Oriol Morales-Ibanez 3 , Juan Caballeria 6,3 , Pere<br />

Gines 6,3 , Ekihiro Seki 5 , Pau Sancho-Bru 3 , Ramon Bataller 1,3 ;<br />

1 Division of Gastroenterology and Hepatology, Departments of<br />

Medicine and Nutrition and Bowles Center For Alcohol Studies,<br />

University of North Carolina at Chapel Hill, Chapel Hill, NC;<br />

2 UNC Human Pluripotent Stem Cell Core, Neuroscience Center<br />

and Department of Genetics, University of North Carolina at<br />

Chapel Hill, Chapel Hill, NC; 3 Institut d’Investigacions Biomediques<br />

August Pi i Sunyer (IDIBAPS), CIBER de Enfermedades Hepaticas y<br />

Digestivas (CIBERehd), Barcelona, Spain; 4 Liver Unit-Internal Medicine<br />

Department, Vall d’Hebron University Hospital, Vall d’Hebron<br />

Institut de Recerca, Barcelona, Spain; 5 Division of Gastroenterology,<br />

Department of Medicine, Cedars-Sinai Medical Center, Los<br />

Angeles, CA; 6 Liver Unit, Hospital Clínic, Barcelona, Spain<br />

Alcoholic hepatitis (AH) is the most severe form of alcoholic<br />

liver disease and targeted therapies are urgently needed. We<br />

recently showed that severe AH is characterized by progenitor<br />

cell expansion and inefficient hepatic regeneration. Because<br />

LPS is a major molecular driver in AH and its serum levels<br />

correlate with patient outcome, we investigated whether the<br />

LPS-TLR4 pathway mediates progenitor cell expansion. We first<br />

assessed the potential role of LPS in patients with biopsy-proven<br />

AH (n=28). LPS serum levels correlated with disease severity<br />

(ABIC score; r=0.48, p=0.02) and markers of progenitor cell<br />

expansion (KRT23, r=0.58, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 863A<br />

ulation of LPC markers gene expression (EpCam, KRT7 and<br />

KRT23). Importantly, LPS stimulated gene and protein expression<br />

of LPCs markers. Interestingly, human LPCs also showed<br />

a profibrogenic profile, as indicated by increased expression<br />

of collagen and α-smooth muscle actin (7.3 and 3.2-fold,<br />

respectively; p5 MTs among parents was observed in 219 (21.7%), 305<br />

(30.3%) and 109 (10.8%) cases, respectively. 59 (5.8%) and<br />

261 (25.9%) cases had a history of cirrhosis and alcoholism<br />

among parents, respectively. Alcoholic cirrhotics with FH of<br />

MTs had significantly younger age at diagnosis (42.8±9.3<br />

vs 48.3±10.4 years, p


864A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

high utilization of governmental programs. More effective early<br />

identification, prevention and treatment of ETOH liver disease<br />

are needed.<br />

Disclosures:<br />

Rolland C. Dickson - Advisory Committees or Review Panels: Biotest; Speaking<br />

and Teaching: gilead<br />

The following authors have nothing to disclose: Jennifer Saad, Neil Volk, Shawn<br />

Shah, Joseph Shatzel, Harley Friedman<br />

1330<br />

Gut Bacterial Load Influences Neutrophil Infiltration<br />

And Liver Inflammation In An Acute-On-Chronic Alcohol<br />

Feeding Model In Mice<br />

Patrick Lowe, Benedek Gyongyosi, Abhishek Satishchandran,<br />

Arvin Iracheta-Vellve, Aditya Ambade, Gyongyi Szabo; Medicine,<br />

UMass Medical School, Worcester, MA<br />

Background/Aims: The triggers for alcoholic hepatitis involve<br />

both metabolic and inflammatory danger signals. Previous <strong>studies</strong><br />

showed that gut-derived LPS activates Kupffer cells, but the<br />

effect of the gut microbiome on neutrophil infiltration, a major<br />

factor in alcoholic hepatitis, is not known. In this study, we<br />

investigated the role of gut bacterial load on alcohol-induced<br />

liver inflammation and immune cell infiltration. Methods: 6-8<br />

week-old C57BL/6 female mice received 10 days of 5% ethanol<br />

in liquid diet (Lieber DeCarli) followed by one ethanol<br />

gavage (5g ethanol/kg body weight) or pair-fed diet followed<br />

by sugar gavage. Some mice received an antibiotic cocktail<br />

(ampicillin, neomycin, metronidazole and vancomycin) twice<br />

daily orally five days prior to and during alcohol consumption.<br />

Mice were sacrificed 9 hours after the alcohol gavage.<br />

Results: The alcohol-induced increase in gut bacterial load was<br />

reduced 18-fold in antibiotic-treated mice as measured by bacterial<br />

culture of cecum stool samples. Alcohol feeding increased<br />

liver TNFα (4-fold), MCP-1 (8-fold) and CXCL1 (15-fold) mRNA<br />

compared to control diet and gut bacterial reduction eliminated<br />

these increases. Serum MCP-1 protein was increased in alcohol-fed<br />

mice (1409 pg/mL compared to 818 pg/mL in pairfed)<br />

while antibiotic-treated alcohol-fed animals showed no<br />

elevation in serum MCP-1 (689 pg/mL) compared to pair-fed<br />

animals. Further, we found elevated E-selectin mRNA and MPO<br />

immunohistochemical staining, indicators of neutrophil activation,<br />

in livers of alcohol-fed compared to control mice and<br />

these neutrophil markers were significantly reduced in antibiotic-treated<br />

alcohol-fed mice. This suggests a reduction in liver<br />

neutrophil infiltration and activation upon antibiotic treatment.<br />

Unlike inflammation, liver injury was not dependent on gut<br />

bacterial load as ALT levels were elevated in alcohol-fed mice<br />

compared to pair-fed controls with and without antibiotic treatment<br />

(177 and 190 IU/L, respectively). However, antibiotics<br />

did provide protection from lipid accumulation as alcohol-fed<br />

antibiotic treated mice had decreased liver Oil-Red O staining<br />

compared with non-treated alcohol-fed mice. Conclusions: Our<br />

results suggest that gut bacterial load correlates with the extent<br />

of liver inflammation, neutrophil infiltration, systemic cytokine<br />

increase and steatosis following alcohol consumption, while<br />

alcohol-induced liver damage (increased ALT) is independent<br />

of gut bacterial load. These observations highlight the importance<br />

of gut microbiome in the dynamics of neutrophil infiltration<br />

and indicate a role of gut- and pathogen-derived signals in<br />

neutrophil activation in alcoholic hepatitis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Patrick Lowe, Benedek Gyongyosi,<br />

Abhishek Satishchandran, Arvin Iracheta-Vellve, Aditya Ambade, Gyongyi<br />

Szabo<br />

1331<br />

Tributyrin, a butyrate pro-drug, attenuates alcohol-induced<br />

inflammation and neutrophil-mediated hepatic<br />

injury<br />

Hridgandh Donde 1 , Smita Ghare 1 , Jingwen Zhang 1 , Swati Joshi-<br />

Barve 1 , Craig J. McClain 1,2 , Shirish Barve 1 ; 1 Department of Medicine/GI,<br />

University of Louisville, Louisville, KY; 2 Robley Rex VAMC,<br />

Louisville, KY<br />

Background: Emerging evidence has shown that neutrophils<br />

and hepatic inflammation are major drivers of alcoholic liver<br />

disease (ALD). Neutrophils infiltration is critically involved in<br />

the liver dysfunction and cellular injury in ALD. Hence, attenuation<br />

of these may prove to be beneficial in the treatment of ALD.<br />

Our recent work indicated that chronic alcohol consumption<br />

decreased intestinal short chain fatty acids, particularly butyrate,<br />

which has a major impact on the development of intestinal<br />

barrier dysfunction and ALD. The aim of this study was to investigate<br />

whether oral administration of tributyrin (Tb; a butyrate<br />

prodrug) results in protection against ALD by targeting alcohol<br />

induced inflammation and its effects in the liver. Tb is a triglyceride<br />

that is rapidly absorbed and metabolized to butyrate<br />

and has favorable pharmacokinetics compared with butyrate,<br />

with low toxicity. Methods: In the present study, we used a<br />

mouse model of ALD to examine the effects of Tb (2g/kg) oral<br />

administration on ethanol-induced changes in hepatic inflammation<br />

and injury. 8–10-week old C57BL/6 male mice were<br />

pair-fed the Lieber-DeCarli liquid diet (Research Diets, NJ). containing<br />

alcohol (AF, n =8) or isocaloric maltose dextrin (PF, n<br />

=8) for 2 and 7 weeks. Tb was administered to a sub-group of<br />

alcohol-fed animals by oral gavage for 5 days/week. Serum<br />

and liver tissue samples were analyzed for hepatic inflammation<br />

and injury. Results: At the end of 7 weeks Tb significantly<br />

attenuated the ethanol-induced neutrophil infiltration as shown<br />

by choline esterase staining. Quantification of neutrophil infiltration<br />

by myeloperoxidase (MPO) activity showed similar<br />

results wherein Tb administered mice had a significant drop<br />

in MPO activity as compared to alcohol-fed mice at 7 weeks<br />

but not in 2 weeks. F4/80 staining of liver sections revealed<br />

increased macrophage infiltration in 7 weeks of alcohol feeding<br />

whereas Tb administration markedly blocked macrophage<br />

infiltration. Further, neutrophil chemoattractant chemokines<br />

such as CXCL-2 and CCL-2 were significantly upregulated in<br />

alcohol-fed mice. Importantly, Tb administration significantly<br />

attenuated alcohol-induced induction of these pro-inflammatory<br />

chemokines. Additionally, alcohol-induced TNF-α, a hallmark<br />

pro-inflammatory cytokine was also attenuated by Tb. Notably,<br />

attenuation of neutrophil infiltration and inflammation by Tb<br />

also led to a significant decrease in liver injury as indicated by<br />

serum AST & ALT levels. Conclusion: The present work demonstrates<br />

that Tb prevents ethanol-induced inflammation and neutrophil-mediated<br />

hepatic injury, and has therapeutic potential<br />

in the prevention/treatment of ALD.<br />

Disclosures:<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

Shirish Barve - Speaking and Teaching: Abbott<br />

The following authors have nothing to disclose: Hridgandh Donde, Smita Ghare,<br />

Jingwen Zhang, Swati Joshi-Barve


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 865A<br />

1332<br />

Differential microRNA expression profiles in patients<br />

with alcoholic hepatitis (AH): Preliminary findings from<br />

the Translational Research and Evolving Alcoholic Hepatitis<br />

Treatment (TREAT) consortium<br />

Suthat Liangpunsakul 1 , Naga P. Chalasani 1 , Guanglong Jiang 2 ,<br />

Yunlong Liu 2 , Vijay Shah 3 , Arun J. Sanyal 4 , Svetlana Radaeva 5 ,<br />

David W. Crabb 1 ; 1 Division of Gastroenterology/Hepatology,<br />

Indiana University School of Medicine, Indianapolis, IN; 2 Department<br />

of Medical and Molecular Genetics, Center for Computational<br />

Biology and Bioinformatics, Indiana University, Indianapolis,<br />

IN; 3 Division of Gastroenterology and Hepatology, Department of<br />

Medicine, Mayo Clinic College of Medicine, Rochester, MN; 4 Division<br />

of Gastroenterology and Hepatology, Department of Medicine,<br />

Virginia Commonwealth University, Richmond, VA; 5 NIH/<br />

NIAAA, Rockville, MD<br />

Background: It is poorly understood why only a subset of subjects<br />

who drink excessively will develop AH. microRNAs (miR-<br />

NAs) are major regulators of multiple biological processes.<br />

Alterations in the miRNAs might play a role in the pathogenesis<br />

of AH. Aim: To determine differentially expressed serum miR-<br />

NAs in subjects with AH compared to heavy drinking controls<br />

without liver disease. Methods: AH was diagnosed as those<br />

with (i) alcohol consumption of >40 g/d for women and >60<br />

g/d for men for a minimum of 6 mos and (ii) total bilirubin>2<br />

mg/dL and AST > 50 U/L. Controls were heavy drinkers with<br />

normal hepatic panel and no evidence of liver disease. RNA<br />

was isolated from the serum of 19 AH (mean age 45 yrs, 50%<br />

women, and 90% White) cases and 20 controls (mean age 44<br />

yrs, 50% women, and 70% White) and custom miRNA profiling<br />

was done by microarray analyses (LC Sciences, Houston,<br />

TX). Results: Array analysis consisted of 2,615 probes. 111<br />

miRNAs were > 2 fold differentially expressed between both<br />

groups with false discovery rate < 0.05. Table 1 describes top<br />

miRNAs with increased and decreased expression in AH compared<br />

to heavy drinking controls. Validated targets for these<br />

miRNAs were sought in the literature: Of the up-regulated miR-<br />

NAs, miR-765 targets HNF4, a transcription factor regulating<br />

many genes determining liver differentiation. Of the down-regulated<br />

miRNAs, the following targets have been reported: IL1<br />

beta for miR-376c; mTOR, c-Met (hepatocyte growth factor),<br />

junB, and IkB kinase for miR-199a; TGF beta and IRS1 for<br />

miR-487; Wnt/beta catenin for miR-432b; and multiple innate<br />

immunity pathway genes for miR-146b. Conclusion: We identified<br />

the serum miRNAs that differentially expressed between<br />

heavy drinkers and AH subjects. Several of these miRNAs have<br />

been reported to be involved in regulation of processes known<br />

to be important in the pathogenesis of AH, such as cytokine<br />

signaling, hepatic growth and differentiation, and innate immunity,<br />

suggesting the changes may provide clues to mechanisms<br />

of liver injury in AH, as well as serving as biomarkers.<br />

Table 1: Serum miRNAs differentially expressed between patients<br />

with AH and heavy drinking controls<br />

Disclosures:<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: Suthat Liangpunsakul, Guanglong<br />

Jiang, Yunlong Liu, Vijay Shah, Svetlana Radaeva, David W. Crabb<br />

1333<br />

Oral pentamidine (VLX103) prevents the development<br />

of alcoholic liver disease in mice<br />

Ghulam Ilyas 1 , Enpeng Zhao 1 , Kathryn Tanaka 3 , Francois Ravenelle<br />

2 , Mark J. Czaja 1 ; 1 Medicine, Albert Einstein College of<br />

Medicine, Bronx, NY; 2 Verlyx Pharma Inc, Montreal, QC, Canada;<br />

3 Pathology, Albert Einstein College of Medicine, Bronx, NY<br />

In alcoholic liver disease (ALD) abnormal lipid metabolism and<br />

hepatocyte injury are driven not only by the direct toxic effects<br />

of alcohol and its metabolites, but also by an overactive innate<br />

immune response generated by stimulation from intestinal lipopolysaccharide<br />

(LPS). Activated macrophages generate factors<br />

such as the proinflammatory cytokine tumor necrosis factor<br />

(TNF) which in the setting of alcohol exposure become cytotoxic<br />

to hepatocytes. Previously we have demonstrated that a<br />

novel form of oral pentamidine, VLX103, significantly reduces<br />

liver injury and improves survival from galactosamine/LPS. The<br />

efficacy of VLX103 in this form of toxic liver injury, together<br />

with the known anti-inflammatory and LPS-binding properties<br />

of pentamidine, suggested that VLX103 may be effective<br />

in the treatment of ALD. We therefore tested the hypothesis<br />

that VLX103 would reduce liver injury in the NIAAA mouse<br />

model of ALD. Methods: A modification of the NIAAA model<br />

was employed in C57BL/6 mice that consisted of 10 days of<br />

control or ethanol diet followed by administration of a single<br />

binge dose of maltose or ethanol, respectively. During the diet<br />

feeding mice received vehicle or 75 mg/kg of VLX103 by<br />

daily gavage. Mice were analyzed 9 h after the maltose/<br />

ethanol gavage. Results: Treatment with VLX103 significantly<br />

decreased the degree of ethanol-induced steatosis as indicated<br />

by a 34% reduction in triglyceride accumulation (64.8 vs. 97.7<br />

mg/g liver; P


866A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1334<br />

Inflammatory and apoptotic markers in alcoholic hepatitis<br />

patients: interim analyses<br />

Leila Gobejishvili 1 , Vatsalya Vatsalya 1 , Mohammad K. Mohammad<br />

1 , Shirish Barve 1 , Craig J. McClain 1,2 ; 1 Department of Medicine/GI,<br />

University of Louisville, Louisville, KY; 2 Robley Rex VAMC,<br />

Louisville, KY<br />

This study is part of a NIAAA UO1 consortium to identify novel<br />

biomarkers and unique drug targets for treatments for AH. 17<br />

male and female acute AH patients within the age group of<br />

21-70 yrs. and 10 healthy controls were included in this initial<br />

pilot investigation. Clinical data for demographic, drinking<br />

history and liver injury markers, MELD, Maddrey DF, and CTP<br />

were recorded. Among AH patients, 6 had a MELD score in a<br />

non-severe range (≤ 19) and 11 were in the severe range (≥<br />

20). Milliplex analysis of serum analytes including cytokines,<br />

soluble FasL, Osteopontin (OPN) and TNF-related apoptosis-inducing<br />

ligand (TRAIL) was performed. A major finding of<br />

our study demonstrates significant inverse correlation of TRAIL<br />

expression with severity of AH represented by MELD and neutrophillia.<br />

AH patients often have hepatic polymorphonuclear<br />

leukocyte infiltration and neutrophilia which plays a major role<br />

in determining whether patients develop acute inflammation.<br />

Importantly, resolution of acute inflammation involves neutrophil<br />

apoptosis mediated by TRAIL. In view of this, association<br />

of the decreasing levels of TRAIL with increasing severity of AH<br />

suggests that this could be a significant pathogenic mechanism<br />

that contributes to the development and severity of AH. In comparison,<br />

there was significant direct correlation of sFasL levels<br />

with the severity of AH in our patient population. Fas/FasL<br />

mediated apoptosis plays a major role in hepatocyte death<br />

that occurs in the proinflammatory cytokine milieu of alcoholic<br />

hepatitis. Similar increase in circulating sFasL has been<br />

reported in severe AH patients (Maddrey score ≥32). Further,<br />

as reported previously, AH patients showed significantly higher<br />

serum Osteopontin levels compared to controls and positively<br />

correlated with MELD. Overall, the findings of our initial pilot<br />

study have begun to identify relevant AH-specific pathogenic<br />

determinants that could serve as prognostic biomarkers and<br />

targets for intervention.<br />

Significant inverse correlation between serum TRAIL levels and<br />

MELD in AH patients.<br />

1335<br />

Zinc deficency inactivate hepatic mitochondrial biogensis<br />

pathway in rats chronically fed alcohol: A mechanism<br />

of perturbed hepatic mitochondrial respiratory<br />

complexes and ROS generation<br />

Qian Sun, Wei Zhong, Wenliang Zhang, Zhanxiang Zhou; Nutrition,<br />

University of North Carolina at greensboro, Kannapolis, NC<br />

Clinical <strong>studies</strong> demonstrated that patients with alcoholic liver<br />

disease (ALD) have a decreased hepatic zinc level, however,<br />

the effect of zinc deficiency on the function of hepatic mitochondria<br />

has not been well studied. The present study was<br />

undertaken to determine if zinc deficiency causes mitochondrial<br />

electron transport chain (ETC) defect and if defected ETC<br />

function could increase hepatic oxidative stress in the presence<br />

of fatty acid and acetaldehyde. Wistar rats were pair-fed with<br />

the Lieber-DeCarli control or ethanol diet for 5 month. Chronic<br />

alcohol exposure significantly increased hepatic triacylglycerol,<br />

free fatty acid and 4-hydroxynonenal (4HNE) levels, meanwhile<br />

hepatic mitochondrial 4HNE level was also increased<br />

and mitochondrial zinc level was reduced. In addition, hepatic<br />

mitochondrial respiratory complex I, III, IV and V were significantly<br />

decreased by chronic alcohol exposure, which was in<br />

consistence with decreased hepatic ATP production. Mitochondrial<br />

biogenesis signal transduction pathway was impaired<br />

by chronic alcohol feeding as indicated by decreased AMPK,<br />

PGC1αa, NRF1 and TFAM levels and mitochondrial DNA<br />

expression. In order to define the link between zinc deficiency<br />

and the defect of mitochondrial respiratory complexes, hepG2<br />

cells were treated with TPEN for 6h. The results indicated that<br />

zinc deficiency significantly decreased mitochondrial respiratory<br />

complex I, III, IV expression. In addition, AMPK, PGC1α,<br />

NRF1 and TFAM levels, and mitochondrial DNA expression<br />

were significantly decreased by TPEN treatment. Consequently,<br />

mitochondrial respiratory complex I, III, IV was silenced by<br />

shRNA, respectively. The transfected hepG2 cells all showed<br />

a decrease in mitochondrial membrane potential and an<br />

increase in free radical production after challenged with 500<br />

mM linoleic acid or 200mM acetaldehyde for 6h. These results<br />

suggest that alcohol exposure induced hepatic zinc deficiency<br />

could inactivate mitochondrial biogenesis signal transduction<br />

pathway and decrease mitochondrial DNA production, which,<br />

in turn, decreases mitochondrial complex protein expression.<br />

The defect of mitochondrial respiratory complexes could<br />

worsen alcohol consumption induced free radical production<br />

while metabolizing fatty acid and acetaldehyde.<br />

Disclosures:<br />

The following authors have nothing to disclose: Qian Sun, Wei Zhong, Wenliang<br />

Zhang, Zhanxiang Zhou<br />

Disclosures:<br />

Shirish Barve - Speaking and Teaching: Abbott<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Leila Gobejishvili, Vatsalya Vatsalya,<br />

Mohammad K. Mohammad<br />

1336<br />

Nuclear receptor Rev-Erbα functions as a transcriptional<br />

activator of Chop to control SHP mediated alcoholic<br />

fatty liver<br />

zhihong yang 1 , Hiroyuki Tsuchiya 2 , Sangmin Lee 1 , Yuxia Zhang 3 ,<br />

Li Wang 1 ; 1 PNB, university of Connecticut, Hamden, CT; 2 School<br />

of Medicine, university of Utah, Salt Lake City, UT; 3 Pharmacology,<br />

Toxicology & Therapeutics, University of Kansas Medical Center,<br />

Kansas City, KS<br />

[Purpose] Excessive consumption of alcohol results in development<br />

steatosis and steatohepatitis, which can progress to<br />

cirrhosis and hepatocellular carcinoma. The endoplasmic reticulum<br />

(ER) stress contributes to metabolic disturbances and is<br />

associated with alcoholic liver disease (ALD). This project aims<br />

at identifying new molecular mechanisms in nuclear receptor


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 867A<br />

mediated circadian regulation of alcoholic fatty liver. [Methods]<br />

The ethanol-binge model (NIAAA model) used in the<br />

current study is described previously. Briefly, two months old<br />

C57BL/6 (WT) and nuclear receptor Shp-/- mice were fed<br />

with Lieber-DeCarli diet containing 5% ethanol ad libitum or<br />

pair- fed with the isocaloric control diet for 10 days. On day<br />

11, mice were oral-gavaged with a single dose of 5g of ethanol/kg<br />

of body weight (EtOH) or maltose dextrin (CTRL). After<br />

nine hours, the serum and liver samples were collected every<br />

six hours for twenty-four hours, subjected to mRNA analysis by<br />

qPCR and RNA-seq or protein expression analysis by Western<br />

Blotting. [Results] Liver TG levels, lipid accumulation and glycogen<br />

increased significantly in WT mice after ethanol binge<br />

compared to control group fed isocaloric diet, but decreased<br />

in Shp -/- mice. Surprisingly, Shp-/- mice receiving control diet<br />

exhibited higher basal levels of TG and lipid accumulation.<br />

Srebp1c mRNA and protein expression were also elevated<br />

in Shp -/- mice, but were reduced in EtOH group, as well as<br />

Fasn, one of Srebp1c’s targets. ER stress markers p-eIf2a and<br />

Atf4 protein increased due to alcohol treatment in WT mice<br />

and the increase is abrogated in Shp-/- mice. The transcription<br />

factor C/EBP homologous protein (Chop) is an evidenced<br />

downstream target of ATF4. Unexpectedly, the basal Chop<br />

expression was higher in Shp-/- mice receiving control diet,<br />

Chop was decreased following alcohol binge in both WT and<br />

Shp-/- mice and the expression was in a circadian fashion.<br />

Rev-Erbα functions as a negative-regulator of several clock proteins.<br />

Interestingly, its levels exhibited a similar changes as<br />

Chop. We further demonstrated that Rev-Erbα activated Chop<br />

promoter activity and induced Chop mRNA and protein expression,<br />

which was inhibited by co-expression of Shp. In addition,<br />

knockdown Rev-Erbα by shRNA reduced Srebp1c protein<br />

cleavage. [Conclusions] Our study identified a novel function<br />

of Rev-Erbα to modulate SHP mediated Chop expression and<br />

Srebp1c cleavage, thus providing a new regulatory mechanism<br />

in circadian clock control of alcoholic fatty liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: zhihong yang, Hiroyuki Tsuchiya,<br />

Sangmin Lee, Yuxia Zhang, Li Wang<br />

1337<br />

The Alcoholic Hepatitis Histology Score Performs Poorly<br />

in a Cohort of Severe Alcoholic Hepatitis in the United<br />

States<br />

Gene Y. Im 1 , Aparna Goel 1 , Stephen C. Ward 2 , Yujin Hoshida 1 ,<br />

Thomas D. Schiano 1 , Scott L. Friedman 1 , Swan N. Thung 2 ; 1 Medicine,<br />

Division of Liver Diseases, Icahn School of Medicine at Mount<br />

Sinai, New York, NY; 2 Pathology, Icahn School of Medicine, New<br />

York, NY<br />

Prediction models for alcoholic hepatitis (AH) that are commonly<br />

used only include clinical variables. The Alcoholic Hepatitis<br />

Histology Score (AHHS) is a newly derived and validated<br />

histologic prediction model. Our aim was to evaluate the performance<br />

of the AHHS in a real-world cohort of severe AH<br />

patients in the United States. Methods: Patients hospitalized<br />

at a high volume liver transplantation (LT) center for severe<br />

AH were prospectively identified from 1/2012 to 6/2015 for<br />

our AH database. Patients with severe AH on liver histology<br />

obtained early in their hospitalization and with a Maddrey’s<br />

discriminant function (DF) ≥32 were included. Liver histology<br />

slides were reviewed by expert hepatobiliary pathologists<br />

who were blinded to patient characteristics. Clinical data<br />

were reviewed to calculate DF, Lille, model for end-stage liver<br />

disease (MELD), Glasgow alcoholic hepatitis score (GAHS)<br />

and age, bilirubin, INR, creatinine (ABIC) scores at presentation.<br />

The primary outcome of analysis was 90-day mortality or<br />

LT. Results: Over the 3.5 year study period, 120 consecutive<br />

patients with severe AH were admitted or transferred to our<br />

liver service and prospectively evaluated. Twenty-six patients<br />

had available liver tissue for analysis. The median age was 44<br />

years (IQR 38-53) and two-thirds Caucasian with female predominance.<br />

Median DF, MELD, Lille, GAHS and ABIC scores<br />

were 78 (IQR 58-93), 33 (IQR 26-39), 0.970 (IQR 0.41-<br />

0.99), 7.5 (7-9) and 8.9 (7.8-9.7), respectively, indicative of<br />

a cohort with high risk of mortality. Standard medical care for<br />

advanced liver disease was provided, with less than one-quarter<br />

receiving glucocorticoid-based therapies and half ineligible<br />

due to severe illness. The cumulative mortality or LT rates at<br />

day 30, 60, 90 and 180 were 42%, 46%, 65% and 69%,<br />

respectively. The AHHS model performed poorly in predicting<br />

90-day mortality or LT, with an area under the curve (AUC) of<br />

0.57. The Lille and GAHS models performed well with AUCs<br />

of 0.83 and 0.71, respectively, while MELD and ABIC had<br />

AUCs of 0.56 and 0.55, respectively. Conclusions: The AHHS<br />

model performed poorly in predicting 90-day mortality or LT in<br />

a real-world cohort of severe AH patients in the United States.<br />

As a static histologic score, the AHHS may lack the precision to<br />

capture the dynamic clinical complexity of severe AH.<br />

Disclosures:<br />

Thomas D. Schiano - Advisory Committees or Review Panels: salix, merck, gilead,<br />

pfizer; Grant/Research Support: galectin, massbiologics, biotest<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

The following authors have nothing to disclose: Gene Y. Im, Aparna Goel, Stephen<br />

C. Ward, Yujin Hoshida, Swan N. Thung<br />

1338<br />

Genetic Ablation of MitoNEET Alleviates Alcoholic Steatosis/Steahepatitis<br />

in Mice<br />

Alvin Jogasuria 1 , Werner J. Geldenhuys 1 , Jiayou Wang 1,2 , Xudong<br />

Hu 1,3 , Kwangwon Lee 1 , Prabodh Sadana 1 , Jiashin Wu 1 ,<br />

Min You 1 ; 1 Pharmaceutical Sciences, Northeast Ohio Medical<br />

University (NEOMED), Rootstown, OH; 2 Department of Anatomy,<br />

Guangzhou University of Chinese Medicine, Guangzhou, China;<br />

3 Department of Biology, Shanghai University of Traditional Chinese<br />

Medicine, Shanghai, China<br />

MitoNEET (CDGSH iron-sulfur domain-containing protein 1<br />

or CISD1) is an outer mitochondrial membrane protein that<br />

donates 2Fe-2S clusters to apo-acceptor proteins, and regulates<br />

mitochondrial matrix iron metabolism. Clinically, aberrant<br />

hepatic iron metabolism is frequently occurred in patients<br />

with alcoholic liver disease. It is also reported that, in vitro<br />

cultured hepatocytes, ethanol along with fructose induces a<br />

large increase in the expression of mitoNEET, which primed the<br />

fructose and ethanol exposed hepatocytes for TNFα-induced<br />

necroptosis. Therefore, in the present study, using a mitoNEET<br />

knock out (MitoNEETKO) mouse model, we aimed to investigate<br />

the in vivo functional role of mitoNEET in the development<br />

of alcoholic steatosis/steatohepatitis and explore the underlying<br />

mechanisms. Alcoholic fatty liver injury was achieved<br />

by pair feeding wild-type (WT) and MitoNEETKO mice with<br />

Lieber-DeCarli ethanol-containing diets for 4-wks. Remarkably,<br />

we found that chronically ethanol-fed MitoNEETKO mice were<br />

resistant to ethanol-induced fatty liver injury demonstrating by


868A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

dramatically reduced hepatic triglycerides, decreased hepatic<br />

cholesterol amount and normalized serum liver enzymes compared<br />

to ethanol-fed WT mice. Our <strong>studies</strong> revealed that genetically<br />

ablation of mitoNEET in mice alleviated alcoholic fatty<br />

liver injury by way of turning on multiples signaling pathways,<br />

including stimulated ileal fibroblast growth factor 15 (FGF15)<br />

synthesis, activated a pre-mRNA splicing regulator SLU7, and<br />

suppressed hepatic lipin-1-medaited phosphatidate phosphatase<br />

(PAP) activity. Altogether, our novel findings suggest that<br />

mitoNEET plays a pivotal role in development and progression<br />

of alcoholic fatty liver injury in mice. Hence, Pharmacological<br />

or nutritional modulation of mitoNEET may be beneficial for<br />

the prevention or treatment of human alcoholic steatosis/steatohepatitis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Alvin Jogasuria, Werner J.<br />

Geldenhuys, Jiayou Wang, Xudong Hu, Kwangwon Lee, Prabodh Sadana,<br />

Jiashin Wu, Min You<br />

1339<br />

Cross talk between hepatocytes and macrophages in<br />

Alcoholic Liver Disease: which language do they use?<br />

Veronica Marin 1 , Kyle L. Poulsen 2 , Laura E. Nagy 2,3 , Claudio<br />

Tiribelli 1,4 , Natalia Rosso 1 ; 1 Fondazione Italiana Fegato, Trieste,<br />

Italy; 2 Department of Pathobiology, Cleveland Clinic, Cleveland,<br />

OH; 3 Department of Molecular Medicine, Case Western Reserve<br />

University, Cleveland, OH; 4 Department of Medical Sciences, University<br />

of Trieste, Trieste, Italy<br />

Macrophage migration inhibitory factor (MIF) is a key cytokine<br />

involved several inflammatory diseases that, depending on<br />

the target cell and inammatory context, can engage different<br />

receptors. It has been reported that interaction between MIF<br />

and CD74 could exert hepatoprotective effects. In the present<br />

study, we explored specifically MIF and CD74 expression in<br />

hepatocytes and macrophage in response to ethanol (EtOH)<br />

exposure. Co-culture of hepatocytes (HuH7) and differentiated<br />

macrophages (THP1+100nM PMA) were exposed to 25mM<br />

EtOH for 24h; monocultures of each cell type were used as<br />

controls (CTRL). Gene expression of MIF, CD74 and TNF-α<br />

was analyzed by real time PCR. The amount of MIF and TNF-α<br />

released in the cultured-media was quantified by ELISA. CD74<br />

protein expression was detected with anti-human CD74-FITC<br />

antibody by flow cytometry. In hepatocytes monoculture, EtOH<br />

induces an up-regulation of TNF-α and MIF mRNA expression,<br />

not translated in cytokines’ release. Interestingly, CD74 expression<br />

(gene and protein) was barely detected in these cells.<br />

On the other hand, in macrophages monoculture EtOH did<br />

not induce relevant changes in any of the parameters under<br />

study. When co-cultured, EtOH increased MIF release (with<br />

no changes at mRNA level).both in hepatocytes and macrophages.<br />

Interestingly, co-cultured hepatocytes showed a significant<br />

up-regulation of CD74 mRNA expression compared with<br />

monoculture CTRL and the same trend was also confirmed at<br />

CD74 protein level in both co-cultured cell types. Regarding<br />

TNF-α even if mRNA was dramatically upregulated its release<br />

was unchanged in hepatocytes and significantly decreased in<br />

macrophages. In conclusion, these data clearly demonstrate<br />

that there is a differential cellular response to EtOH when<br />

hepatocytes and macrophages are cultured together. We<br />

hypothesize that EtOH-challenged hepatocytes (in presence<br />

of macrophages) might be the main source of MIF production<br />

and that CD74 play a determinant role in this mechanism. The<br />

reduction in macrophages TNF-α release lead to the speculation<br />

that in this model MIF through CD74 interaction might<br />

exert hepatoprotective effects.<br />

Table 1- Summary of the obtained data both in monoculture and<br />

co-culture system in response to ethanol exposure<br />

a p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 869A<br />

careful cut-offs from HC, was lower than in literature, adipopenia<br />

was more pronounced.<br />

Disclosures:<br />

The following authors have nothing to disclose: Onan Perez-Hernandez, Geraldine<br />

del Carmen Quintero-Platt, Carlos Jorge-Ripper, Camino Maria Fernandez-Rodriguez,<br />

Maria Blanca Monereo Muñoz, Maria Jose Sanchez-Perez,<br />

Emilio Gonzalez-Reimers, Ruben Hernandez-Luis<br />

Data expressed as mean± SD/ median (IQR) # (C vs. DC), *(HC<br />

vs.C), $ (HC vs. DC)<br />

Disclosures:<br />

Puneet Puri - Advisory Committees or Review Panels: Health Diagnostic Laboratory<br />

inc<br />

The following authors have nothing to disclose: Varsha Shasthry, Jaya Benjamin,<br />

Chethan Kalal, Lovkesh Anand, Ankit Bhardwaj, Vanshja Pandit, Ankur Arora, S.<br />

Rajesh, Viniyendra Pamecha, Vikas Jain, Guresh Kumar, Y.K Joshi, Shiv K. Sarin<br />

1341<br />

The role of proinflammatory cytokines, adipokines and<br />

lipid peroxidation in short term mortality in patients<br />

with Severe Alcoholic Hepatitis<br />

Onan Perez-Hernandez 1,2 , Geraldine del Carmen Quintero-Platt 1 ,<br />

Carlos Jorge-Ripper 1,2 , Camino Maria Fernandez-Rodriguez 1,2 ,<br />

Maria Blanca Monereo Muñoz 1 , Maria Jose Sanchez-Perez 1 ,<br />

Emilio Gonzalez-Reimers 1,2 , Ruben Hernandez-Luis 1,2 ; 1 Medicina<br />

Interna, Hospital Universitario de Canarias, San Cristóbal de La<br />

Laguna, Spain; 2 Medicina Interna, Universidad de La Laguna, San<br />

Cristobal de La Laguna, Spain<br />

INTRODUCTION Acute alcoholic hepatitis is a serious disease<br />

associated with a high short term mortality ranging from 20 to<br />

50%. The factors involved in mortality in patients with acute<br />

alcoholic hepatitis are still controversial. It is well known that<br />

acute alcoholic hepatitis is a disease in which cytokine activation<br />

and oxidative damage play important roles. The aim of<br />

this study is to prospectively assess the relationship between<br />

cytokines, lipid peroxidation, and adipokines with mortality<br />

of patients with alcoholic hepatitis at 30, 90, and 180 days.<br />

PATIENTS AND METHODS Sixty patients with severe alcoholic<br />

hepatitis (Maddrey index ≥32) were successively included.<br />

Cytokine levels (IL-4, IL-6, IL-8, TNF-α, INF-γ, adiponectin,<br />

resistin, leptin, insulin and MDA) were measured within the first<br />

48 hours after admission and one week later. Patients were followed-up<br />

during 180 days; 12 patients died within the first 30<br />

days, 21 within 90 days, and 23 within 180 days. RESULTS<br />

The mean age was 51.9 years (SD=±10.7). Serum levels at<br />

admission of TNF-α were significantly higher among those<br />

who died at admission (p=0.041). IL-4 were significantly lower<br />

among those who died within 30 days (p=0.048) whereas<br />

levels of IL-8 were associated with an increased mortality at<br />

90 and 180 days (p=0.022 and 0.027, respectively). Serum<br />

MDA levels at admission were significantly associated with<br />

mortality at admission, 30, 90, and 180 days. Serum leptin<br />

levels were also associated with mortality at 180 days and<br />

insulin levels were related to mortality at admission, 30, 90,<br />

and 180 days. CONCLUSION Twenty percent of patients died<br />

within the first 30 days after admission; 35.5% died within 90<br />

days, and 38.7% died within 180 days. Increased proinflammatory<br />

cytokine levels at admission and especially increased<br />

MDA levels are significantly associated with short-term mortality<br />

among alcoholics. Lower IL-4 levels were also associated<br />

with mortality. No relation was observed between mortality<br />

and adiponectin levels but indeed between mortality and leptin<br />

levels. Higher mortality was observed among those with low<br />

insulin levels.<br />

1342<br />

Nosocomial Infection in Severe Alcoholic Hepatitis<br />

Onan Perez-Hernandez 1,2 , Geraldine del Carmen Quintero-Platt 1 ,<br />

Emilio Gonzalez-Reimers 1,2 , Maria Jose Sanchez-Perez 1 , Pedro<br />

Abreu-Gonzalez 3 , Maria Jose de la Vega-Prieto 4 , Carlos Jorge-Ripper<br />

1,2 , Francisco Santolaria-Fernandez 1,2 ; 1 Medicina Interna,<br />

Hospital Universitario de Canarias, San Cristóbal de La Laguna,<br />

Spain; 2 Medicina Interna, Universidad de la Laguna, San Cristobal<br />

de La Laguna, Spain; 3 Fisiologia, Universidad de La Laguna,<br />

San Cristobal de La Laguna, Spain; 4 Laboratorio Central, Hospital<br />

Universitario de Canarias, San Cristobal de La Laguna, Spain<br />

INTRODUCTION Both alcoholism and liver cirrhosis increases<br />

the risk of infection. In acute alcoholic hepatitis, especially<br />

in severe cases (Maddrey ≥32), intestinal permeability is<br />

increased. This fact triggers an greater risk of bacterial translocation<br />

and, on the one hand, more infections and on the other,<br />

increased lipid peroxidation mediated proinflammatory activation.<br />

OBJECTIVE The aim of this study is to prospectively assess<br />

the nosocomial infections in severe acute alcoholic hepatitis<br />

(sAAH) and its prognostic value. Furthermore, we examine<br />

the relationship between inflammatory cytokines, adipokines<br />

and lipid peroxidation with nosocomial infections. MATERIAL<br />

AND METHOD Sixty-two patients with sAAH were successively<br />

included. Nosocomial infections were registered. Cytokine levels<br />

(IL-4, IL-6, IL-8, TNF-α, INF-γ, adiponectin, resistin, leptin,<br />

insulin and MDA) were measured within the first 48 hours<br />

after admission and one week later. Patients were followed-up<br />

during 180 days. RESULTS The mean age was 52.2 years<br />

(±10.7). Maddrey’s index on admission was 55.6 (±19.8).<br />

29.5% of patients had a infection during hospitalization<br />

(66.7% were pneumonia), which caused 42.9% of deaths. A<br />

higher percentage of patients with nosocomial infection died<br />

at 180 days (58.8 vs 28.9%; p = 0.03). Infection appeared<br />

after 22.8 days (±16.7). There was a trend to hyponatremia<br />

at admission was more frequent (68.8 vs 31.3%; p = 0.064).<br />

After the first week, infected patients had a worsening of bilirubin<br />

(76.5 vs 48.8%; p=0.048), ABIC≥8 (73.3 vs 40.5%;<br />

p=0,032), platelets 7.91; p=0.031). CONCLUSION Nosocomial<br />

infections were a common complication in patients admitted for<br />

sAAH and were the main cause of death. Data exist relating of<br />

develop it as unfavorable evolution in the first week, measured<br />

by ABIC index, and worsening of bilirubin levels. Thrombocytopenia<br />

and hypofibrinogenemia also associated with increased<br />

risk. Infected patients had lower adiponectin levels, both at<br />

admission and first week. Finally, we also observed lipid peroxidation<br />

was higher in this group.<br />

Disclosures:<br />

The following authors have nothing to disclose: Onan Perez-Hernandez, Geraldine<br />

del Carmen Quintero-Platt, Emilio Gonzalez-Reimers, Maria Jose Sanchez-Perez,<br />

Pedro Abreu-Gonzalez, Maria Jose de la Vega-Prieto, Carlos<br />

Jorge-Ripper, Francisco Santolaria-Fernandez


870A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1343<br />

Effects of Alcohol and Abstinence on Liver Stiffness as<br />

measured by Transient Elastography<br />

Izabella Pokorski 1 , Yang Wu 2 , Martin Weltman 2 , Guy D. Eslick 3 ;<br />

1 Centre for Addiction Medicine, Nepean Blue Mountains Local<br />

Health District, Penrith, NSW, Australia; 2 Gastroenterology and<br />

Hepatology, Nepean Hospital, Penrith, NSW, Australia; 3 Surgery,<br />

The University of Sydney, Penrith, NSW, Australia<br />

The purpose of this study is to examine the relationship between<br />

alcohol consumption, monitored abstinence, relapse and liver<br />

stiffness as measured by Transient Elastogprahy (TE) to elucidate<br />

the best time to accurately record liver fibrosis in patients<br />

with Alcoholic Liver Disease (ALD). TE is a potential alternative<br />

to biopsy to measure liver stiffness, an indirect estimation of<br />

liver fibrosis. Patients presenting for voluntary detoxification<br />

of their alcohol dependence at Nepean Hospital were consecutively<br />

approached from June 2014 until May 2015. Liver<br />

stiffness measurements were recorded on admission, discharge<br />

and at least 7 days post discharge. Relapse information after<br />

discharge was recorded. Blood tests as well as BMI, hepatitis<br />

B and C status were collated for sub-analysis. 123 patients<br />

were recruited.13% were excluded due to their body habitus<br />

and early discharge limited second scan data. 76% of participants<br />

were male (53/70) and 24% of participants were female<br />

(17/70). 70 patients had 2 valid scans, with an average mean<br />

reduction of -2.25kPa in liver stiffness from day of admission<br />

to second TE scan (statistically significant P>0.001). 47% of<br />

patients showed a decrease in at least 1 fibrosis stage (statistically<br />

significant P < 0.001).The average length between admission<br />

and discharge scan was 4 days (range 1-31). Patients<br />

were put into 2 groups based on days between TE scans (3<br />

days and under, more than 3 days). The difference in TE score<br />

for patients who were scanned after 3 days and above was<br />

more statistically significant (P < 0.001) compared to patients<br />

scanned in 3 days and under (P =0.004). 19 (27%) patients<br />

attended a third follow-up scan with a mean time of 96 days<br />

from discharge. Difference in measurements was not statistically<br />

significant. Fibrosis severity among patients with ALD not<br />

infected with HCV (n=40) was variable (F0-1 = 55%, F2-F3<br />

= 35%, F4 = 10%). 19% of patients reported having been<br />

diagnosed with HCV with 46% of these patients initially found<br />

to have cirrhosis (14.8kPa – 66.4 kPa). The study suggests that<br />

liver stiffness as measured by TE declines in patients with ALD,<br />

Hepatitis C and Cirrhosis while undergoing monitored abstinence<br />

from alcohol. A decrease in stage of fibrosis is clinically<br />

significant. The small number of patients presenting for a third<br />

TE scan after discharge (n=19) has made it difficult to evaluate<br />

the effect of relapse of alcohol consumption. However, it is<br />

clear that ongoing consumption of alcohol facilitates over-reading<br />

of TE scores. To obtain the most accurate TE reading, a<br />

scan should be performed at least 3-5 days post complete<br />

alcohol abstinence.<br />

Disclosures:<br />

The following authors have nothing to disclose: Izabella Pokorski, Yang Wu,<br />

Martin Weltman, Guy D. Eslick<br />

1344<br />

Microparticles Are Markers Of Immune Cell Stress And<br />

Predict Severity And Outcomes In Patients With Alcoholic<br />

Liver Disease<br />

Sukriti Sukriti 1 , Jaswinder S. Maras 1 , Shvetank Sharma 1 , Madhumita<br />

Premkumar 2 , Sukanta Das 1 , Shabir Hussain 1 , Guresh Kumar 1 ,<br />

Naminita Gogoi 3 , Ashok K. Choudhury 2 , Chinmay Mukhopadhyay<br />

3 , Nirupma Trehanpati 1 , Shiv K. Sarin 1,2 ; 1 Research, Institute<br />

of Liver and Biliary Sciences, New Delhi, India; 2 Hepatology,<br />

Institute of Liver and Biliary Sciences, New Delhi, India; 3 Special<br />

Center for Molecular Medicine, JNU, New Delhi, India<br />

Background and Aims: Microparticles (MPs) are the membrane<br />

bound vesicles released during cellular stress. Disease<br />

specific MP signatures may aid disease progression or treatment<br />

outcomes. Severe alcoholic hepatitis (SAH) has a rapid<br />

progression and poor response to current therapies. Reliable<br />

tools, preferably noninvasive, are needed for early assessment<br />

of progression and treatment response. We investigated the<br />

correlation between MPs released in SAH, steroid responders<br />

(R) and non-responders (NR). Patients and Methods: MPs<br />

were isolated from plasma using differential ultracentrifugation<br />

followed by flow cytometry. MPs were enumerated by addition<br />

of known size beads (0.22,0.44,0.88,1.45um). This size<br />

gate was combined with Annexin V staining and, various cell<br />

specific markers were used. Reference counting beads were<br />

added to determine absolute count of MPs/ml. MPs were determined<br />

at day 0 and 7 in all 40 patients with SAH; untreated<br />

(n=8), steroid responder (R)(n=20), NR (n=12), and healthy<br />

controls (n=20). The MP counts were correlated with MELD<br />

score, Lille score, serum bilirubin and ALT/AST. Results: MPs<br />

associated with HSCs (CD34+), macs (CD68+) and Tcells<br />

(CD3+CD8+) were higher in NR at day0 (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 871A<br />

1345<br />

An alarmin cytokine IL-33 inhibits type I NKT cells and<br />

neutrophil accumulation into liver and mediates attenuation<br />

of alcoholic liver disease<br />

Igor Maricic, Ekihiro Seki, Idania Marrero, Vipin Kumar; Medicine,<br />

University of California, La Jolla, CA<br />

Natural killer T (NKT) cells, comprised of at least two distinct<br />

subsets, type I and type II, recognize different lipid antigens<br />

presented by CD1d molecules and are involved in inflammatory<br />

liver diseases. We have recently found that type I but not<br />

type II NKT cells become activated and mediate liver injury following<br />

chronic plus binge feeding of Lieber-DeCarli liquid diet<br />

in male C57BL/6 mice. Here we have analyzed the hepatic<br />

cytokine gene expression using quantitative PCR, and found<br />

that IL-33 is significantly upregulated during alcoholic liver disease<br />

(ALD). Interestingly elevated levels of IL-33 are dependent<br />

upon the presence of type I NKT cells as no upregulation was<br />

observed in Jα18-/- mice. Additionally, inhibition of type I NKT<br />

cells by sulfatide or retinoids also blunts IL-33 levels. Next we<br />

examined the effect of IL-33 treatment on ALD. Surprisingly i.p.<br />

administration of IL-33 (2 μg/mouse) resulted in the amelioration<br />

of liver injury whereas anti-IL-33 (5 μg/mouse) treatment<br />

leads to exacerbation of liver enzymes and hepatosteatosis<br />

following alcohol ingestion. Notably IL-33 inhibits the proliferation<br />

and effector function of type I NKT cells in both in vitro<br />

and in vivo assays. Immunohistological analysis of liver sections<br />

showed that the IL-33 administration leads to a significant<br />

inhibition of Ly6G+ neutrophils accumulation into liver.<br />

In contrast anti-IL-33 treatment results in a significant decrease<br />

in neutrophils and an increase in F4/80+ macrophage populations.<br />

These <strong>studies</strong> indicate that in this model of ALD, innate<br />

cytokines such as IL-33 secreted by liver cells have a protective<br />

role from injury and is regulated by interactions with type I NKT<br />

cells and myeloid cells. Collectively these results suggest that<br />

complex interactions of different innate cells play a central role<br />

in mediating inflammatory liver disease and may offer novel<br />

therapeutic targets.<br />

Disclosures:<br />

Ekihiro Seki - Consulting: Merck, Tobira; Grant/Research Support: Nippon Zoki<br />

Vipin Kumar - Management Position: GRI bio<br />

The following authors have nothing to disclose: Igor Maricic, Idania Marrero<br />

1346<br />

Macrophage Migration Inhibitory Factor is Protective in<br />

a Model of Chronic-Binge Ethanol Feeding in Mice<br />

Kyle L. Poulsen 3 , Natalia Rosso 1 , Veronica Marin 1 , Megan<br />

R. McMullen 3 , Adam R. Morris 3 , Claudio Tiribelli 1,2 , Laura E.<br />

Nagy 3,4 ; 1 Fondazione Italiana Fegato-Centro Studi Fegato, Italian<br />

Liver Foundation - Liver Research Center, Trieste, Italy; 2 Clinica<br />

Patologie del Fegato, Universita degli Studi di Trieste, Trieste, Italy;<br />

3 Pathobiology, Center for Liver Disease Research - Cleveland Clinic<br />

Foundation, Cleveland, OH; 4 Gastroenterology, Center for Liver<br />

Disease Research - Cleveland Clinic Foundation, Cleveland, OH<br />

Chronic alcohol abuse is a leading cause of preventable morbidity<br />

and mortality worldwide. Ongoing and continued alcohol<br />

intake by alcohol abusers is the most significant risk factor<br />

associated with the development and progression of Alcoholic<br />

Liver Disease (ALD), which encompasses a spectrum of liver-associated<br />

pathologies. Macrophage Migration Inhibitory Factor<br />

(MIF), a regulator of innate immunity with chemokine- and<br />

cytokine-like activities, was previously identified as a key contributor<br />

to the early stage of ALD after chronic ethanol feeding<br />

to mice. While chronic ethanol feeding in mice results in mild<br />

liver injury and steatosis, the recently developed Chronic-Binge<br />

(Gao-binge) ethanol feeding protocol may better model of the<br />

more severe condition of acute alcoholic hepatitis in humans.<br />

The goal of the current study was to investigate the role of MIF<br />

in liver injury in response to Chronic-Binge ethanol exposure<br />

utilizing female C57Bl/6J and MIF -/- mice. In contrast to the<br />

protection of MIF-/- mice from chronic ethanol feeding, MIF -<br />

/-<br />

mice were not protected from ethanol-mediated liver injury<br />

in response to chronic-binge exposure, exhibiting exacerbated<br />

liver injury as indicated by plasma ALT and AST activities.<br />

Expression of pro-inflammatory cytokines TNFα and IL-1β, as<br />

well as chemokines MCP-1 and MIP2, were enhanced in livers<br />

of MIF -/- mice, further implicating MIF-mediated protection<br />

following Chronic-Binge ethanol feeding. Adhesion molecules<br />

CD62E and ICAM-1 were also enhanced in MIF -/- mice suggesting<br />

that MIF limits leukocyte infiltration into the liver parenchyma.<br />

Flow-cytometric analysis of liver non-parenchymal cells<br />

confirmed that Chronic-Binge ethanol feeding increased infiltration<br />

of CD45 + CD11B + Ly6C + (monocyte/ macrophage) and<br />

CD45 + CD11B + Ly6G + (neutrophil) cells into the liver. While<br />

MIF did not affect recruitment of Ly6C + cells, Ly6G + cells were<br />

increased in MIF -/- mice, consistent with a role for neutrophils<br />

in mediating inflammation in response to chronic binge ethanol<br />

exposure. Taken together, these results suggest that MIF<br />

protects from ethanol-mediated liver damage by limiting pro-inflammatory<br />

cytokine/chemokine synthesis, adhesion marker<br />

expression and neutrophil infiltration following Chronic-Binge<br />

ethanol feeding.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kyle L. Poulsen, Natalia Rosso,<br />

Veronica Marin, Megan R. McMullen, Adam R. Morris, Claudio Tiribelli, Laura<br />

E. Nagy<br />

1347<br />

Creatine Supplementation Does Not Prevent Alcoholic<br />

Steatosis<br />

Dan Feng 1,2 , Ryan W. Barton 2 , Paul G. Thomes 3 , Dean J. Tuma 1,2 ,<br />

Natalia A. Osna 1,2 , Kusum K. Kharbanda 1,2 ; 1 Research Service,<br />

Veterans Affairs Nebraska-Western Iowa Health Care System,<br />

Omaha, NE; 2 Department of Internal Medicine, university<br />

Nebraska Medical Center, Omaha, NE; 3 Liver Pathobiology Laboratory,<br />

Cannon Research Center, Carolinas Healthcare System,<br />

Charlotte, NC<br />

In our ongoing investigation in understanding the pathogenesis<br />

of alcohol induced injury, we have reported that chronic<br />

ethanol intake lowers the hepatocellular S-adenosylmethionine<br />

(SAM) to S-adenosylhomocysteine (SAH) ratio. The reduced<br />

ratio significantly impairs the activities of several SAM-dependent<br />

methyltransferases, leading to the generation of many<br />

hallmark features of alcoholic liver injury (Kharbanda, Sem<br />

Liv Dis, 2009). One such methyltransferase, guanidinoacetate<br />

methyltransferase (GAMT), is severely impaired after ethanol<br />

exposure that results in reduced hepatic creatine production.<br />

This causes detrimental consequences in the liver but also affect<br />

distal organs such as the heart, muscle and brain that depend<br />

on a steady supply of creatine from the liver. Since creatine<br />

supplementation has been recently reported to prevent high-fat<br />

diet-induced hepatic steatosis (Deminice, J Nutr, 2011), we<br />

sought to examine whether creatine supplementation could prevent<br />

alcoholic liver injury. Further, since GAMT is a major consumer<br />

of SAM, we were also interested in examining whether<br />

creatine supplementation could spare this metabolite to preserve<br />

hepatocellular SAM:SAH ratio as well as restore the<br />

depleted hepatic and extrahepatic creatine levels. Adult male<br />

Wistar rats were pair-fed the Lieber DeCarli control or ethanol<br />

diet in the presence or absence of 1% creatine in these respec-


872A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tive diets for 4-5 weeks. At the end of the feeding regimen, the<br />

rats were sacrificed and blood, livers and relevant extrahepatic<br />

tissues were collected and processed for biochemical and histological<br />

analyses. We observed micro- and macro-vesicular<br />

steatosis and a 4 fold-increased triglyceride accumulation in<br />

the livers of rats fed the ethanol diet compared to the controls.<br />

Creatine supplementation neither prevented alcoholic<br />

steatosis nor decreased elevated plasma ALT levels. The lower<br />

hepatocellular SAM:SAH ratio seen in the ethanol-fed rats was<br />

also not normalized when these rats were fed the creatine<br />

supplemented ethanol diet nor were SAM levels increased in<br />

these rats. However, a >10-fold increased level of creatine<br />

was observed in the liver, serum, muscles and hearts of rats fed<br />

the creatine-supplemented control and ethanol diet. Overall,<br />

dietary creatine supplementation did not prevent alcoholic liver<br />

injury despite its known efficacy in preventing high-fat diet-induced<br />

steatosis. To conclude, creatine is ineffective in protecting<br />

against the development of alcoholic steatosis. Betaine, a<br />

pro-methylating agent that maintains hepatocellular SAM:SAH<br />

and prevents alcoholic steatosis still remains our best option for<br />

this treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Dan Feng, Ryan W. Barton, Paul<br />

G. Thomes, Dean J. Tuma, Natalia A. Osna, Kusum K. Kharbanda<br />

1348<br />

The L-carnitine alleviate hepatic fibrosis in a non-alcoholic<br />

steatohepatitis<br />

Hajime Sunagozaka 1 , Masao Honda 1 , Taro Yamashita 1 , Hikari<br />

Okada 1 , Naoki Oishi 1 , Tetsuro Shimakami 1 , Kazuya Kitamura 1 ,<br />

Kuniaki Arai 1 , Yoshio Sakai 1 , Tatsuya Yamashita 1 , Naoto<br />

Nagata 2 , Toshinari Takamura 2 , Eishiro Mizukoshi 1 , Shuichi<br />

Kaneko 1 ; 1 Department of gastroenterology, kanazawa university<br />

Hospital, Kanazawa, Japan; 2 Department of Disease Control and<br />

Homeostasis, Kanazawa University Graduate School of Medical<br />

Sciences, kanazawa Ishikawa, Japan<br />

BACKGROUND & AIMS Non-alcoholic steatohepatitis (NASH)<br />

is a known metabolic disorder of the liver. No treatment has<br />

been conclusively shown conclusively to improve NASH or<br />

prevent disease progression. The function of L-carnitine to<br />

modulates the lipid profile, oxidative stress, and inflammatory<br />

responses. In this study, we evaluated the efficacy of L-carnitine<br />

for the prevention of NASH. METHODS Eight-week-old male<br />

C57BL/6J mice were divided into four groups: (1) basal diet,<br />

(2) atherogenic high-fat (Ath+HF) diet, (3) Ath+HF+0.5% carnitine<br />

diet, and (4) Ath+HF+1% carnitine diet. Liver histology<br />

and gene expression profiles were evaluated at base line and<br />

after 12, 30weeks of L-carnitine administration. Furthermore,<br />

in the clinical research, we administered 1800 mg L-carnitine<br />

for 24 weeks to the patients who have beenwere diagnosed<br />

histologically with histologically NASH for 24 weeks.<br />

We evaluated liver enzymes, lipid and glucose profiles, and<br />

histological scores at the start and end of treatment. Gene and<br />

protein expression were was evaluated by qRT-PCR and western<br />

blotting, respectively. Metabolome analysis was performed<br />

with mice mouse and Hhuman liver tissue samples to detect the<br />

changes of in metabolic path ways with induced by L-carnitine<br />

administration. RESULTS the L-carnitine significantly decreased<br />

liver weight and triglyceride TG and total cholesterol TC levels<br />

compared with the Ath+HF groups in a dose- dependent<br />

manner. Furthermore, the L-carnitine, as compared with the<br />

Ath+ & HF group, was associated with significant reductions<br />

in hepatic steatosis, lobular inflammation, and fibrosis score<br />

at 12 and 30 weeks in a dose- dependent manner. qRT-PCR<br />

analysis demonstrated that L-carnitine significantly reduced<br />

the alteredation of expression of the pro-fibrotic, inflammatory,<br />

and fat synthesis-associated genes, whereas the expression<br />

of PPAR-alpha expression was significantly increased. In<br />

Human researchthe NASH patients, L-carnitine -treated patients<br />

showedadministration significantly improvedments in the liver<br />

enzymes, and histological scores at the end of the treatment<br />

period. In metabolome analysis, the pentose phosphate pathway<br />

was accelerated with L-carnitine, whereas the glycolysis<br />

pathway was not activated directlyory. In this pathway, a<br />

large amount of NADPH was produced and the glutathione<br />

redox ratio was increased in the liver tissue. CONCLUSIONS<br />

L-carnitine supplementation improved hepatic steatosis, inflammation,<br />

fibrosis. L-carnitine could reduce an oxidativeon stress<br />

through the modification of the pentose phosphate pathway<br />

in the NASH liver. L-carnitine could be serve as a candidate<br />

therapeutic strategy for NASH.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Hajime Sunagozaka, Masao<br />

Honda, Taro Yamashita, Naoki Oishi, Tetsuro Shimakami, Kazuya Kitamura,<br />

Kuniaki Arai, Yoshio Sakai, Tatsuya Yamashita, Naoto Nagata, Toshinari<br />

Takamura, Eishiro Mizukoshi<br />

1349<br />

Functional role of the Lin28/let-7 system during alcoholic<br />

liver injury<br />

Kelly McDaniel 2,1 , Tami Annable 2,1 , Yuyan Han 3,2 , Tianhao<br />

Zhou 3,2 , Julie Venter 3,2 , Ying Wan 1,2 , Jessica S. Garner 1 , Nan<br />

Wu 3,2 , Shannon S. Glaser 3,2 , Heather L. Francis 1,2 , Gianfranco<br />

Alpini 3,2 , Fanyin Meng 1,2 ; 1 Scott & White Hospital, Texas A&M<br />

HSC College of Medicine, Temple, TX; 2 Central Texas Veterans<br />

Healthcare System, Temple, TX; 3 Texas A&M HSC College of Medicine,<br />

Temple, TX<br />

Background: Alcoholic liver disease (ALD) is a serious health<br />

concern affecting millions of patients each year. Progression of<br />

alcoholic liver disease (ALD) involves steatosis, inflammation<br />

and subsequent fibrosis leading to cirrhosis of the liver. microR-<br />

NAs including let-7 family have the therapeutic potentials to<br />

recover the liver injury and fibrosis. Lin28 has been shown<br />

to bind to the let-7 pre-microRNA and subsequently block the<br />

maturation of let-7. Our aim is to evaluate the role of lin28/<br />

let-7 system in alcoholic-induced fibrotic liver disease. Methods:<br />

Lin28a conditional mutant mice (Lin28a tm1.2Gqda/J ) and WT<br />

controls underwent chronic and binge ethanol feeding (NIAAA<br />

model). RNA was extracted from whole liver and analyzed<br />

via PCR array and qPCR for fibrosis markers and microRNA.<br />

PCR array data was further analyzed with the GO Enrichment<br />

Analysis tool. Serum from mice was analyzed via ELISA for<br />

TGF-beta levels. Immunohistochemistry (IHC) was performed<br />

on liver sections for fibrosis markers and morphological analysis.<br />

Results: The total liver histopathology score significantly<br />

increased in ethanol WT group relative to control WT mice,<br />

and decreased in Lin28 mutant group relative to WT control<br />

mice with ethanol treatment. By PCR array, ethanol treated<br />

Lin28a mutant mice showed increase in genes associated with<br />

low-density lipoprotein particle receptor, inflammation, vitamin<br />

D synthesis and calidiol 1-monooxygenase activity compared<br />

to Lin28a mutant mice. Lin28a mutant mice treated with ethanol<br />

also showed a decrease in genes associated with primary<br />

miRNA processing and foregut morphogenesis compared to<br />

Lin28a mutant mice. Compared to wild type animals treated<br />

with ethanol, Lin28a mutant mice treated with ethanol showed


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 873A<br />

decreased levels of genes associated with wound healing,<br />

MMP regulation and angiogenesis. Quantitative PCR demonstrated<br />

a significant decrease in alpha-SMA and TIMP3 and a<br />

significant increase in let-7a in Lin28a mutant ethanol treated<br />

mice compared to Lin28a mutant controls. Furthermore, H&E<br />

staining of liver sections showed increased vascularization<br />

and steatosis in all animals treated with ethanol compared<br />

to untreated animals. Significantly reduced Sirius red staining<br />

was observed in ethanol treated Lin28a mutant mice relative<br />

to WT ethanol controls, and in Lin28a mutant mice compared<br />

to wild type animals. Conclusions: Lin28a/let-7 axis regulates<br />

innate inflammation and fibrotic prone processes and diminishes<br />

alcoholic liver injury. Modulation of Lin28a/let-7 system<br />

could be the potential therapeutic approach for alcoholic liver<br />

disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kelly McDaniel, Tami Annable,<br />

Yuyan Han, Tianhao Zhou, Julie Venter, Ying Wan, Jessica S. Garner, Nan Wu,<br />

Shannon S. Glaser, Heather L. Francis, Gianfranco Alpini, Fanyin Meng<br />

1350<br />

Elevated leukocyte count combined with radiologic evidence<br />

of a nodular liver surface in uninfected patients<br />

strongly predicts histologic alcoholic hepatitis, obviating<br />

the need for liver biopsy<br />

Nitzan Roth 1 , Behnam Saberi 2 , Jared Macklin 3 , John Donovan 1 ,<br />

Andrew Stolz 1 , Gary C. Kanel 4 , Neil Kaplowitz 1 ; 1 Division of GI/<br />

Liver Diseases, University of Southern California, Los Angeles, CA;<br />

2 Division of Gastroenterology and Hepatology, The Johns Hopkins<br />

University, Baltimore, MD; 3 Department of Medicine, University of<br />

Southern California, Los Angeles, CA; 4 Department of Pathology,<br />

University of Southern California, Los Angeles, CA<br />

Background: The clinical presentation of alcoholic hepatitis<br />

(AH) can be mimicked by other alcoholic liver diseases. Our<br />

aim was to identify the clinical features that predict having histologic<br />

evidence of AH in order to define a profile that would<br />

obviate the need for liver biopsy. Methods: An expert hepatopathologist<br />

blinded to clinical data retrospectively reviewed<br />

the liver biopsies of 95 active alcoholic patients hospitalized<br />

for presumed severe AH with a discriminant function of >=32<br />

and a bilirubin level of >=5 mg/dL. Patients were defined as<br />

having definite histologic evidence of AH if lobular inflammation<br />

and hepatocyte damage as indicated by moderate to<br />

severe ballooning or easily seen Mallory-Denk bodies were<br />

found. Patients were defined as having no AH if there was<br />

no ballooning or Mallory-Denk bodies and absent or minimal<br />

lobular inflammation. Patients were defined as having possible<br />

AH if they did not meet the histologic criteria for either definite<br />

AH or no AH. Results: Fifty patients (53%) had definite AH and<br />

33 patients (35%) had possible AH. Of the 12 patients (12%)<br />

without AH, five had cirrhosis; the other seven with non-cirrhotic<br />

liver disease had alcoholic foamy degeneration or alcoholic<br />

fatty liver with cholestasis. Thirty-day survival was 94%<br />

for definite AH, 91% for possible AH, and 100% for those with<br />

no AH. Liver surface nodularity on imaging was 98% sensitive<br />

and 38% specific for histologic cirrhosis. In an analysis limited<br />

to uninfected patients with definite AH or no AH, greater leukocyte<br />

count at admission and liver surface nodularity were<br />

independent predictors of definite AH on biopsy (P=14x10 9 /L had<br />

a sensitivity of 43% and a specificity of 100% for diagnosing<br />

definite AH. If a nodular liver surface was present, a leukocyte<br />

count of >=10x10 9 /L had a sensitivity of 80% and a specificity<br />

of 100% for diagnosing definite AH. Using these cut offs,<br />

44% of our cohort met clinical criteria for AH, including 12<br />

of the 33 patients (36%) initially classified as possible AH on<br />

biopsy. Conclusions: The combination of an elevated leukocyte<br />

count and a nodular liver surface in the absence of infection<br />

retrospectively identified patients with a high likelihood of having<br />

histologic AH and suggests that a liver biopsy may not be<br />

necessary in this subset. For patients with suspected severe AH<br />

who do not fulfill these criteria, liver biopsy remains important,<br />

as 23% of these patients have no histologic AH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nitzan Roth, Behnam Saberi,<br />

Jared Macklin, John Donovan, Andrew Stolz, Gary C. Kanel, Neil Kaplowitz<br />

1351<br />

Molecular chaperone Hsp90 regulates cell-specific<br />

induction of epigenetic genes important in macrophage<br />

activation and lipid metabolism during alcoholic liver<br />

injury<br />

Pranoti Mandrekar, Aditya Ambade, Arlene Lim; Medicine, University<br />

of Massachusetts Medical School, Worcester, MA<br />

Background and Aims: Acetylation and methylation of chromatin-modifying<br />

enzymes and epigenetic genes by alcohol<br />

influence transcriptional activity of genes and impact alcoholic<br />

liver disease (ALD). Cellular stress induced molecular chaperones<br />

particularly hsp90 are being recognized as mediators of<br />

environmental impacts on epigenetic states and transgenerational<br />

inheritance. We hypothesized that chronic alcohol exposure<br />

modulates expression of chromatin modifying enzymes<br />

chaperoned by hsp90 during liver injury. Methods: To test<br />

our hypothesis, C57BL/6 mice were subjected to the NIAAA-<br />

Gao chronic-binge alcohol feeding model using Lieber-deCarli<br />

diet with 5% v/v ethanol. A single injection of hsp90 inhibitor,<br />

17-DMAG [17-Dimethylamino-ethylamino-17-demethoxygeldanamycin]<br />

was administered (30-50 mg/kg BW) i.p.<br />

before the binge. Serum ALT, liver cytokine expression and<br />

steatosis were assessed. Hsp90α was estimated in liver nuclear<br />

fractions. Chromatin modifying enzyme expression was analyzed<br />

in whole livers, isolated hepatocytes and Kupffer cells<br />

using PCR arrays. Results: Gene ontology analysis reveals<br />

alterations in histone acetyltransferase, histone methyltransferase,<br />

SET domain proteins and histone deacetylases in the<br />

alcoholic liver. Five genes upregulated include ATF2 (histone<br />

acetyltransferase), PRMT6 and SETD7 (histone methyltransferases),<br />

RPS6KA3 (kinase) and HDAC3 (histone deacetylase)<br />

whereas HDAC9 (histone deacetylase) was downregulated<br />

during ALD. Analysis of expression in isolated hepatocytes and<br />

Kupffer cells exhibits cell type specific regulation of expression.<br />

ATF2 was exclusively upregulated (2 fold) in alcohol exposed<br />

Kupffer cells while PRMT6 (3.2 fold) increased in hepatocytes.<br />

HDAC9, on the other hand, was down regulated (7 fold) in<br />

alcohol exposed hepatocytes. HDAC3, SETD7 and RPS6KA3<br />

were increased in KCs and hepatocytes. Inhibition of hsp90<br />

using 17-DMAG after chronic alcohol exposure significantly<br />

alleviated liver injury measured by reduced serum ALT and liver<br />

triglycerides. 17-DMAG altered expression of ATF2, PRMT6,<br />

HDAC3 and HDAC9 in the liver. Phosphorylation of ATF2 in<br />

LPS stimulated macrophages activates transcription of pro-inflammatory<br />

cytokines. 17-DMAG treatment after alcohol feeding<br />

prevented an increase in ATF2 expression (P


874A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

hsp90 impacts epigenetic chromatin modifying enzymes<br />

important in innate immune activation and lipid metabolism in<br />

a cell-specific manner in ALD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pranoti Mandrekar, Aditya<br />

Ambade, Arlene Lim<br />

1352<br />

Cyr61 protection of hepatocytes from alcohol-induced<br />

endoplasmic reticulum stress<br />

Sang Geon Kim, Tae Hyun Kim, Hyun Joo Kim; College of Pharmacy,<br />

Seoul National University, Seoul, Korea (the Republic of)<br />

The endoplasmic reticulum (ER), the major organelle responsible<br />

for the biosynthesis, folding, assembly and modification<br />

of soluble proteins and membrane proteins, senses cellular or<br />

extracellular stresses. Since the liver plays a key role in energy<br />

metabolism and toxicant detoxification as the portal organ, a<br />

variety of harmful stimuli, including imbalance of energy supply,<br />

alcohol intake, oxidative stress, and inflammatory mediators<br />

induces ER stress, which may facilitate hepatocyte dysfunction<br />

and injury. Cysteine-rich angiogenic protein 61 (Cyr61), a<br />

member of matricellular cysteine-rich protein (CCN) family,<br />

regulates cell survival, differentiation and inflammation. This<br />

study investigated whether Cyr61 inhibits alcoholic steatohepatitis<br />

(ASH), and if so, what the underlying mechanism is using<br />

the database of ASH patients, animal and hepatocyte models.<br />

Of the CCN family members, the transcript levels of CCN1<br />

encoding for Cyr61 were uniquely and markedly diminished<br />

in patients with alcoholic hepatitis. In the liver tissue or hepatocytes<br />

prepared from mice fed on Lieber-DeCarli alcohol diet for<br />

4 weeks or exposed to binge alcohol, Cyr61 expression levels<br />

were substantially decreased with reciprocal increases of fatty<br />

acid synthase and acetyl-CoA carboxylase. Cyr61 knockdown<br />

augmented ER stress in hepatocytes, as indicated by changes<br />

in ER stress markers, CHOP, Grp78, XBP-1 and IRE-1a. Conversely,<br />

overexpression of Cyr61 suppressed ER stress. Moreover,<br />

Cyr61 knockdown promoted PARP cleavage with the<br />

induction of CYP2E1. Cyr61 overexpression diminished ethanol-induced<br />

PARP-1 cleavage. Overall, our results demonstrate<br />

that Cyr61 repression by alcohol intake promotes hepatocyte<br />

injury through increase of ER stress, providing a novel link<br />

between Cyr61 and ER stress in the progression of ASH.<br />

Disclosures:<br />

Sang Geon Kim - Consulting: Nature’s Sunshine Products, Inc<br />

The following authors have nothing to disclose: Tae Hyun Kim, Hyun Joo Kim<br />

1353<br />

IL-13Rα1 signaling and M2 macrophages but not Th2 T<br />

cells drive progression of CCL 4<br />

-induce fibrosis<br />

Shih-Yen Weng 1 , Xiaoyu Wang 1 , Yong Ook Kim 1 , Leonard Kaps 1 ,<br />

Yilang Tang 2 , Olena Molokanova 1 , Jeff R. Crosby 3 , Michael L.<br />

McCaleb 3 , Brombacher Frank 4 , Tobias Bopp 5 , Hans-Joerg Schild 5 ,<br />

Ari Waisman 2 , Detlef Schuppan 1,6 ; 1 Institute of Translational<br />

Immunology, University Medicine, Johannes Gutenberg University,<br />

Mainz, Germany, Mainz, Germany; 2 Institute for Molecular<br />

Medicine, Mainz, Germany; 3 Isis Pharmaceuticals, Carlsbad,<br />

CA; 4 Institute of Infectious Disease and Molecular Medicine, Cape<br />

Town, South Africa; 5 Institute of Immunology and Research Center<br />

for Immunotherapy at UMC Mainz, Mainz, Germany; 6 Beth Israel<br />

Deaconess Medical Center, Boston, MA<br />

Background and aims: Liver fibrosis progression and regression<br />

are modulated by cells of the immune system. CD4 T cells<br />

regulate functional macrophages polarization by secreting<br />

cytokines such as IFNg, IL-12, IL-4 and IL-13. To investigate the<br />

roles of IL-4 and IL-13 in liver fibrosis development, systemic<br />

IL-4/IL-13 double KO mice (IL-4/IL13 -/- ) and T cell specific deletion<br />

of IL-4/IL-13 (IL-4/IL13 ∆CD4 ) and IL-4Rα (IL-4Rα ∆CD4 ) were<br />

generated and subjected to liver fibrosis analysis. Methods:<br />

Liver fibrosis progression was modeled by administration of<br />

oral CCL 4<br />

in increasing doses for 6 weeks. Antisense oligonucleotides<br />

(ASO) were injected intraperitoneally at 40mg/kg 3<br />

times per week during the 6 weeks of CCL 4<br />

treatment . Results:<br />

Histology analysis of IL-4/IL-13 -/- mice at 6 week of CCL 4<br />

treatment<br />

revealed significant reduction of collagen deposition indicated<br />

by less Sirius Red-stained area (50% decrease) in. This<br />

was accompanied by 6-fold reduction of procollagen α1(I)<br />

and α-SMA transcripts, and a decreased ALT indicating suppressed<br />

liver injury. IL-4/IL-13 -/- mice also showed an increase<br />

in CD4 and CD8 T cells, monocytes and macrophages/Kupffer<br />

cells as assessed by FACS. However, IL-4/IL-13 ΔCD4 mice and<br />

IL-4Rα ΔCD4 mice showed no significant difference in liver fibrosis<br />

compared to control mice. Therefore Th2 cells appear to<br />

play only a minor role in CCL 4<br />

-induced fibrosis, suggesting<br />

other fibrogenic cells responding to IL-4/13. IL-13Rα1 was<br />

previously characterized as an M2 macrophage marker. We<br />

therefore applied an ASO against IL-13Rα1 to CCL 4<br />

-treated<br />

mice. Compared to control ASO treated mice, liver fibrosis as<br />

assessed by collagen quantification and Sirius red morphometry,<br />

was significantlyy attenuated by this treatment. Conclusions:<br />

We demonstrate an important role of IL-4/IL-13 signaling<br />

in CCL 4<br />

-induced fibrogenesis. However, IL-4/13 producing<br />

Th2 T cells do not play a significant role in fibrosis progression,<br />

while M2 macrophages, responsive to IL-13, apparently produced<br />

by other sources, are central to fibrosis progression. This<br />

qualifies IL-13 and IL-13Rα1 as important targets for antifibrotic<br />

therapies.<br />

Disclosures:<br />

Jeff R. Crosby - Employment: ISIS Pharmaceuticals<br />

Michael L. McCaleb - Employment: Isis Pharmaceuticals; Stock Shareholder: Isis<br />

Pharmaceuticals<br />

The following authors have nothing to disclose: Shih-Yen Weng, Xiaoyu Wang,<br />

Yong Ook Kim, Leonard Kaps, Yilang Tang, Olena Molokanova, Brombacher<br />

Frank, Tobias Bopp, Hans-Joerg Schild, Ari Waisman, Detlef Schuppan


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 875A<br />

1354<br />

Myofibroblastic conversion of portal fibroblasts and<br />

mesothelial cells isolated from adult mouse liver<br />

Ingrid A. Lua, Kinji Asahina; University of Southern California,<br />

Department of Pathology, Keck Scool of Medicine, Los Angeles,<br />

CA<br />

Purpose: Hepatic stellate cells (HSCs) are believed to be the<br />

major source of myofibroblasts in liver fibrosis; however, other<br />

cellular sources, such as portal fibroblasts (PFs) in the portal<br />

triad and mesothelial cells (MCs) on the liver surface, have also<br />

been suggested. It is unknown how these cells contribute to<br />

fibrogenesis due to insufficient availability of markers and isolation<br />

methods. Methods: We isolated HSCs, PFs and MCs from<br />

collagen1a1 (Col1a1)-GFP transgenic mouse livers by fluorescence-activated<br />

cell sorting (FACS) and identified their markers<br />

by microarray analysis. Primary HSCs, PFs and MCs were<br />

cultured in the presence of TGF-β1 or PDGF-BB to examine their<br />

myofibroblastic conversion and proliferation. Results: Immunohistochemistry<br />

showed that Col1a1-GFP mouse livers broadly<br />

express GFP in quiescent HSCs, PFs, and MCs. Combining<br />

vitamin A (VitA) lipid autofluorescence with GFP expression,<br />

we separated the VitA+ HSCs and VitA-GFP+ population from<br />

the Col1a1-GFP livers through FACS. We further subtracted<br />

glycoprotein M6A (GPM6A)+ MCs from the heterogeneous<br />

VitA-GFP+ population and obtained a PF-enriched fraction as<br />

VitA-GFP+GPM6A- cells. Microarray analysis identified selective<br />

expression of Reelin in HSCs. PFs exclusively expressed<br />

Ectonucleoside triphosphate diphosphohydrolase-2 (ENTPD2/<br />

NTPDase2). Isolated PFs exhibited fibroblastic morphology<br />

without storing VitA lipids in culture. MCs showed a round<br />

morphology and expressed CD200, mesothelin, podoplanin,<br />

and uroplakin 1b. HSCs, PFs, and MCs that were isolated from<br />

normal Col1a1-GFP livers transformed into myofibroblasts that<br />

expressed α-smooth muscle actin induced by TGF-β1 in culture.<br />

TGF-β1 suppressed Cyclin D mRNA expression in PFs but not in<br />

HSCs and MCs. By contrast, PDGF-BB induced Cyclin D mRNA<br />

only in HSCs. Conclusion: HSCs, PFs, and MCs have the potential<br />

to differentiate into myofibroblasts, and their proliferation is<br />

regulated differently by TGF-β1 and PDGF-BB.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ingrid A. Lua, Kinji Asahina<br />

1355<br />

Selective disruption of dynamin GTPase activity in HSC<br />

increases murine fibrogenesis in vivo and promotes HSC<br />

activation in vitro<br />

Qian Ding 1 , Ruisi Wang 1 , Thiago de Assuncao 1 , Meng Yin 2 ,<br />

Sheng Cao 1 , Usman Yaqoob 1 , Richard Ehman 2 , Robert C. Huebert<br />

1 , Vijay Shah 1 ; 1 Gastroenterology Research Unit, Mayo Clinic,<br />

Rochester, MN; 2 Department of Radiology, Mayo Clinic, Rochester,<br />

MN<br />

Background: Dynamin-2 (Dyn2) is a large GTPase responsible<br />

for endocytosis. A point mutation of Dyn2 at lysine 44 to<br />

alanine disrupts dynamin GTPase activity (Dyn2K44A). Prior<br />

<strong>studies</strong> using a mouse we generated that conferred Dyn2K44A<br />

expression selectively in endothelial cells impaired angiogenesis<br />

in vivo by disrupting receptor tyrosine kinase endocytosis<br />

and signaling. We hypothesized that expression of Dyn2K44A<br />

in hepatic stellate cells (HSC) would similarly disrupt tyrosine<br />

kinase dependent HSC activation and fibrogenesis in vivo.<br />

Methods and Results: Dyn2K44A fl/fl mice were crossed with<br />

Collagen 1-Cre (Col1 Cre ) mice to generate offspring with HSC<br />

selective expression of Dyn2K44A (Col1 Cre /Dyn2K44A fl/fl ).<br />

Primary HSC isolates showed Dyn2K44A expression by realtime<br />

PCR. Col1 Cre /Dyn2K44A fl/fl mice and littermate controls<br />

were subjected to carbon tetrachloride (CCl 4<br />

) or olive oil vehicle<br />

for 6 weeks, or bile duct ligation (BDL) or sham operation<br />

for 3 weeks to induce fibrosis and then sacrificed. Contrary<br />

to our hypothesis, immunofluorescence for α-smooth muscle<br />

actin (α-SMA) and fibronectin (FN), and sirius red staining from<br />

fixed liver tissues was significantly greater in response to both<br />

BDL and CCl 4<br />

in Col1 Cre /Dyn2K44A fl/fl mice compared with<br />

littermate controls (~1.5-fold; p


876A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

exosome binding to HSC by RGD (p < 0.05) but not by RGE.<br />

Binding of PKH26-stained HSC-derived exosomes to PKH67-<br />

stained HSC, assessed over 12 hrs, was also dose-dependently<br />

inhibited by EDTA (0-500 μM; p < 0.05). Since these data suggested<br />

exosome-HSC interactions involve integrins and since<br />

all five αV integrins (β1, β3, β5, β6, β8) and two β1 integrins<br />

(α5, α8) are RGD-dependent, recipient HSC were transfected<br />

for 24 hrs with siRNA to αV or β1 and then incubated with<br />

PKH26-stained exosomes for 24 hrs. Each treatment alone was<br />

effective in blocking >95% exosome binding. Finally, binding<br />

of HSC-derived exosomes to recipient HSC was reduced by<br />

heparin but not chondroitin sulfate (100μg/ml) while exosomal<br />

miR-214-mediated suppression of CTGF 3’-UTR activity, accomplished<br />

by 24-hr co-culture of miR-214-transfected donor HSC<br />

with CTGF 3’-UTR luciferase reporter-transfected recipient HSC<br />

was reversed in the presence of heparin but not chondroitin<br />

sulfate (100μg/ml) (p < 0.05). Thus exosome-HSC interactions<br />

or downstream gene activity are heparin- or integrin-dependent,<br />

the latter of which requires expression of αV and/or β1<br />

subunits by recipient HSC.<br />

Disclosures:<br />

David Brigstock - Stock Shareholder: FibroGen<br />

The following authors have nothing to disclose: Li Chen, Ruju Chen, Sherri Kemper<br />

1357<br />

Potential of connective tissue growth factor as a novel<br />

therapeutic target against hepatic fibrosis<br />

Yuki Makino, Hayato Hikita, Yugo Kai, Yasutoshi Nozaki, Tasuku<br />

Nakabori, Yoshinobu Saito, Satoshi Tanaka, Ryotaro Sakamori,<br />

Naoki Hiramatsu, Tomohide Tatsumi, Tetsuo Takehara; Department<br />

of Gastroenterology and Hepatology, Osaka University<br />

Graduate School of Medicine, Suita, Japan<br />

Background and aim: Connective tissue growth factor (CTGF)<br />

is a multifunctional matricellular protein reported to be up-regulated<br />

in the liver of patients with chronic liver diseases. However,<br />

the association of CTGF and liver fibrosis has not been<br />

fully studied. We analyzed the significance of CTGF and its<br />

potential as a therapeutic target in hepatic fibrosis. Methods/<br />

Results: Using clinical resected livers of 93 patients under the<br />

approve of Research Ethics Committee, we revealed that CTGF<br />

expression levels in non-tumorous liver tissue increased along<br />

with the progression of fibrotic degrees of the liver and positively<br />

correlated with those of other fibrosis-related genes such<br />

as α-SMA, TGF-β1, and Col1a1. In a mouse experiment, CTGF<br />

was up-regulated in the liver 3 weeks after bile duct ligation<br />

(BDL). CTGF expression levels showed positive correlation with<br />

those of other fibrosis-related genes listed above. Immunochemistry<br />

of CTGF after BDL showed positive reaction both in parenchymal<br />

cells (PC) and non-parenchymal cells (NPC). Thus, we<br />

subsequently isolated PC and NPC from mice underwent BDL<br />

or sham, and compared respective CTGF expression levels<br />

between BDL and sham groups. CTGF was up-regulated both<br />

in PC and NPC in BDL groups. In vitro experiments revealed<br />

that treatment of TGF-β, an inducing factor of CTGF, increased<br />

CTGF expression and secretion not only in hepatocyte cell<br />

lines, Huh7 and PLC/PRF/5 cells but in hepatic stellate cell<br />

(HSC) cell lines, LX-2 cells. These findings indicated that both<br />

hepatocytes and HSC produce CTGF under the stimulation of<br />

BDL. We finally analyzed the potential of CTGF as a therapeutic<br />

target in liver fibrosis. In vitro CTGF treatment induced the<br />

activation and collagen production in LX-2 cells, suggesting<br />

CTGF has a promoting effect for liver fibrosis. We subsequently<br />

evaluated the degree of fibrosis after BDL in 3 strains of conditional<br />

CTGF knockout mice; HSC-specific CTGF deficient mice<br />

(GFAP-Cre+ CTGF flox/flox mice; CTGFΔHSC mice), hepatocyte-specific<br />

CTGF deficient mice (Alb-Cre+ CTGF flox/flox<br />

mice; CTGFΔHep mice), and poly IC-induced PC and NPC<br />

CTGF deficient mice (MX1-Cre+ CTGF flox/flox mice; CTG-<br />

FΔPC+NPC mice). As a result, compared with Cre- littermates,<br />

CTGF expression levels in the liver were significantly reduced<br />

to 55.8% and 27.7% in CTGFΔHep and ΔPC+NPC mice,<br />

respectively. No difference was observed in CTGFΔHSC mice.<br />

Liver fibrosis was attenuated only in CTGFΔPC+NPC mice, evidenced<br />

by reduced Collagen expression and Sirius Red-stained<br />

area. Conclusion: CTGF is produced from both PC and NPC<br />

and promotes hepatic fibrosis. CTGF is a promising therapeutic<br />

target in liver fibrosis.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yuki Makino, Yugo Kai, Yasutoshi<br />

Nozaki, Tasuku Nakabori, Yoshinobu Saito, Satoshi Tanaka, Ryotaro Sakamori,<br />

Naoki Hiramatsu, Tomohide Tatsumi<br />

1358<br />

Exosome-mediated activation of TLR3 in stellate cells<br />

stimulates IL-17A production of γδ T cells in liver fibrosis<br />

of mice<br />

Wonhyo Seo, Hyuk-Soo Eun, So Yeon Kim, Seol-Hee Park, Jong-<br />

Min Jeong, Won-Mook Choi, Myung-Ho Kim, Won-IL Jeong;<br />

KAIST, Daejeon, Korea (the Republic of)<br />

Background: During liver injury, hepatocytes secrete numerous<br />

exosomes including diverse types of self-RNAs. Recently,<br />

self-noncoding RNA has been considered as an activator to tolllike<br />

receptor 3 (TLR3). However, the roles of hepatic exosome<br />

and its detection by TLR3 have not been fully understood in the<br />

pathogenesis of liver fibrosis. In the present study, we identified<br />

that exosome-induced activation of TLR3 in hepatic stellate cells<br />

(HSCs) stimulated production of interleukin-17A (IL-17A) in γδ<br />

T cells, subsequently leading to acceleration of liver fibrosis<br />

in mice. Methods: Liver injury and fibrosis was induced with<br />

single or 2-week injection of carbon tetrachloride (CCl 4<br />

) in wild<br />

type (WT) and TLR3-deficient (TLR3 -/- ) mice. Liver immune cells,<br />

HSCs and hepatocytes were freshly isolated for flow cytometry<br />

and in vitro experiments. Results: In acute and chronic liver<br />

injuries, increased IL-17A production was mainly detected in<br />

hepatic γδ T cells of WT mice, whereas those of TLR3 -/- mice did<br />

not show any increase of IL-17A production in both liver injuries.<br />

In addition, liver fibrosis of TLR3 -/- mice was significantly<br />

attenuated compared with WT mice. More interestingly, IL-17A<br />

producing γδ T cells were in close contact with activated HSCs,<br />

suggesting implication of HSCs and TLR3 in IL-17A production<br />

of γδ T cells. Furthermore, in vitro treatments of exosomes<br />

derived from CCl 4<br />

- or ethanol-treated hepatocytes significantly<br />

increased expression of IL-1β, IL-23 and IL-17A in HSCs, and<br />

remarkable increase of IL-17A production in γδ T cells was<br />

observed as they were co-cultured with exosome-treated WT<br />

HSCs or conditioned medium from TLR3-activated WT HSCs.<br />

However, all these finding were not detected as γδ T cells were<br />

co-cultured with exosome-treated HSCs of IL-17A -/- or TLR3 -<br />

/-<br />

mice. Using reciprocal bone marrow transplantation of WT<br />

and TLR3 -/- mice, we demonstrated TLR3 deficiency in HSCs<br />

contributed to decreased liver fibrosis and IL-17A production<br />

in γδ T cells. Conclusion: In liver injury, exosome-mediated activation<br />

of TLR3 in HSCs accelerates liver fibrosis by enhancing<br />

IL-17A production of γδ T cells, which might be associated<br />

with IL-17A production of HSCs by unknown self-TLR3 ligands


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 877A<br />

released from damaged hepatocytes. Therefore, TLR3 might be<br />

a novel therapeutic target for liver fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Wonhyo Seo, Hyuk-Soo Eun, So<br />

Yeon Kim, Seol-Hee Park, Jong-Min Jeong, Won-Mook Choi, Myung-Ho Kim,<br />

Won-IL Jeong<br />

Effects of ASK1 inhibitors on fatty acid metabolism<br />

1359<br />

Efficacy of an ASK1 Inhibitor to Reduce Fibrosis and<br />

Steatosis in a Murine Model of NASH is Associated with<br />

Normalization of Lipids and Hepatic Gene Expression<br />

and a Reduction in Serum Biomarkers of Inflammation<br />

and Fibrosis<br />

Satyajit Karnik 1 , Michael R. Charlton 2 , Li Li 1 , Michelle Nash 1 ,<br />

Maisoun Sulfab 1 , David Newstrom 1 , Erik G. Huntzicker 1 , Dorothy<br />

French 1 , Zachary D. Goodman 3 , Tracy Shafizadeh 4 , Steve Watkins<br />

4 , David Breckenridge 1 , Daniel Tumas 1 ; 1 Biology Research,<br />

Gilead Sciences, Foster City, CA; 2 Gastroenterology and Hepatology,<br />

Mayo Clinic, Rochester, MN; 3 Center for Liver Diseases,<br />

Inova Fairfax Hospital, Falls Church, VA; 4 Metabolon Inc, West<br />

Sacramento, CA<br />

BACKGROUND & SIGNIFICANCE: ASK1 is a kinase that is activated<br />

by various stimuli including oxidative stress, TGF-β, and<br />

hyperglycemia. ASK1 induces apoptosis, fibrosis, and metabolic<br />

dysfunction by activating p38 and JNK1. We previously<br />

demonstrated that ASK1 inhibition is efficacious in preclinical<br />

models of NASH and liver fibrosis. GS-4997, a first-in-class,<br />

oral small molecule ASK1 inhibitor, is being investigated in a<br />

Phase 2 clinical trial in NASH. Here, we further characterize<br />

the efficacy of ASK1 inhibition in a murine model of NASH by<br />

evaluating hepatic gene expression, hepatic and serum lipids,<br />

and serum biomarkers and cytokines. METHODS: NASH was<br />

induced in male C57BL/6 mice by administration of a diet<br />

high in fat, cholesterol, and sugar for 240 days. Animals were<br />

subsequently treated with either placebo or a selective, small<br />

molecule inhibitor of ASK1 for 90 days (n=15/group). Endpoints<br />

included steatosis evaluated by morphometry, clinical<br />

pathology, liver hydroxyproline levels, liver and plasma lipids<br />

by chromatography, hepatic gene expression by microarray,<br />

serum levels cytokines and fibrosis biomarkers by Luminex and<br />

ELISA. RESULTS: ASK1 inhibition reduced hepatic fibrosis, steatosis,<br />

and insulin resistance. ASK1 inhibitor treated animals<br />

had increased lipid metabolism and decreased de novo lipogenesis<br />

vs. controls as evaluated by lipid profiling in liver.<br />

ASK1 inhibitor treated animals had a 26% decrease in lipogenesis,<br />

a 21% decrease in stearoyl-CoA desaturase, and a<br />

78% increase in delta-5 desaturase relative to placebo controls<br />

as reflected in the plasma lipidome (Table). Consistent with<br />

these effects, gene networks regulating fatty acid synthesis,<br />

lipid metabolism, and cholesterol biosynthesis were differentially<br />

expressed in ASK1 inhibitor treated animals vs. placebo<br />

controls. Serum levels of IL-6, and the fibrosis markers osteopontin,<br />

hyaluronic acid, and TIMP-1 were reduced by 50%,<br />

35%, 33% and 41%, respectively, in ASK1 inhibitor treated<br />

animals vs. controls. CONCLUSIONS: The therapeutic benefit<br />

of an ASK1 inhibitor in reducing hepatic steatosis and fibrosis<br />

was associated with normalization of fatty acid synthesis and<br />

lipid metabolism. These data enhance our understanding of the<br />

mechanisms of action of ASK1 inhibition and further support<br />

evaluation of the ASK1 inhibitor, GS-4997, in patients with<br />

NASH.<br />

Disclosures:<br />

Satyajit Karnik - Consulting: Monsoon Dx; Employment: Gilead Sciences<br />

Michael R. Charlton - Grant/Research Support: GIlead Sciences, Merck, Janssen,<br />

AbbVie, Novartis<br />

Li Li - Employment: Gilead Sciences<br />

Michelle Nash - Employment: Gilead Sciences<br />

Erik G. Huntzicker - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Dorothy French - Employment: Gilead Sciences; Stock Shareholder: Genentech<br />

- Roche<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

David Breckenridge - Employment: Gilead Sciences<br />

Daniel Tumas - Employment: Gilead Sciences, Inc<br />

The following authors have nothing to disclose: Maisoun Sulfab, David Newstrom,<br />

Tracy Shafizadeh, Steve Watkins<br />

1360<br />

Endothelial Niche Derived Notch Signaling Pathway Balanced<br />

Liver Regeneration and Fibrosis in Distinct Ways<br />

Lin Wang; Hepatic Surgery, Fourth Military Medical University,<br />

Xi`an, China<br />

Abstract: Traumatic or toxic damage to the liver is often<br />

associated with liver regeneration and fibrosis. Liver sinusoidal<br />

endothelial niche has been addressed to modulate liver<br />

regeneration and fibrosis. Previously we found, Notch signaling<br />

was involved into the determination of liver sinusoidal<br />

endothelial cell (SEC) proliferation and differentiation by an<br />

angiocrine manner, and consequently influenced liver regeneration.<br />

However, the specific modulations and cell communications<br />

by endothelial niche-derived Notch signaling on<br />

liver reconstruction are still unknown. In this study, by using<br />

endothelial specific Notch Intracellular Domain (NICD) activated<br />

transgenic models (VeCad-CreERT-NICD) to persistently<br />

activate Notch signal, we firstly found SEC fenestration and<br />

sieve plates dramatically decreased both in vivo and in vitro.<br />

Meanwhile, sinusoidal perfusion and SEC endocytosis ability<br />

were significantly compromised compared to the control.<br />

Intriguingly, incapable hepatocyte proliferation, activated<br />

hepatic stellate cells (HSC) and liver malfunction were more<br />

remarkable in VeCad-CreERT-NICD mice. In 70% hepatectomy<br />

models, hepatocyte regeneration was largely blocked by NICD<br />

activation. By contrast, hepatic fibrosis induced by CCl 4<br />

injection<br />

was apparently aggravated in Notch activated models.<br />

Excitingly, in different co-culture systems, isolated primary SEC<br />

with loss of fenestrations by Notch activation disrupted wild<br />

type hepatocyte proliferation, but initiated HSC activation. To<br />

illustrate the underlying mechanisms, we detected Wnt2, which<br />

was considered as a mitogen of hepatocytes, was downregulated<br />

in VeCad-CreERT-NICD SEC. Overexpression of Wnt2<br />

successfully restored incapable hepatocyte proliferation in<br />

Notch activated mice. Besides, prominently increased ET1 was<br />

noticed in Notch activated SEC and potentially accelerated<br />

liver fibrosis in a Notch-ET1 dependent manner. Conclusion:<br />

We demonstrated Notch signal as a bi-directional switch in<br />

balancing hepatic regeneration and fibrosis through sinusoidal<br />

endothelial niche derived pro-regenerative Notch-Wnt2 versus<br />

pro-fibrotic Notch-ET1 pathways.


878A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Lin Wang<br />

1361<br />

Liver fatty acid-binding protein is required for Perilipin<br />

5 to restore lipid droplets in activated hepatic stellate<br />

cells<br />

Jianguo Lin 1 , Elizabeth P. Newberry 2 , Nicholas O. Davidson 2 ,<br />

Anping Chen 1 ; 1 Pathology, Saint Louis University, St. Louis, MO;<br />

2 Medicine, Washington University in St. Louis, St. Louis, MO<br />

Hepatic stellate cell (HSC) activation is coupled to the loss of<br />

lipid droplets (LDs) which are specialized organelles composed<br />

of neutral lipids surrounded by perilipins (Plins), among which<br />

Plin5 plays a key role in regulating aspects of intracellular<br />

trafficking, signaling and cytoskeletal organization in hepatocytes.<br />

Recent work in Plin5 ko mice suggests a role in high fat<br />

diet induced hepatic lipotoxicity. We now show that high fat<br />

diet induced liver fibrosis is accompanied by >70% reduction<br />

in HSC Plin5 and that spontaneous activation of primary<br />

HSCs produces temporally coincident loss of Plin5 expression<br />

and LD depletion. Since modulating lipid content in HSCs is<br />

a suggested strategy for inhibiting HSC activation and treatment<br />

of hepatic fibrosis we asked whether exogenous Plin5<br />

expression in primary HSCs would reverse the activation phenotype<br />

and promote LD formation. Lentiviral Plin5 expression<br />

in primary wild type (WT) mouse HSCs restored LDs, increased<br />

lipid content and suppressed activation (~2-fold reduced procollagen<br />

expression), coincident with ~2 fold upregulation of<br />

both PPARgamma and L-Fabp expression. Since HSC specific<br />

expression of L-Fabp promotes FA accumulation and LDs and<br />

inhibits HSC activation in vitro, we next asked if L-Fabp expression<br />

is required for Plin5 rescue of the activated phenotype.<br />

Lentiviral Plin5 transduction of HSCs from L-Fabp –/– mice failed<br />

to restore lipogenic enzyme expression (FAS, SREBP1c) and<br />

showed significantly attenuated induction of PPARgamma (WT<br />

~5-fold induction vs KO ~2.5 fold). Lentiviral Plin5 transduction<br />

also failed to induce LD formation in HSCs from L-Fabp –/– mice<br />

and failed to reverse the activation phenotype (no change in<br />

aSMA or procollagen 1 mRNA, vs significant ~2-fold reduction<br />

in WT HSCs). To further verify the functional role of L-Fabp in<br />

Plin5 mediated pathways, we introduced adenoviral L-Fabp<br />

into HSCs from L-Fabp –/– mice, which completely rescued the<br />

ability of lentiviral Plin5 to restore LD formation and increase<br />

lipid content. Control, lacZ transduced HSCs from L-Fabp –/–<br />

mice failed to demonstrate increased lipid content or visible LDs<br />

after lentiviral Plin5 transduction. Taken together, our results<br />

indicate that Plin5 expression in HSCs plays a critical role in<br />

the activation phenotype both in vivo and in vitro. Furthermore,<br />

our findings demonstrate that HSC L-Fabp has a critical role in<br />

the pathways by which Plin5 modulates lipogenic gene expression,<br />

restoring LD formation and overall lipid content and in<br />

turn modulating HSC activation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jianguo Lin, Elizabeth P. Newberry,<br />

Nicholas O. Davidson, Anping Chen<br />

1362<br />

Deletion of Wntless in macrophages exacerbates liver<br />

fibrosis and the ductular reaction in chronic liver injury<br />

Katharine Irvine 1 , Andrew D. Clouston 1 , Antje Blumenthal 3 , Elizabeth<br />

E. Powell 1,2 ; 1 Centre for Liver Disease Research, The University<br />

of Queensland, Brisbane, QLD, Australia; 2 Department of<br />

Gastroenterology and Hepatology, Princess Alexandra Hospital,<br />

Brisbane, QLD, Australia; 3 UQ Diamantina Institute, The University<br />

of Queensland, Brisbane, QLD, Australia<br />

Introduction: Macrophages play critical roles in liver regeneration,<br />

fibrosis development and resolution. They are among the<br />

first responders to liver injury, and are implicated in orchestrating<br />

the fibrogenic response via multiple mechanisms. Macrophages<br />

are intimately associated with the activated hepatic<br />

progenitor cell (HPC) niche, or ductular reaction, that develops<br />

in parallel with fibrosis. Among the many macrophage-derived<br />

mediators implicated in liver disease progression, a key role<br />

for macrophage-derived Wnt proteins in driving pro-regenerative<br />

HPC activation towards a hepatocellular fate has been<br />

suggested. Wnt proteins, in general, however, have been<br />

associated with both pro- and anti-fibrogenic activities in the<br />

liver and other organs. We investigated the role of macrophage-derived<br />

Wnt proteins in fibrogenesis and HPC activation<br />

in murine models of chronic liver disease. Methods: We<br />

generated mice deficient in Wnt secretion from macrophages<br />

(MΦ) by deleting the Wls gene, which is essential for Wnt<br />

secretion, in LysM-Cre-expressing cells. Liver injury and fibrosis<br />

were induced by administration of thioacetamide (TAA, 12<br />

wks) in drinking water, or a choline-deficient ethionine-supplemented<br />

diet (CDE, 6 wks). Fibrosis and the ductular reaction<br />

were assessed histologically, and gene expression in whole<br />

liver and flow cytometry-sorted macrophages and HPC was<br />

assessed in MΦ-Wls-knockout mice and littermate controls.<br />

Results: Fibrosis and HPC activation were exacerbated in<br />

MΦ-Wls mice compared to littermate controls in TAA and CDE<br />

treated mice (TAA, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 879A<br />

1363<br />

Mast cells interact with proliferating cholangiocytes to<br />

activate hepatic stellate cells and promote fibrosis via<br />

TGF-β1 signaling during cholestatic injury<br />

Lindsey Kennedy 2 , Laura Hargrove 1 , Jennifer Owens 2 , Heather<br />

L. Francis 2,1 ; 1 Scott and White Memorial Hospital, Temple, TX;<br />

2 Research, Central Texas Veteran’s Health Care System, Temple,<br />

TX<br />

Background: Hepatic fibrosis is marked by activation of hepatic<br />

stellate cells (HSCs) and the release of TGF-β1. Cholestatic<br />

injury is a precursor to liver fibrosis and, following injury cholangiocytes<br />

acquire a neuroendocrine phenotype and interact<br />

with HSCs to promote fibrosis. Mast cells (MCs) infiltrate the<br />

liver following injury and release pro-fibrogenic factors including<br />

histamine that increases biliary proliferation and fibrosis.<br />

We tested the hypotheses that (i) MCs induce biliary proliferation,<br />

HSC activation and promote fibrosis and (ii) inhibition of<br />

MC-derived histamine decreases fibrosis via TGF-β1. Methods:<br />

To demonstrate that MCs promote fibrosis, wild-type (WT) and<br />

HDC -/- mice (characterized by lower MC count) were subjected<br />

to sham surgery or bile duct ligation (BDL) and injected with<br />

cultured MCs via tail vein. Further, WT and MDR2 -/- mice (a<br />

model of Primary Sclerosing Cholangitis characterized by<br />

increased MC count) were treated with cromolyn sodium to<br />

block MC-derived histamine. We collected liver blocks, serum,<br />

cholangiocytes and biliary supernatants. Biliary mass and proliferation<br />

were evaluated by immunohistochemistry for CK-19<br />

and PCNA, respectively. Fibrosis was evaluated in tissue sections<br />

by Sirius Red/Fast Green Collagen staining and by qPCR<br />

in total liver and cholangiocytes for α-SMA, fibronectin, collagen<br />

type 1a and TGF-β1. HSC activation was evaluated by<br />

qPCR in total liver and immunofluorescent staining in tissues<br />

for synaptophysin 9 (SYP9). TGF-β1 secretion was measured in<br />

serum from all groups. Downstream from TGFβ-1, we measured<br />

SMAD2 and SMAD3 by qPCR and western blotting in total<br />

liver and cholangiocytes. In vitro, cultured MCs were transfected<br />

with HDC shRNA to decrease histamine secretion and<br />

subsequently co-cultured with cholangiocytes or HSCs prior to<br />

measuring fibrosis markers, proliferation and TGF-β1 secretion.<br />

Results: Injected mast cells were found in close proximity to proliferating<br />

bile ducts. Biliary proliferation, fibrosis and HSC activation<br />

were increased in HDC -/- and BDL HDC -/- mice following<br />

MC injection. Inhibition of MC-derived histamine decreased<br />

biliary proliferation, fibrosis, TGF-β1 secretion and SMAD2/3<br />

expression in MDR2 -/- mice. In vitro, knockdown of MC HDC<br />

decreased (i) MC TGF-β1 secretion, (ii) biliary proliferation and<br />

fibrotic marker expression and (iii) HSC activation and TGF-β1<br />

secretion. Conclusion: MCs are recruited to proliferating cholangiocytes<br />

and promote HSC activation increasing fibrosis via<br />

TGF-β1/SMAD signaling. Inhibition of MC-derived histamine<br />

decreases fibrosis and regulation of MC mediators may be a<br />

therapeutic target for fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Lindsey Kennedy, Laura Hargrove,<br />

Jennifer Owens, Heather L. Francis<br />

1364<br />

Hepatocyte autophagy impairment promotes ductular<br />

reaction and fibrogenesis by hepatic stellate cells<br />

Youngmin A. Lee 1 , Luke A. Noon 1,2 , Tingfang Lee 1 , Marie-Luise<br />

Berres 3,6 , Fatemeh P. Parvin-Nejad 1 , Kemal M. Akat 4 , Hsin I<br />

Chou 1 , Varinder Athwal 1,7 , M. Isabel Fiel 5 , Ronald E. Gordon 5 ,<br />

Scott L. Friedman 1 ; 1 Liver Diseases, Icahn School of Medicine at<br />

Mount Sinai, New York, NY; 2 Centro de Investigación Principe<br />

Felipe (CIPF), Valencia, Spain; 3 Department of Internal Medicine<br />

III, University Hospital, RWTH Aachen, Aachen, Germany; 4 Laboratory<br />

of RNA Molecular Biology, HHMI/Rockefeller University,<br />

New York, NY; 5 Department for Pathology, Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 6 Department of Oncological<br />

Sciences, Icahn School of Medicine at Mount Sinai, New York,<br />

NY; 7 Institute of Human Development, Faculty of Medicine and<br />

Human Sciences, University of Manchester, Manchester, United<br />

Kingdom<br />

Background Ductular reaction (DR), an expansion of ductular<br />

progenitor-like cells, is often correlated with fibrosis in chronic<br />

liver disease and fulminant liver injury. Murine models in which<br />

autophagy, a homeostatic lysosomal pathway, is abrogated<br />

in hepatocytes by deletion of the autophagy related protein 7<br />

(Atg7) display a prominent DR with marked elevation of liver<br />

transaminases, steatohepatitis and fibrosis, with eventual tumorigenesis.<br />

Although hepatic stellate cells (HSCs) are associated<br />

with the DR, their interactions have not been well characterized<br />

and their relative contribution to fibrosis is uncertain. Our<br />

aim was to determine the contribution of ductular and HSCs<br />

to fibrosis when hepatocyte autophagy is impaired. Methods:<br />

Olig1-Cre:Atg7 F/F mice, which have selective hepatocyte<br />

depletion of the autophagy protein Atg7, were analyzed by<br />

immunohistochemistry (IHC), qRT-PCR and microarray for fibrogenic<br />

markers. Primary HSCs and ductular cells were isolated<br />

following collagenase perfusion and gradient purification.<br />

Enrichment of respective cell populations was confirmed by<br />

flow analysis and IHC. Gene expression was then analyzed by<br />

qRT-PCR. To identify fibrogenic cells in vivo, Olig1-Cre:Atg7 F/F<br />

mice were crossed to transgenic Col1a1-GFP mice, in which<br />

GFP expression is regulated by the Collagen1a1 promoter.<br />

Livers of Tg(Col1-GFP)cKO(Olig1-Cre:Atg7 F/F ) were analyzed<br />

by IHC for GFP as well as markers for ductular cells (EpCAM,<br />

CD133) and HSCs (desmin). Results: Olig1-Cre:Atg7 f/f mice<br />

develop extensive ductular reaction (Epcam + , CD133 + by IHC)<br />

and marked spontaneous fibrosis at 3 months, as assessed by<br />

sirius red staining (8.4 ± 2.4 vs 1.1±0.2 % fibrotic tissue area<br />

in controls), with significant fibrogenic gene induction in whole<br />

liver RNA based on qRT-PCR and array. Strikingly, collagen<br />

fibers were closely aligned with ductular cells (EpCAM + , Ck7 + )<br />

that were also intimately associated with HSCs (desmin + ).<br />

HSCs and ductular cells isolated from Olig1-Cre:Atg7 F/F both<br />

expressed aSMA, a marker of activated HSCs, and Col1,<br />

whereas none was expressed in control livers. In transgenic<br />

Col1a1-GFP reporter mice with hepatocyte autophagy loss,<br />

GFP expression was greater in desmin positive cells than in<br />

cells expressing progenitor markers (EpCAM, CD133), establishing<br />

HSCs as the primary fibrogenic cells. Conclusion:<br />

Hepatic fibrosis is a prominent and novel feature of hepatocyte<br />

Atg7 KO mice. HSCs intimately associated with ductular cells<br />

are the main fibrogenic cell population. These findings help<br />

clarify the basis of fibrogenesis associated with the DR and<br />

establish an experimental model for elucidating cell-cell interactions<br />

in this response.


880A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

The following authors have nothing to disclose: Youngmin A. Lee, Luke A. Noon,<br />

Tingfang Lee, Marie-Luise Berres, Fatemeh P. Parvin-Nejad, Kemal M. Akat, Hsin<br />

I Chou, Varinder Athwal, M. Isabel Fiel, Ronald E. Gordon<br />

1365<br />

Hepatocyte- and hepatic stellate cell-specific deletion of<br />

liver fatty acid binding protein (L-Fabp) reveals divergent,<br />

cell-type specific roles in hepatic lipid trafficking<br />

and fibrogenesis<br />

Elizabeth P. Newberry 1 , Susan M. Kennedy 1 , Gianfranco Alpini 3 ,<br />

Anping Chen 2 , Nicholas O. Davidson 1 ; 1 Washington University<br />

School of Medicine, St Louis, MO; 2 St. Louis University, St Louis,<br />

MO; 3 Central Texas Veterans Health Care System and Texas A&M<br />

HSC College of Medicine, Temple, TX<br />

Liver fatty acid binding protein (L-Fabp) modulates fatty acid<br />

(FA) trafficking, high fat diet-induced obesity, and steatosis. We<br />

previously showed that germline L-Fabp –/– mice have reduced<br />

FA uptake and incorporation into complex lipid in both hepatocytes<br />

and stellate cells (HSC), with attenuated diet-induced<br />

hepatic steatosis and fibrosis. Since L-Fabp is expressed in<br />

both hepatocytes and HSC, we examined its cell-specific role in<br />

lipid trafficking and fibrotic injury. Germline L-Fabp –/– mice fed<br />

a trans-fat, fructose (TFF) diet exhibit reduced steatosis and liver<br />

fibrosis, with a 70% decrease in Col1a1 and αSMA mRNAs.<br />

Hepatocyte-specific (Alb-Cre) L-Fabp deletors (Hep KO, >95%<br />

decreased L-Fabp mRNA) exhibit reduced hepatic TG (f/f, 560<br />

± 71mg/mg; Hep-KO, 284 ± 91; p=0.046), with decreased<br />

liver size and expression of Col1a1 and αSMA (~50%), suggesting<br />

that HSC activation in TFF-fed mice is linked to hepatocyte<br />

lipid accumulation. In contrast, HSC-specific (GFAP-Cre)<br />

L-Fabp deletion (HSC KO, ~75% reduction) did not affect liver<br />

size, hepatic steatosis or expression of Col1a1, but αSMA<br />

mRNA was reduced. Because GFAP Cre may target other cell<br />

types, we verified that WT cholangiocytes express undetectable<br />

L-Fabp. Freshly isolated HSCs from HSC-KO contain abundant<br />

lipid droplets, in contrast to germline L-Fabp –/– HSC which are<br />

lipid droplet depleted. Together these data indicate that TFF<br />

diet induced fibrotic injury is linked to hepatocyte TG content<br />

and suggest that lipid trafficking to HSCs may reflect FA compartmentalization<br />

in hepatocytes. In a distinct model of liver<br />

injury, carbon tetrachloride (CCl 4<br />

) administration produced<br />

worse injury in L-Fabp –/– mice vs C57BL/6J controls (ALT WT,<br />

666 ± 194 IU/L; KO, 1247 ± 179), with increased hepatocyte<br />

proliferation and enhanced expression of fibrogenic genes,<br />

yet no difference in fibrosis (WT, 2.6 ± 0.5%; KO, 2.4 ± 0.3,<br />

n=6). We then used HSC- and Hep- KOs to understand how<br />

L-Fabp deletion exacerbates CCl 4<br />

-induced liver injury. CCL 4<br />

-<br />

treated HSC KO mice exhibited higher serum ALT (f/f, 733<br />

± 306 IU/L; HSC-KO, 1450 ± 270, p=0.09) and hepatocyte<br />

proliferation (f/f, 7.2 ± 3.1 % BrdU+; HSC-KO 16.6 ± 1.6,<br />

p=0.02) with ~3-fold increased Col1a1 and αSMA expression.<br />

In contrast, CCl 4<br />

-treated Hep KO mice exhibited no difference<br />

in ALT (548 ± 186 IU/L) or hepatocyte proliferation<br />

versus controls. Together these data indicate that HSC L-Fabp<br />

expression plays a protective role against CCl 4<br />

injury, whereas<br />

hepatocyte L-Fabp promotes TFF diet-induced steatotic injury.<br />

Furthermore, the findings point to an unanticipated, cell specific<br />

role for L-Fabp in high fat diet induced versus toxin-mediated<br />

liver injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elizabeth P. Newberry, Susan M.<br />

Kennedy, Gianfranco Alpini, Anping Chen, Nicholas O. Davidson<br />

1366<br />

Fibrous network composition and cell heterogeneity<br />

modulate long-range force transmission by cells in<br />

matrices<br />

Robert N. Kent, Jude D. Han, Rebecca G. Wells; Gastroenterology,<br />

University of Pennsylvania, Philadelphia, PA<br />

Long-range force transmission is implicated in the large-scale<br />

architectural rearrangements occurring in fibrotic diseases like<br />

pulmonary fibrosis and liver cirrhosis. It was previously shown<br />

that matrix organized in a fibrous network is required and<br />

that collagen becomes aligned and compacted as part of this<br />

process. We hypothesized that the characteristics of the fibrous<br />

network as well as the heterogeneity of the cell population<br />

modulate long-range force transmission between cells. Fibroblasts<br />

(murine 3T3) were cultured as spheroids by suspension<br />

in a hanging droplet over 5 days. These were seeded in pairs<br />

500-1000 μm apart on thick gels of either denatured collagen<br />

(gelatin) or type I collagen (Col1) containing varying concentrations<br />

of type III collagen (Col3) or plasma fibronectin (pFN).<br />

Heterotypic spheroids containing varying ratios of primary rat<br />

hepatic stellate cells (HSCs) and hepatocytes were cultured in<br />

hanging droplets for 3 days and seeded on Col1 gels. Once<br />

seeded, spheroids were incubated or imaged via time-lapse<br />

microscopy for 24 hours and then fixed. Collagen alignment<br />

and compaction were visualized using second harmonic generation<br />

imaging, and the extent of alignment was quantified by<br />

calculating the anisotropy index in ImageJ (NIH). Gels without<br />

spheroids showed minimal collagen organization or alignment<br />

by second harmonic generation imaging. Col1 gels with spheroids<br />

showed marked alignment, contraction, and compaction<br />

of collagen along the axis between spheroids, with migration of<br />

cells along the compacted fibrils. Spheroids of 3T3 fibroblasts<br />

seeded on gelatin substrates showed minimal organization,<br />

suggesting that collagen cross-linking is required for cell-mediated<br />

network reorganization. On substrates consisting of 60%<br />

Col1 and 40% Col3, 3T3 spheroids generated significantly<br />

less fiber alignment when compared to spheroids on pure Col1<br />

substrates (p < 0.01). Col1 gels containing pFN (10% v/v),<br />

however, yielded a significant increase in alignment relative to<br />

pure Col1 (p < 0.05). Spheroids of primary HSCs reorganized<br />

the fibrous network such that the anisotropy index was significantly<br />

larger than the no-spheroid control (p < 0.05). The inclusion<br />

of hepatocytes in ratios of 1:20 and 1:1 hepatocyte:HSC<br />

yielded even greater network anisotropies than spheroids of<br />

HSCs alone (p < 0.01, 0.05). Collagen cross-linking, Col3 and<br />

pFN content, and heterogeneity of cell populations modulate<br />

long-range force transmission in fibrous networks. Determining<br />

the role of matrix cross-linking and cell heterogeneity in matrix<br />

reorganization may contribute to understanding the underlying<br />

mechanisms driving bridging in liver cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Robert N. Kent, Jude D. Han,<br />

Rebecca G. Wells


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 881A<br />

1367<br />

Dual combination therapy directed against lysyl oxidase-like<br />

2 (LOXL2) and apoptosis signal-regulating<br />

kinase 1 (ASK1) potently inhibits fibrosis and portal<br />

hypertension in a new mouse model of PSC-like liver<br />

disease<br />

Naoki Ikenaga 1 , Susan B. Liu 1 , Zhen-Wei Peng 1 , Andrew E.<br />

Greenstein 2 , Dorothy French 2 , Victoria Smith 2 , Yury Popov 1 ;<br />

1 Gatroenterology, Beth Israel Deaconnes, Boston, MA; 2 Gilead<br />

Sciences, Inc, Foster City, CA<br />

BACKGROUND/AIMS: We have previously characterized<br />

novel LOXL2 and ASK1 inhibitors as effective monotherapies<br />

for liver fibrosis in vitro and in vivo. They impact distinct,<br />

non-overlapping pro-fibrogenic signal transduction pathways,<br />

with LOXL2 mediating collagen cross-linking and ASK1 promoting<br />

hepatocyte injury and fibrogenic stellate cell activation<br />

via the p38 and JNK pathways. Here, the therapeutic efficacy<br />

of combined anti-LOXL2 and anti-ASK1 agents was evaluated<br />

in a mouse model of biliary fibrosis and portal hypertension.<br />

METHODS: We developed an improved BALBc.Mdr2-/- mouse<br />

model resembling human primary sclerosing cholangitis (PSC)<br />

with rapidly progressive fibrosis and early-onset portal hypertension<br />

(Ikenaga et al, 2015). 6 weeks of ASK1 inhibitor (GS-<br />

444217, 0.15% in diet), anti-LOXL2 monoclonal antibody<br />

(AB0023 mAb/GS-607601, 30mg/kg/week i/p), or their<br />

combination were administered to 6 weeks old BALB/c.Mdr2-<br />

/- mice with pre-established fibrosis (n=9-11/group). Portal<br />

venous pressure (PVP) was measured via direct cannulation of<br />

the portal vein with a micro-tip pressure monitor at the end of<br />

the study. Liver fibrosis was evaluated by histology and biochemical<br />

determination of collagen. RESULTS Both anti-LOXL2<br />

and anti-ASK1 mono-therapies significantly improved histological<br />

signs of fibrosis, with reduced hepatic collagen deposition<br />

(by 37 and 38% of hydroxyproline, respectively) in BALB/c.<br />

Mdr2-/- mice compared to placebo control group (p


882A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

sues and accelerates regression of CCl 4<br />

-induced fibrosis by<br />

stimulating expression of matrix metalloproteinase (MMP)-13<br />

and MMP-9. A subsequent study has identified opioid growth<br />

factor receptor-like 1 (OGFRL1) as a G-CSF-induced novel<br />

bone marrow cells-derived factor. To establish the molecular<br />

bases for future therapeutic application, we examined, in the<br />

present study, tissue distribution of endogenous OGFRL1 and<br />

the therapeutic effects of exogenous OGFRL1 administration.<br />

We also characterized the functions of OGFRL1 by generating<br />

its null mice. Methods: Total RNA was extracted from various<br />

organs/tissues of mice, and expression levels of Ogfrl1 gene<br />

were quantified using real time RT-PCR. Human bone marrow<br />

mesenchymal stem cells were infected with recombinant<br />

OGFRL1-expressing lentiviruses and injected into the spleen of<br />

CCl 4<br />

-treated nude mice. Hepatic expression of Mmp-13 and<br />

Mmp-9 mRNAs was examined by real time RT-PCR, and liver<br />

regeneration was estimated by BrdU uptake into hepatocytes.<br />

Ogfrl1 gene targeting was conducted using the CRISPR/Cas9<br />

system with a guide RNA set in the first exon of Ogfrl1 gene.<br />

Results: Endogenous OGFRL1 was expressed predominantly in<br />

neural tissues and, to the lesser extent, in hematopoietic organs<br />

such as bone marrow and spleen. Among the peripheral blood<br />

cells, monocytes and granulocytes were the major sources of<br />

OGFRL1. Liver tissues contained a small but significant amount<br />

of Ogfrl1 mRNA, which was decreased following repeated<br />

CCl 4<br />

injections. When fibrotic mice underwent partial hepatectomy,<br />

expression levels of endogenous Ogfrl1 gene were<br />

further declined. An intrasplenic injection of OGFRL1-expressing<br />

cells stimulated hepatic expression of MMP-13 and MMP-9,<br />

and increased the BrdU uptake into hepatocytes in fibrotic liver<br />

after partial hepatectomy. It also caused a remarkable increase<br />

in gene expression of cyclins and other cell cycle-related factors<br />

as well as that of AFP. OGFRL1 null mice exhibited no abnormalities<br />

in their developmental process or liver tissue architecture.<br />

However, mitosis of hepatocytes was delayed following<br />

partial hepatectomy in OGFRL1-null mice compared with wild<br />

type littermates. Conclusions: OGFRL1 plays a critical role in<br />

hepatocyte proliferation, and administration of OGFRL1-expressing<br />

cells could be a potential regeneration therapy for<br />

advanced liver fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takayo Yanagawa, Hideaki Sumiyoshi,<br />

Sachie Nakao, Masato Ohtsuka, Minoru Kimura, Yousuke Chiba, Shintaro<br />

Satake, Tadashi Moro, Hiromi Chikada, Akihide Kamiya, Yutaka Inagaki<br />

1370<br />

Interleukin 15 regulates transforming growth factor<br />

β1-induced collagen production in hepatic stellate cells<br />

by controlling a network of suppressors of collagen<br />

transcription<br />

Kohtaro Ooka 3 , Huanzi Han 1 , Holger Fey 1 , Costica Aloman 1 , Jingjing<br />

Jiao 2 ; 1 Division of Gastroenterology/Liver Diseases (MC716),<br />

University of Illinois at Chicago, Chicago, IL; 2 University of Texas<br />

MD Anderson Cancer Center, Houston, TX; 3 Medicine, University<br />

of Illinois at Chicago, Chicago, IL<br />

Background & Aim: Interleukin 15 (IL15) signaling is unique<br />

in that IL15 binds to the high affinity IL15 receptor a subunit<br />

(IL15Rα) which it shuttles to the cell surface and stimulates<br />

opposing cells through the β/γ receptor complex. Our group<br />

has identified a protective role for IL15 signaling in murine models<br />

of liver fibrosis using IL15Rα knockout mice (IL15RαKO).<br />

In addition to the established effects of IL15 on maintaining<br />

natural killer (NK) cell homeostasis, a cell type well known to<br />

limit fibrogenesis, a direct anti-fibrotic effect of IL15 signaling<br />

on radio-resistant liver cells has also been observed. We aim<br />

to understand in detail the mechanism of the anti-fibrotic effect<br />

of IL15 signaling on hepatic stellate cells (HSC), the major type<br />

of fibrogenic cells in the liver,. Methods: HSCs were isolated<br />

from wild type (WT) and IL15RαKO mice by gradient centrifugation<br />

followed by depletion of contaminating leukocytes and<br />

endothelial cells. Purity of HSC was validated by the absence<br />

of mRNA for CD31 and CD45. HSC from WT and IL15RαKO<br />

were assessed by quantitative real time PCR (qPCR) and western<br />

blot to measure fibrogenic markers and well-known negative<br />

regulators of collagen transcription. Results: IL15RαKO<br />

HSCs display enhanced fibrogenesis both in the steady-state<br />

and under transforming growth factor β1 (TGF-β1) stimulation<br />

as shown by qPCR of collagen 1A1 (COL1A1), collagen<br />

3A1 (COL3A1) and platelet-derived growth factor receptor<br />

(PDGFR). Treatment with the IL15/IL15Rα complex reduced<br />

expression of COL1A1 and alpha smooth muscle actin (αSMA)<br />

in IL15RαKO HSCs and reduced the magnitude of αSMA<br />

expression induced by TGF-β1. Expression of several negative<br />

regulators involved in the collagen expression pathway including<br />

TP53, serine-threonine kinase receptor-associated protein<br />

(STRAP) and ski-like (SKIL) were lower in IL15RαKO HSCs and<br />

were further down-regulated by TGF-β1 treatment. Conclusion:<br />

IL15 signaling regulates suppressors of collagen transcription<br />

to have a direct anti-fibrotic effect on HSCs. Our study highlights<br />

a possible therapeutic effect of IL15 and IL15/IL15Rα in<br />

liver fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kohtaro Ooka, Huanzi Han,<br />

Holger Fey, Costica Aloman, Jingjing Jiao<br />

1371<br />

NLRP3 inflammasome modulation of hepatic stellate cell<br />

activation: role in liver fibrosis<br />

Alexander Wree 1,2 , Matthew D. McGeough 1 , Davide Povero 1 ,<br />

Akiko Eguchi 1 , Hal M. Hoffman 1 , Ariel E. Feldstein 1 ; 1 Pediatrics,<br />

UCSD, La Jolla, CA; 2 Internal Medicine III, University Hospital,<br />

RWTH-Aachen, Aachen, Germany<br />

Background: NLRP3 inflammasome activation has been<br />

increasingly recognized as an important mechanism of liver<br />

damage. Our group has recently demonstrated that isolated<br />

NLRP3 inflammasome activation is sufficient to induce liver<br />

inflammation and fibrosis (Hepatology 2014 Mar). Here we<br />

tested the hypothesis that the NLRP3 inflammasome in hepatic<br />

stellate cells (HSCs) can directly modulate their activation and<br />

contribution to liver fibrosis. Methods: An immortalized human<br />

HSC line (LX-2) was incubated with the NLRP3 inflammasome<br />

inducers, LPS (as initial priming signal) and ATP (as activating<br />

signal). Additionally, primary HSCs were isolated from<br />

Nlrp3 D301NneoR/+ knock-in mice on a C57BL/6 background<br />

which were crossed to C57BL/6.Cg-Tg(Cre/Esr1)–Estrogen<br />

Receptor Cre (CreT) mice allowing for temporal control of<br />

NLRP3 activation. In a third set of experiments, mutant Nlrp3<br />

was activated in the aforementioned mice in vivo for over 16<br />

weeks. Activation of the NLRP3 inflammasome and HSCs was<br />

assessed via fluorescent imaging techniques and mRNA analysis.<br />

Results: LX-2 cells incubated with LPS (1ug/ml, over 3h)<br />

showed increased mRNA levels of Nlrp3 (1.5 fold compared<br />

to control), Incubation with ATP (5mM) resulted in an additional<br />

increase (2.0 fold). Collagen-1 mRNA levels were only<br />

raised in LX-2 cells when incubated with LPS and ATP (1.5<br />

fold compared to control or LPS only). Fluorescently labeled<br />

active Caspase 1 was present in LX-2 cells when incubated<br />

with LPS + ATP. The NLRP3 inflammasome was induced in<br />

vitro in cultured primary HSCs via 4OH-tamoxifen treatment.<br />

Immunofluorescence for F-actin showed cytoskeleton reorga-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 883A<br />

nization as early as 6h after 4OH-tamoxifen incubation with<br />

mRNA levels of Nlrp3, alpha smooth muscle actin and tissue<br />

inhibitor of metalloproteinase 1 being increased. Activation of<br />

the NLRP3 inflammasome in Nlrp3 D301N/+ /CreT mice over 16<br />

weeks resulted in hepatomegaly while tamoxifen-injected control<br />

mice showed regular liver weights (liver to body weight %<br />

- 6.1% ± 0.32% vs. 4.6% ± 0.03%, p < 0.05). Analysis of liver<br />

histology revealed increased inflammation with neutrophilic<br />

infiltration and increased mRNA levels of myeloperoxidase<br />

(fold-change to WT 52.8 ± 13.75, p


884A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

by the same parameters as above. We conclude that simultaneous<br />

targeting of peripheral CB 1<br />

R and iNOS by novel hybrid<br />

inhibitors offers improved efficacy and safety in the pharmacotherapy<br />

of liver fibrosis.<br />

Disclosures:<br />

Resat Cinar - Employment: National Institutes of Health; Patent Held/Filed:<br />

National Institutes of Health<br />

George Kunos - Patent Held/Filed: US Patent<br />

The following authors have nothing to disclose: Malliga Iyer, Ziyi Liu, Tony Jourdan,<br />

Zongxian Cao, Katalin Erdelyi, Grzegorz Godlewski, Gergo Szanda, Jie<br />

Liu, Pal Pacher, Kenner Rice<br />

1375<br />

Proof of liver regeneration after resolution of fibrosis by<br />

siRNA against HSP47 encapsulated in vitamin A-coupled<br />

liposome employing CCl 4<br />

mice engineered with<br />

Sox9-Cre ERT2 /ROSA-YFP genes<br />

Naoki Fujitani 1 , Akihiro Yoneda 2 , Yuya Murakami 1 , Miyuki<br />

Nishimura 1 , Miyono Miyazaki 3 , Keiko Kajiwara 3 , Kenjiro<br />

Minomi 3 , Yasuaki Tamura 2 , Yoshiro Niitsu 1 ; 1 School of Medicine,<br />

Department of Molecular Exploration, Sapporo Medical University,<br />

Sapporo, Japan; 2 Department of Molecular Therapeutics, Center<br />

for Food and Medical Innovation, Institute for the Promotion of<br />

Business-Regional Collaboration, Sapporo, Japan; 3 Hokkaido Laboratory,<br />

Nitto Denko Corporation, Sapporo, Japan<br />

Background and Aim We have previously shown that hepatic<br />

fibrosis could be resolved by the administration of siRNA<br />

against HSP47 encapsulated in vitamin A-coupled liposome<br />

(VA-lip-siRNA HSP47) in various liver cirrhosis rodent models<br />

(Sato et al. Nat. Biotechnol. 2008). However, a real clinical<br />

benefit of such anti-fibrosis modality would be a recovery of<br />

impaired liver function with tissue regeneration. We therefore,<br />

in the present study intended to demonstrate that fibrotic liver<br />

is capable to regenerate after resolution of fibrosis, employing<br />

bile duct ligation (BDL) model rats, dimthylnitrosoamine (DMN)<br />

model rats and carbon tetrachloride (CCl 4<br />

) model of transgenic<br />

mice carrying Sox9-Cre ERT2 /ROSA-YFP genes. Methods BDL<br />

and DMN models of rats were established as reported earlier<br />

(Sato et al. Nat. Biotechnol. 2008). Transgenic mice with<br />

Sox9-Cre ERT2 /ROSA-YFP genes were produced by repeated<br />

crossing of Sox9-Cre ERT2 mice onto ROSA-YFP mice obtained<br />

from The Jackson Laboratory. Fibrosis was induced with the<br />

intraperitoneal administration of CCl 4<br />

for 10 weeks. VA-lipsiRNA<br />

HSP47 was formulated as described earlier (Sato et al.<br />

Nat. Biotechnol. 2008). Immuno-stainings for Sox9, αSMA,<br />

EpCAM, YFP, HNF4α, BrdU and PCNA were carried out using<br />

each specific antibodies. Results and discussion In the liver of<br />

BDL rats, EpCAM positive; progenitors were localized at portal<br />

area and were surrounded by deposited collagen and αSMA<br />

positive cells, activated stellate cells (aHSC) as revealed by<br />

immune-staining. In the liver of DMN rats which were administrated<br />

with VA-lip-siRNA HSP47 for 4 times, both aHSC and<br />

progenitors markedly diminished, and PCNA staining onto<br />

HNF4α positive hepatocytes was significantly increased as<br />

compared with that of control rats, suggesting a concept that<br />

in the cirrhotic liver progenitors surrounded by aHSCs forming<br />

niche like structure underwent differentiation into hepatocytes<br />

upon destruction of niche structure with administration<br />

of VA-lip-siRNA HSP47. To verify this concept, fate-tracing of<br />

progenitor after resolution of fibrosis was conducted with Sox9-<br />

Cre ERT2 /ROSA-YFP mice. When the CCl 4<br />

mice were injected<br />

with tamoxifen, followed by the treatment of VA-lip-siRNA<br />

HSP47 for 10 times, initiating at 10 th week of CCl 4<br />

administration,<br />

and YFP expression was examined at 13 th week of<br />

CCl 4<br />

administration, HNF4α positive hepatocytes were clearly<br />

immune-stained for YFP. Thus, present study demonstrates that<br />

cirrhotic liver is capable of undergoing regeneration with our<br />

anti-fibrosis treatment even under the condition of continuous<br />

exposure to hepatotoxic agent.<br />

Disclosures:<br />

Miyuki Nishimura - Grant/Research Support: Nitto Denko<br />

The following authors have nothing to disclose: Naoki Fujitani, Akihiro Yoneda,<br />

Yuya Murakami, Miyono Miyazaki, Keiko Kajiwara, Kenjiro Minomi, Yasuaki<br />

Tamura, Yoshiro Niitsu<br />

1376<br />

Oral hepatotropic DPP4 inhibitors attenuate NASH and<br />

bilary fibrosis in part through macrophage modulation<br />

Xiaoyu Wang 1 , Shih-Yen Weng 1 , Yong Ook Kim 1 , Thomas Klein 3 ,<br />

Detlef Schuppan 1,2 ; 1 Institue of Translational Immunology, Mainz,<br />

Germany; 2 Division of Gastroenterology, Boston, MA; 3 Boehringer<br />

Ingelheim Pharma, Biberach an der Riss, Germany<br />

Background and aims: Non-alcoholic steatohepatitis (NASH) is<br />

characterized by steatosis, lobular inflammation and progressive<br />

parenchymal fibrosis. Glucagon like peptide 1 (GLP-1)<br />

analogues are proven drugs to treat type 2 diabetes and insulin<br />

resistance. GLP-1 levels are increased by agents that inhibit cell<br />

surface dipeptidyl peptidase-4 (DPP-4) which is ubiquitously<br />

expressed and rapidly inactivates secreted GLP-1. Here, we<br />

studied the effect of two hepatotropic DPP4-inhibitors on liver<br />

inflammation and fibrosis in models of NASH and biliary fibrosis.<br />

Methods: Linagliptin and Sitagliptin were administered<br />

daily by oral gavage to Mdr2KO mice and to C57BL/6mice<br />

fed a methionine and choline deficient (MCD) diet for 6 weeks.<br />

Hepatic fibrosis was assessed by morphometric analysis of<br />

Sirius red stained collagen and measurement of hydroxyproline<br />

content. Hepatic inflammation and fibrosis were assessed by<br />

semiquantitative immunohistochemistry. Fibrosis and inflammation<br />

related transcript levels were measured by quantitative<br />

real-time polymerase chain reaction (qPCR). Ex vivo analysis<br />

of hepatic inflammatory cells was done by FACS. Results: In<br />

the MCD model, both Linagliptin and Sitagliptin lead to a significantly<br />

attenuated hepatic fat accumulation a reduction of<br />

transcript levels of SREBP-1c, FAS and LPL, lowered serum ALT,<br />

AST, ALP and LDH and significantly suppressed fibrosis and<br />

inflammation related transcripts (aSMA, procollagen α1(I),<br />

TIMP-1, TGFβ1, MMP-13, CD68, CCL3 and TNFα) compared<br />

to vehicle-treated controls. This was accompanied by a significant<br />

decrease of collagen area (Sirius red), αSMA, procollagen<br />

type III, CD68, F4/80, Caspase 3, as determined by semiquantitive<br />

immunohistochemistry. Boih DPPIV-inhibitors significantly<br />

decreased hepatic CD11b + Ly6C + high (proinflammatory monocytes/macrophages)<br />

and total F4/80 + cells (macrophages,<br />

Kupffer cells) compared to mice on the MCD diet without treatment.<br />

In nonsteatotic, spontaneously fibrotic Mdr2KO mice 10<br />

mg/kg/day of Linagliptin significantly decreased procollagen<br />

α1(I), TGFβ1, TIMP-1, MMP-8 transcript levels, and increased<br />

putatively anti-fibrotic MMP-9 and -13. Conclusion: In mice fed<br />

the MCD diet for 6 weeks, the DPPIV-inhibitors significantly<br />

decreased inflammation and apoptosis. In both the MCD and<br />

Mdr2KO model Linagliptin and Sitagliptin mildly decreased<br />

liver injury and fibrosis through inhibition of inflammatory and<br />

apoptosis pathways and via a decrease of proinflammatory<br />

and profibrotic monocytes-macrophages.<br />

Disclosures:<br />

Thomas Klein - Employment: Boehringer Ingelheim<br />

Detlef Schuppan - Consulting: Boehringer Ingelheim, Conatus, GLG, Merck,<br />

Mitsubishi-Tanabe, Takeda, Silence, Glenmark, Isis; Grant/Research Support:<br />

Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Xiaoyu Wang, Shih-Yen Weng,<br />

Yong Ook Kim


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 885A<br />

1377<br />

Branched-chain amino acids prevent hepatic fibrosis<br />

and the development of liver tumors by suppressing<br />

TGF-β signaling<br />

Kai Takegoshi 1 , Masao Honda 1 , Hikari Okada 1 , Riuta Takabatake<br />

1 , Takayoshi Shirasaki 1 , Naoto Matsuzawa 1 , Toshinari<br />

Takamura 1 , Takuji Tanaka 2 , Shuichi Kaneko 1 ; 1 Kanazawa University,<br />

Kanazawashi, Japan; 2 Tohkai Cytopathology Institute, Gifu,<br />

Japan<br />

BACKGROUND & AIMS Oral supplementation with branchedchain<br />

amino acids (BCAAs) in patients with liver cirrhosis<br />

potentially suppresses the incidence of hepatocellular carcinoma<br />

and improves event-free survival. However, the detailed<br />

mechanisms by which BCAAs act on hepatic fibrosis have not<br />

been elucidated fully. METHODS BCAAs were administered to<br />

atherogenic and high-fat (Ath&HF) diet-induced nonalcoholic<br />

steatohepatitis (NASH) model mice and platelet-derived growth<br />

factor C transgenic (Pdgf-c Tg) mice. Liver histology, tumor<br />

incidence, and GeneChip expression profiles were evaluated.<br />

To investigate the mechanisms of the effect of BCAAs, in vitro<br />

<strong>studies</strong> were performed using a human stellate cell line (Lx-2),<br />

rat liver progenitor cell-like cells (WB-F344), mouse primary<br />

hepatic stellate cells (MHSC), and mouse primary hepatocytes<br />

(MPH). RESULTS Mice fed the Ath&HF diet for 30 weeks developed<br />

liver fibrosis, whereas mice fed the Ath&HF diet containing<br />

BCAAs showed markedly reduced fibrosis. After 60 weeks,<br />

73.5% of mice fed the Ath&HF diet developed liver tumors,<br />

while BCAAs reduced tumor incidence to 30.8%. BCAAs also<br />

prevented liver fibrosis of Pdgf-c Tg mice and reduced tumor<br />

incidence. Gene expression analysis demonstrated the significant<br />

resolution of pro-fibrotic genes, pro-oncogenic genes, cancer<br />

stem cell markers (EpCAM and CD90), and impaired lipid<br />

metabolism gene expression by BCAA supplementation. In<br />

vitro, in stellate cells, BCAAs restored the transforming growth<br />

factor (TGF)-β1-stimulated expression of pro-fibrotic genes and<br />

inhibited the trans-differentiation of MHSC to myofibroblast-like<br />

cells. In hepatocytes, BCAAs prevented TGF-β1 and palmitate<br />

co-stimulated apoptosis signaling and restored impaired fatty<br />

acid β oxidation gene expression in MPH. Moreover, in progenitor<br />

cells, BCAAs reduced the TGF-β1-stimulated spheroid<br />

formation of WB-F344 cells. We found BCAAs inhibited TGF-β<br />

signaling by preventing the nuclear localization of the histone<br />

acetylating transcription factor NF-Y. Knockdown of NF-YA<br />

by small interfering RNA attenuated the TGF-β1-stimulated<br />

expression of p-AKT and p-Smad2/3. CONCLUSIONS BCAA<br />

supplementation prevents hepatic fibrosis and the development<br />

of liver tumors in the NASH mouse model and Pdgf-c Tg mice,<br />

possibly by suppressing TGF-β signaling through an interaction<br />

with NF-Y. These results highlight a new mechanism of the<br />

anti-fibrotic and anti-tumor effect of BCAAs in the liver and<br />

could be utilized for the treatment of chronic liver disease.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Kai Takegoshi, Masao Honda,<br />

Riuta Takabatake, Takayoshi Shirasaki, Naoto Matsuzawa, Toshinari Takamura,<br />

Takuji Tanaka<br />

1378<br />

Membrane type I-matrix metalloprotease-cleaved collagens<br />

derived from hepatic stellate cells play critical role<br />

in liver regeneration after partial hepatectomy in rats<br />

Akihiro Yoneda 1,2 , Miyuki Nishimura 2 , Naoko Urushihara 2,3 ,<br />

Tomoko M. Sudo 2,3 , Kenjiro Minomi 2,3 , Yasuaki Tamura 1 , Yoshiro<br />

Niitsu 2 ; 1 Department of Molecular Therapeutics, Center for Food &<br />

Medical Innovation, Institute for the Promotion of Business-Regional<br />

Collaboration, Hokkaido University, Sapporo, Japan; 2 Department<br />

of Molecular Target Exploration, School of Medicine, Sapporo<br />

Medical University, Sapporo, Japan; 3 Hokkaido Laboratory, Nitto<br />

Denko Corporation, Sapporo, Japan<br />

Background and Aims: The liver is well known to exhibit a<br />

remarkable ability to regenerate after partial heptatectomy<br />

(PHx). Liver regeneration is regulated by various events such as<br />

the remodeling of extracellular matrices (ECMs) and proliferation<br />

of hepatic cells stimulated by growth factors and cytokines.<br />

However, the precise molecular mechanism remains controversial.<br />

Since hepatic stellate cells (HSCs) are main producer of<br />

ECMs and also hepatic growth factor (HGF) in the damaged<br />

liver, HSCs has been postulated to play an important role in<br />

liver regeneration. We have recently demonstrated that membrane<br />

type I-matrix metalloprotease (MT1-MMP) of activated<br />

HSCs cleaved collagens secreted from themselves to expose<br />

RGD motif which interacted with integrins, thereby transduced<br />

survival signal and sustained their proliferation in an autocrine<br />

manner (Birukawa et al. J. Biol. Chem. 2014). The purpose of<br />

this study was to explore the possibility that MT1-MMP-cleaved<br />

collagens stimulate liver regeneration after PHx in a paracrine<br />

manner. Methods: Proliferations of hepatic cells after 70% PHx<br />

in rats were examined by immunohistochemistry and flow cytometric<br />

analysis. Hepatocytes isolated from EGFP-transgenic<br />

rats were co-cultured with activated HSCs. In vivo gene silencing<br />

system was performed using vitamin A-coupled liposome<br />

reported previously (Sato et al. Nature Biotechnol. 2008).<br />

Results: Proliferation of hepatocytes was coincide with activation<br />

of HSCs during the early phase of liver regeneration<br />

after 70% PHx. The proliferative hepatocytes localized in the<br />

vicinity of activated HSCs. Co-culture system demonstrated<br />

that activated HSCs stimulated the proliferation of hepatocytes<br />

while gene silencing of MT1-MMP or Hsp47, collagen-specific<br />

chaperone, in activated HSCs completely inhibited their<br />

proliferation. In vivo silencing of MT1-MMP or Hsp47 genes<br />

also resulted in remarkable suppression of liver regeneration<br />

after 70% PHx in rats. MT1-MMP-cleaved collagen I and HGF<br />

induced hepatocyte proliferation via the engagement of integrin<br />

b1 with c-Met tyrosine kinase receptor followed by the phosphorylation<br />

of c-Met and activation of focal adhesion kinase.<br />

We further found that proliferation of Sox9+ hepatic progenitor<br />

cells (HPCs) in close proximity to activated HSCs was transiently<br />

detected during the late phase of liver regeneration after<br />

70% PHx. Co-culture system demonstrated that activated HSCs<br />

induced the proliferation of HPCs while gene silencing of MT1-<br />

MMP or Hsp47 in activated HSCs completely inhibited their<br />

proliferation. Conclusion: MT1-MMP-cleaved collagens from<br />

HSCs play a pivotal role in liver regeneration after PHx.<br />

Disclosures:<br />

Miyuki Nishimura - Grant/Research Support: Nitto Denko<br />

The following authors have nothing to disclose: Akihiro Yoneda, Naoko Urushihara,<br />

Tomoko M. Sudo, Kenjiro Minomi, Yasuaki Tamura, Yoshiro Niitsu


886A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1379<br />

Selective antibody targeting of lysyl oxidase-like 2<br />

(LOXL2) suppresses hepatic fibrosis progression and<br />

accelerates its reversal via crosslinking-dependent and<br />

independent mechanisms<br />

Naoki Ikenaga 1 , Susan B. Liu 1 , Zhen-Wei Peng 1 , Shuhei Yoshida 1 ,<br />

Deanna Sverdlov 1 , Satyajit Karnik 2 , Amanda Mikels-Vigdal 2 , Victoria<br />

Smith 2 , Detlef Schuppan 1 , Yury Popov 1 ; 1 Gatroenterology,<br />

Beth Israel Deaconess Medical Center, Boston, MA; 2 Gilead Sciences,<br />

Inc, Foster City, CA<br />

BACKGROUND/AIMS: LOXL2 plays a role in collagen cross-linking<br />

and epithelial cell differentiation. We studied the role of<br />

LOXL2 in collagen cross-linking and hepatic progenitor cell<br />

(HPC) differentiation, and the therapeutic efficacy of anti-LOXL2<br />

mAb on liver fibrosis progression/reversal in mice. METHODS:<br />

Anti-LOXL2 antibody (AB0023 mAb, 30mg/kg), control antibody<br />

(M64, 30mg/kg) or placebo was administered i.p.<br />

twice a week during thioacetamide (TAA)-induced fibrosis progression<br />

(delayed treatment, week 6 to 12 of TAA) or during<br />

recovery (1-12 weeks off TAA). Therapeutic efficacy in biliary<br />

fibrosis was tested in BALB/c.Mdr2-/- and 3,5-diethoxycarbonyl-1,4-dihydrocollidine(DDC)-fed<br />

mice. Collagen cross-linking<br />

and fibrosis were assessed by histological and biochemical<br />

methods. HPC differentiation was studied in primary EpCAM(+)<br />

cells in vitro. RESULTS: LOXL2 was virtually absent from healthy<br />

but strongly induced in fibrotic livers, with predominant localization<br />

within fibrotic septa. Delayed anti-LOXL2 treatment<br />

of pre-established TAA-induced fibrosis suppressed collagen<br />

crosslinking, and histological signs of bridging fibrosis in the<br />

anti-LOXL2-treated group improved, with a 53% reduction in<br />

connective tissue deposition by morphometry. During recovery,<br />

anti-LOXL2 mAb promoted fibrosis regression, histological<br />

signs of fibrotic septa remodeling (splitting and thinning),<br />

and a 45% decrease in connective tissue area (p=0.021) at<br />

early recovery stages (4 weeks). In Mdr2-/- and DDC-induced<br />

models of biliary fibrosis, anti-LOXL2 mAb achieved significant<br />

antifibrotic efficacy but did not affect collagen cross-linking.<br />

Instead, ductular reaction and proliferation was suppressed<br />

in mice with LOXL2 inhibition, whereas hepatocyte replication<br />

increased. The LOXL2-blocking antibody had a profound effect<br />

on primary EpCAM(+) HPC in vitro, promoting their differentiation<br />

towards a hepatocyte lineage while inhibiting ductal cell<br />

lineage commitment. CONCLUSIONS: LOXL2 mediates collagen<br />

cross-linking and fibrotic matrix stabilization during liver<br />

fibrosis, and regulates hepatic progenitor cell differentiation. A<br />

therapeutic anti-LOXL2 antibody inhibits the progression of liver<br />

fibrosis and promotes fibrosis reversal.<br />

Disclosures:<br />

Satyajit Karnik - Employment: Gilead Sciences<br />

Victoria Smith - Employment: Gilead Sciences Inc<br />

Detlef Schuppan - Consulting: Boehringer Ingelheim, Conatus, GLG, Merck,<br />

Mitsubishi-Tanabe, Takeda, Silence, Glenmark, Isis; Grant/Research Support:<br />

Boehringer-Ingelheim<br />

Yury Popov - Grant/Research Support: Gilead Sciences, Inc, Takeda<br />

The following authors have nothing to disclose: Naoki Ikenaga, Susan B. Liu,<br />

Zhen-Wei Peng, Shuhei Yoshida, Deanna Sverdlov, Amanda Mikels-Vigdal<br />

1380<br />

IL-22 enhances TGF-beta pro-fibrotic function in hepatic<br />

stellate cells in a p38/MAPK dependent manner.<br />

Thomas Fabre 1,2 , Manuel Flores Molina 1,2 , Naglaa H. Shoukry 1,3 ;<br />

1 Centre de recherche du CHUM, Montrèéal, QC, Canada; 2 Département<br />

de microbiologie, immunologie et infectiologie, Université<br />

de Montréal, Montréal, QC, Canada; 3 Département de Médecine,<br />

Université de Montréal, Montréal, QC, Canada<br />

Background & Hypothesis: Activation of hepatic stellate<br />

cells (HSCs) is a key event in liver fibrosis, characterized by<br />

enhanced extracellular matrix (ECM) production and altered<br />

degradation. The immune system may modulate activation of<br />

HSCs through different cytokines. We have previously demonstrated<br />

that IL-17A enhances the expression of profibrotic genes<br />

through upregulation of the TGF-β receptor on hepatic stellate<br />

cells in a JNK-dependent manner (Fabre, Journal of Immunology<br />

2014). IL-22 is another Th17 enigmatic cytokine, from the<br />

IL-10 family, with both pro- and anti-inflammatory properties.<br />

IL-22 deficient mice develop high levels of hepatic inflammation<br />

during acute injury compared to their wild type littermates.<br />

In contrast, IL-22 is also elevated in the sera of patients with<br />

liver cirrhosis and carcinoma. However, in a mouse model<br />

of hepatitis B, IL-22 indirectly induces fibrosis by recruiting<br />

the pro-fibrotic inflammatory Th17 cells. We hypothesized that<br />

IL-22 may modulate activation and induction of the fibrogenic<br />

process in HSCs. Methods & Results: The human HSC line<br />

LX2 and primary human HSCs were stimulated with increasing<br />

doses of IL-22 and compared to TGF-β- and PBS- treated<br />

cells as positive and negative controls, respectively. IL-22 did<br />

not induce activation of HSCs. However, IL-22 enhanced the<br />

response of HSCs to suboptimal doses of TGF-β as observed by<br />

strong induction of alpha-smooth muscle actin (-SMA), collagen<br />

type I (COL1A1) and tissue inhibitor of matrix metalloproteinase<br />

(TIMP-I). IL-22 stimulation, in contrast to IL-17A, did not<br />

enhance cell surface expression of TGF-β-RII. To characterize<br />

the cellular mechanisms by which IL-22 enhances the TGF-β<br />

response we performed RNA-seq on primary human HSCs.<br />

However, pre-treatment of HSCs with IL-22 led to increase<br />

phosphorylation of SMAD2/3 in response to suboptimal doses<br />

TGF-β. Gene expression analysis revealed an increased in activation<br />

of p38/MAPK activity related pathways in IL-22 with<br />

TGF-β lo - as compared to PBS- or TGF-β lo treated cells. Activity<br />

was confirmed at the protein level and was associated with<br />

increased phosphorylation of SMAD2/3. Chemical inhibition<br />

of p38 significantly reduces HSCs activation in response to<br />

IL-22 with TGF-β lo . Conclusion: Our results suggest that IL-22<br />

enhances TGF-β signalling in a p38/MAPK dependent manner<br />

leading to increased activation of HSCs.<br />

Disclosures:<br />

The following authors have nothing to disclose: Thomas Fabre, Manuel Flores<br />

Molina, Naglaa H. Shoukry<br />

1381<br />

MeCP2 exerts global control over the myofibroblast<br />

transcriptome and reveals new regulators of fibrosis.<br />

EVA MORAN-SALVADOR, Pier P. Paoli, Graham R. Smith, Agata<br />

Page, Rachel M. Howarth, Fiona Oakley, Jelena Mann, Derek<br />

Mann; Institute of Cellular Medicine, Newcastle University, Newcastle<br />

upon Tyne, United Kingdom<br />

Background: Hepatic stellate cells (HSCs) are considered<br />

the central extracellular matrix (ECM)-producing cells in the<br />

wound-healing response process driven by a persistent liver<br />

injury. Previous <strong>studies</strong> in our laboratory identified MeCP2<br />

(methyl-CpG binding protein 2) as the epigenetics master reg-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 887A<br />

ulator of HSCs activation in liver fibrogenesis. Nowadays,<br />

the epigenetic machinery embraces an extensive network of<br />

non-coding RNAs (ncRNAs), such as long non coding RNAs<br />

(lncRNAs) and microRNAs (miRNAs). The transcription of<br />

ncRNAs establishes new regulatory systems affecting chromatin<br />

structure and gene expression. The aim of this study was to<br />

identify potential targets and a novel ncRNA regulatory network<br />

controlled by MeCP2 in HSCs. Materials and methods:<br />

RNA from Wt (n=3) and MeCP2-null (n=3) mice HSCs was<br />

subjected to a microarray based profiling analyses (Arraystar<br />

Mouse LncRNA Microarray V2.0 platform) and to a small RNA<br />

next-generation sequencing (NGS) (Illumina platform). Results:<br />

Using the microarray platform we found 161 upregulated and<br />

83 downregulated lncRNAs in MeCP2-null HSCs. We focused<br />

our attention to Gm3893 and AK080187; their deregulated<br />

expression was confirmed by qPCR and correlated with the<br />

expression of their associated gene transcripts (cytokines Ccl-<br />

21b/c and IL11ra2 receptor and the profibrogenic marker<br />

Acta2, respectively). Pathway analysis based on KEGGs<br />

database denoted that most of 284 downregulated mRNAs<br />

in MeCP2-null HSCs were significantly enriched at DNA replication<br />

and cell cycle pathways, while most of 124 upregulated<br />

mRNAs were associated with pathways involved in the<br />

immune system and metabolism. Most importantly, MeCP2-<br />

null HSCs presented a robust downregulation in five members<br />

of the MCM protein complex and a remarkable decrease in<br />

other components of the replisome multisubunit complex. The<br />

resulting altered pattern of mRNAs also emphasized the downregulation<br />

of the hyaluronan synthase 2 (Has2) upon loss of<br />

MeCP2 in HSCs. In liver, Has2 was predominantly expressed<br />

in RNA of HSCs and its protein expression increased over<br />

the HSCs transdifferentiation process. In primary HSCs, small<br />

interfering RNA for Has2 decreased the gene expression of<br />

ECM proteins. Of note, miRNA-328-3p pinpointed from the<br />

14 significantly altered miRNA identified by NGS analyses<br />

as has been proved to target and reduce the hyaluronic acid<br />

receptor CD44 expression. Conclusions: Our findings successfully<br />

evidence a vast ncRNA network controlled by MeCP2<br />

and highlight the complex role that MeCP2 elicits in the HSC<br />

transdifferentiating process.<br />

Disclosures:<br />

The following authors have nothing to disclose: EVA MORAN-SALVADOR, Pier P.<br />

Paoli, Graham R. Smith, Agata Page, Rachel M. Howarth, Fiona Oakley, Jelena<br />

Mann, Derek Mann<br />

1382<br />

Cartilage oligomeric matrix protein participates in the<br />

pathogenesis of liver fibrosis<br />

Fernando Magdaleno 1,2 , Elena Arriazu 2 , Marina Ruiz de Galarreta<br />

2 , Yu Chen 1 , Xiadong Ge 1,2 , Laura Conde de la Rosa 2 , Natalia<br />

Nieto 1,2 ; 1 Pathology, University of Illinois at Chicago, Chicago,<br />

IL; 2 Medicine, Icahn School of Medicine at Mount Sinai, New<br />

York, NY<br />

BACKGROUND & HYPOTHESIS: Liver fibrosis is characterized<br />

by excessive accumulation of extracellular matrix (ECM) proteins,<br />

mainly fibrillar collagen-I, as a result of persistent liver<br />

injury. Cartilage oligomeric matrix protein (COMP) is a glycoprotein<br />

typically found in the ECM of skeletal tissue. Increased<br />

COMP expression has been associated with fibrogenesis in<br />

systemic sclerosis, lung fibrosis and chronic pancreatitis. Since<br />

COMP is expressed in cirrhotic livers and it is significantly<br />

induced in hepatocellular carcinoma we hypothesized that<br />

COMP could induce fibrillar collagen-I protein deposition and<br />

participate in matrix remodeling contributing to the pathophysiology<br />

of liver fibrosis. METHODS: Chronic thioacetamide (TAA)<br />

administration (4 mos) and carbon tetrachloride (CCl 4<br />

) injection<br />

(1 mo) were used as models to induce liver fibrosis in<br />

wild-type (WT) littermates and in global Comp -/- mice. Controls<br />

received water or mineral oil, respectively. To dissect the mechanism<br />

for COMP signaling, in vitro experiments were carried<br />

out with primary rat hepatic stellate cells (HSCs) and qPCR,<br />

gelatin zymography or western blot analysis were performed.<br />

To determine if physical interaction occurred between COMP<br />

and collagen-I and how this conditioned collagen-I cleavage<br />

by matrix metalloproteinases (MMPs), in vitro reconstituted systems<br />

were used followed by immunoprecipitation and immunoblotting.<br />

RESULTS: COMP expression was detected in livers<br />

from control WT littermates and was up-regulated in response<br />

to chronic TAA- or CCl 4<br />

-induced liver injury. TAA-treated or<br />

CCl 4<br />

-injected Comp -/- mice showed less liver damage (ALT and<br />

AST activities, necrosis, inflammation and hepatocyte ballooning<br />

degeneration) and fibrosis compared to their matching<br />

control WT littermates. Challenge of HSCs with recombinant<br />

COMP (rCOMP) up-regulated intra- and extracellular collagen-I<br />

protein expression and increased MMP2, 9 and 13. However,<br />

rCOMP neither affected TGFβ protein, a well-known pro-fibrogenic<br />

factor, nor altered Tnfα nor tissue inhibitor of metalloproteinase-1<br />

(Timp1) mRNAs. Moreover, COMP showed major<br />

ability to bind collagen-I; yet, the binding did not preclude from<br />

cleavage by MMP1, the MMP with the highest affinity for collagen-I<br />

cleavage. Last, we identified that rCOMP significantly<br />

up-regulated collagen-I protein expression in HSCs via CD36<br />

receptor binding and activation of the MEK1/2-pERK1/2<br />

signaling pathway. CONCLUSION: These results suggest that<br />

COMP contributes to liver fibrosis by up-regulating collagen-I<br />

protein synthesis but not its cleavage by MMP1; thus, resulting<br />

in significant collagen-I deposition.<br />

Disclosures:<br />

The following authors have nothing to disclose: Fernando Magdaleno, Elena<br />

Arriazu, Marina Ruiz de Galarreta, Yu Chen, Xiadong Ge, Laura Conde de la<br />

Rosa, Natalia Nieto<br />

1383<br />

IL-10-producing regulatory B cells from chronic hepatitis<br />

B patients suppressed CD4+T cells proliferation and<br />

enhanced regulatory T cells function<br />

Li-sha Cheng, Yun Liu, Wei Jiang; Gastroenterology, Zhongshan<br />

hospital, Shanghai, China<br />

Background & aims: Chronic hepatitis B (CHB) infection is a<br />

leading cause of liver fibrosis, cirrhosis and even hepatocellular<br />

carcinoma. Although hepatitis B virus (HBV) itself is non-cytopathic,<br />

both HBV-related liver damage and viral control are<br />

immune-mediated. There were abundant evidences to believe<br />

that regulatory T (Treg) cells play pivotal roles in the immunopathogenesis<br />

of HBV-related liver diseases. Nowadays, a<br />

new subset of IL-10-producing B cells has been shown to play<br />

vital roles in modulating autoimmune diseases. Here, we evaluated<br />

the role of regulatory B (Breg) cells in the pathogenesis<br />

of HBV-related liver fibrosis (HBV-LF) and assess their impact<br />

on the proliferation of CD4+T cells the immunoregulatory<br />

effects on Treg cells. Methods: A total of 67 treatment-naïve<br />

patients who were diagnosed as chronic HBV (CHB) and<br />

25 healthy donors took part in the study. Firstly, we determined<br />

the changes in number and function of Breg cells in<br />

CHB patients compared with that of healthy controls. For the<br />

in-vitro experiments, purified CD4+T cells were cultured alone<br />

or with autologous Breg cells and their proliferation response<br />

was determined by the thymidine method. Simultaneously, to<br />

elucidate the exact effects of Bregs on Tregs and the potential<br />

methanism, human Breg and Treg cells were co-cultured in the


888A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

stimulation with HBcAg, anti-IL-10 and anti-TGF-β antibody.<br />

Results: B cells and Breg cells were both enriched in patients,<br />

Breg cells frequencies had a close relationship with viral load,<br />

liver inflammation and degree of fibrosis. Stimultaneously, we<br />

demonstrated the phenotype of these Breg cells: memory B cells<br />

(CD24 hi CD27 + B cell)mature B cells(CD24 int CD38 int B cell) and<br />

transitional B cells (CD24 hi CD38 hi B cell), and the phenotype<br />

was pridominantly characterized as CD19 + CD24 hi CD38 hi B<br />

cell. In the co-culture of Breg cells from CHB patients with autologous<br />

CD4 + T cells, the proliferation capacity of CD4 + T cells<br />

was significantly reduced in a dose-dependent manner compared<br />

with controls. In addition, Breg cells from patients were<br />

found to have further enhanced effect on the proliferation of<br />

Treg cells as well as the release of IL-10 and TGF-β cytokines<br />

than such cells from controls. Furthermore, blockage of IL-10<br />

and CTLA-4 but not TGF-β decreased the inhibitory effect of<br />

Breg cells on Treg cells. The regulatory function of Breg cells on<br />

Treg cells was dependent on direct cell-cell contact as well as<br />

IL-10 but not TGF-β-dependent manner. Conclusions: Our data<br />

revealed that elevated Breg cells, which can inhibit the proliferation<br />

of CD4+T cells and regulate Treg cells immunitymaybe<br />

an indicator for the development of HBV-related liver fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Li-sha Cheng, Yun Liu, Wei Jiang<br />

1384<br />

SNAP-23 Heterozygous Mice Are Resistant To Experimental<br />

Fibrosis<br />

Haleigh B. Eubanks, Jessica R. Goree, Michel Fausther, Jonathan<br />

A. Dranoff; Gastroenterology and Hepatology, University of<br />

Arkansas for Medical Sciences, Little Rock, AR<br />

Background. Upon liver injury, quiescent hepatic stellate cells<br />

(HSC) become “activated” and converted to myofibroblasts.<br />

Hepatic stellate cell derived myofibroblasts possess characteristics<br />

such as increased proliferation capability and collagen<br />

synthesis. Several molecular targets have been identified<br />

in liver fibrosis. Most current therapies are aimed towards<br />

inhibiting or impeding accumulation of fibrogenic cells or<br />

deposition of extracellular matrix (ECM) proteins. However,<br />

the general molecular mechanism and proteins that function<br />

in aiding excessive ECM secretion remain to be elucidated.<br />

SNARE (soluble N-ethlmaleimide sensitive fusion attachment<br />

protein receptor) proteins have been proposed to be a conserved<br />

group of machinery that mediate intracellular trafficking<br />

and exocytosis. We found that HSC lack the t-SNARE protein,<br />

synaptosomal associated protein-25 (SNAP-25), but express a<br />

related protein, SNAP-23. Specific Aims. We are testing the<br />

hypothesis that SNAP-23 is essential for matrix release by HSC<br />

myofibroblasts by testing whether mice lacking one allele of the<br />

SNAP-23 gene are resistant to experimental fibrosis. (Of note,<br />

the SNAP-23 knockout mouse is embryonically lethal). Methods.<br />

Wild type and SNAP-23 heterozygous mice were treated<br />

with CCl4 or olive oil (control) for two (early fibrosis) and six<br />

weeks (late fibrosis/early cirrhosis). Fibrosis in liver tissue was<br />

assessed by Trichrome and Sirius red staining. Assessment of<br />

the expression of collagen, tissue inhibitor of metalloproteinase-1<br />

(TIMP-1), and alpha-smooth muscle actin (α-SMA) was<br />

determined by qRT-PCR. At the protein level, immunofluorescence<br />

and western blot were used to assess α-SMA secretion.<br />

Results. SNAP-23 +/- mice receiving CCl4 were significantly<br />

less fibrotic than WT mice; moreover, no SNAP-23 +/- mice<br />

were cirrhotic at the six week time point. Collagen quantification<br />

showed similar amounts of collagen content in WT and<br />

SNAP-23 +/- mice. While there was no difference in collagen<br />

or alpha-SMA mRNA expression between WT and SNAP-23<br />

+/- mice, SNAP-23 +/- mice evidenced decreased collagen<br />

and alpha-SMA protein. Conclusion. Single allelic deletion of<br />

SNAP-23 protected mice from experimental fibrosis/cirrhosis,<br />

likely via inhibition of exocytosis of matrix proteins. The reduction<br />

in alpha-SMA protein in heterozygote mice may reflect<br />

impaired HSC survival when the SNARE complex is disrupted.<br />

Disclosures:<br />

The following authors have nothing to disclose: Haleigh B. Eubanks, Jessica R.<br />

Goree, Michel Fausther, Jonathan A. Dranoff<br />

1385<br />

GP38/PODOPLANIN marks a novel subset of progenitor<br />

cells in chronic liver injury<br />

Henrike Julich 1 , Christoph Eckert 1 , Yong Ook Kim 2 , Hannes<br />

Borchardt 1 , Niklas Klein 1 , Eva-Carina Heier 1 , Thomas Tschernig 3 ,<br />

Detlef Schuppan 2 , Veronika Lukacs-Kornek 1 ; 1 Department of<br />

Medicine II, Saarland University, Homburg, Germany; 2 Institute<br />

of Translational Immunology and Research Center for Immunotherapy,<br />

Institute of Translational Immunology, University Medical<br />

Center, Johannes Gutenberg University, Mainz, Germany; 3 Insitute<br />

of Anatomy and Cell Biology, Saarland University, Homburg,<br />

Germany<br />

BACKGROUND/AIMS: Gp38/podoplanin + stromal cells present<br />

in lymphoid organs play central role in the formation and<br />

reorganization of the extracellular matrix providing the three<br />

dimensional scaffold for immune responses. Additionally,<br />

gp38 + stromal cells are crucial in the functional regulation of T<br />

cells during immune responses (V. Lukacs-Kornek Nat Immunol<br />

2011 12(11):1096-104). Gp38 + cells have been identified<br />

during embryogenesis and more recently in human livers of primary<br />

biliary cirrhosis. Since, little is known about the function<br />

of these cells, we aimed to study the phenotypic and functional<br />

characteristics of gp38 + cells during chronic liver inflammation.<br />

METHODS: Gp38 + stromal cells have been analyzed in models<br />

of biliary fibrosis (MDR2 KO), CCL 4<br />

induced liver fibrosis<br />

(both during progression and regression) and in a model of<br />

steatohepatitis (methionine choline deficient (MCD) diet). Stromal<br />

cells were analyzed via flow cytometry and via confocal<br />

microscopy. Moreover, various subsets of stromal cells were<br />

sorted with high purity (>95%) for gene expression analyses<br />

and gp38 + stromal cells were tested functionally in in vitro<br />

assays. RESULTS: Gp38 + stromal cells are present in the healthy<br />

liver and significantly expand in all three models of liver injury.<br />

During regression of inflammation and fibrosis, the presence<br />

of these cells returns to the levels of healthy animals. Gp38 +<br />

cells localized primarily in the portal area and occasionally<br />

imposed as round ductular structures. Using multicolor FACS<br />

analysis, two subsets of gp38 + stromal cells could be identified:<br />

gp38 hi and gp38 low . The gp38 hi cells expressed multiple<br />

mesenchymal markers while the gp38 low cells could be further<br />

divided using CD133 and EPCAM labeling. Importantly, the<br />

gp38 low subset showed surface markers similar to liver progenitor<br />

(oval) cells, which was confirmed by qPCR array analyses.<br />

Importantly, gp38 low cells displayed inflammatory genes<br />

during liver injury different from classical oval cells identified as<br />

CD45 - gp38 - CD133 + cells. CONCLUSIONS: 1.Gp38 + cells are<br />

present in healthy liver. 2. Gp38 + stromal cells expand during<br />

liver injury. 3. Multiple subsets of gp38 + stromal cells could<br />

be distinguished: the gp38 low cells represent a novel oval cell<br />

subset with a distinct inflammatory gene expression compared<br />

to classical oval cells which after further characterization may<br />

represent a novel therapeutic target for liver inflammation and<br />

fibrosis. Studies were supported by the Alexander von Humboldt<br />

Foundation – Sofja Kovalevskaja Award 2012 to VLK,<br />

and an ERC Advanced Grant to DS.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 889A<br />

Disclosures:<br />

Detlef Schuppan - Consulting: Boehringer Ingelheim, Conatus, GLG, Merck,<br />

Mitsubishi-Tanabe, Takeda, Silence, Glenmark, Isis; Grant/Research Support:<br />

Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Henrike Julich, Christoph Eckert,<br />

Yong Ook Kim, Hannes Borchardt, Niklas Klein, Eva-Carina Heier, Thomas<br />

Tschernig, Veronika Lukacs-Kornek<br />

1386<br />

Characterization of obesity-related modifications in<br />

hepatic stellate cells<br />

Guanhua Xie 1 , Marzena Swiderska-Syn 1 , Mariana V. Machado 1 ,<br />

Mark Jewell 1 , Richard T. Premont 1 , Gregory A. Michelotti 1 , Raffaele<br />

Teperino 2 , Anna Mae Diehl 1 ; 1 Division of Gastroenterology,<br />

Duke University Medical Center, Durham, NC; 2 Institute of Experimental<br />

Genetics (IEG) - HDC, Neuherberg, Germany<br />

Background: Obesity is an independent risk factor for<br />

advanced fibrosis in older NAFLD patients. Paradoxically, ob/<br />

ob mice with genetic leptin deficiency are obese but protected<br />

from liver fibrosis. Hedgehog-responsive myofibroblasts (MF)<br />

derived from hepatic stellate cells (HSC) are the major fibrogenic<br />

cells in obesity-related cirrhosis. The anti-obesogenic hormone,<br />

leptin, promotes transdifferentiation of HSC into MF.<br />

Aim: To determine if genetic leptin deficiency induces inherent<br />

differences in HSC that account for the “fibrosis-protected”<br />

liver phenotype in ob/ob mice. Methods: Primary HSCs were<br />

isolated from wild type (WT) and ob/ob mice by in situ pronase/collagenase<br />

perfusion followed by histodenz density gradient<br />

centrifugation. To exclude the possibility of hepatocyte<br />

contamination, some experiments were repeated in ultrapure<br />

HSCs obtained by additional retinoid-based FACS sorting.<br />

mRNA and/or protein expression of myofibroblastic markers<br />

and Hedgehog (Hh) signaling components/targets were compared<br />

by qRT PCR and FACS analysis when HSCs were either<br />

freshly isolated or cultured for 7 days to induce myofibroblastic<br />

transition. Scratch and Brdu assays were used to examine the<br />

migration and proliferation of HSCs. Studies were repeated<br />

with HSC harvested from leptin-treated ob/ob mice, and with<br />

cultured ob/ob HSC that were treated with leptin or Sonic<br />

Hedgehog (SHh) ligand. Results: Ob/ob HSCs were less myofibroblastic<br />

(e.g., ≥ 50% reduced mRNA and protein expression<br />

of α smooth muscle actin, collagen 1α1, and platelet-derived<br />

growth factor) and had inhibited Hh pathway activity (e.g.,<br />

less patched 1, Gli1, Gli2, and SHh at mRNA and/or protein<br />

level) compared to WT HSCs, both when quiescent and culture-activated.<br />

In contrast, a 2-fold greater elevated expression<br />

of quiescent HSCs markers (peroxisome proliferator-activated<br />

receptor γ and glial fibrillary acidic protein) were found in<br />

ob/ob HSCs. Functional analysis by Brdu assay and migration<br />

assay also showed that ob/ob HSCs were significantly<br />

less proliferative and migratory (p


890A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

would use human liver, or a biologically relevant 3D in vitro<br />

model of human liver cells grown on an extracellular matrix<br />

(ECM) derived from healthy or cirrhotic liver tissue. Aim: The<br />

aim of this study was to develop a rapid protocol for the decellularisation<br />

of small samples of healthy and cirrhotic human<br />

liver and demonstrate repopulation with cultured human liver<br />

cell lines, namely hepatocarcinoma (HepG2) and hepatic stellate<br />

(LX2) cells. Methods: Liver tissue cubes (125mm 3 ) were<br />

dissected from human livers unsuitable for transplantation or<br />

explanted liver. The decellularisation of the liver scaffolds was<br />

completed within 3hrs (healthy) and 5hrs (cirrhotic) of agitation<br />

in decellularization solutions i.e. detergents/enzymes. The<br />

decellularization efficiency was determined by immunohistochemistry<br />

for ECM components and residual DNA, scanning<br />

electron microscopy, biomechanics, as well as DNA and ECM<br />

protein quantification. The liver scaffolds were repopulated by<br />

HepG2 and LX2 for up to 14 days. RESULTS: This innovative<br />

protocol resulted in liver scaffolds with a preserved 3D structure<br />

and ECM composition, while DNA and cellular residues<br />

were successfully removed. The cirrhotic liver structure was<br />

maintained after decellularization and the biomechanical properties<br />

were remarkably different when comparing healthy liver<br />

scaffolds with cirrhotic scaffolds. Liver scaffolds were progressively<br />

repopulated for up to 14 days with LX2 and HepG2 cells<br />

which showed remarkable viability, motility and proliferation<br />

associated with remodelling effects on the surrounding ECM.<br />

Notably, the expression of several genes and proteins involved<br />

in liver fibrosis and cancer was different between the healthy<br />

and cirrhotic 3D-system. CONCLUSION: This is the first report<br />

describing an efficient protocol to completely decellularize<br />

human liver scaffolds obtained from healthy and cirrhotic liver.<br />

This decellularization protocol maintained the natural 3D-structure<br />

and ECM composition and organisation of both healthy<br />

and cirrhotic human liver tissue. This is a key advance in the<br />

development of 3D-technologies for the study of the progression<br />

of human liver fibrosis into cirrhosis and hepatocellular<br />

carcinoma.<br />

Disclosures:<br />

Kevin Moore - Advisory Committees or Review Panels: Servier<br />

Massimo Pinzani - Advisory Committees or Review Panels: Intercept Pharmaceutical,<br />

Silence Therapeutic, Abbot; Consulting: UCB; Speaking and Teaching:<br />

Gilead, BMS<br />

The following authors have nothing to disclose: Giuseppe Mazza, Lisa Longato,<br />

Walid Al-Akkad, Andrea Telese, Luca Urbani, Andrew R. Hall, Benjamin Robinson,<br />

Luca Frenguelli, Oliver Willacy, Marco Curti, Domenico Tamburrino, Gabriele<br />

Spoletini, Massimo Malago, Tu Vinh Luong, Armando E. Del Rio Hernandez,<br />

Paolo De Coppi, Krista Rombouts<br />

1389<br />

Rev-erb and TGF-β Differentially Regulate Autophagy in<br />

Hepatic Stellate Cells<br />

Paul G. Thomes 1 , Nicole A. Feilen 1 , Jennifer H. Benbow 1 , Elizabeth<br />

Brandon-Warner 1 , Cathy Culberson 1 , Terrence M. Donohue 2 ,<br />

Laura W. Schrum 1 ; 1 Liver Pathobiology Laboratory, Carolinas<br />

Medical Center, Charlotte, NC; 2 Omaha VA Medical Center &<br />

Univ. of Nebraska Medical Center, Omaha, NE<br />

Background: Recently, we demonstrated that ligand-activated<br />

nuclear receptor Rev-erbα mitgates the fibrogenic phenotype of<br />

hepatic stellate cells (HSCs) (Li et al., Hepatology, 2014). Reverbα<br />

is also a novel regulator of macroautophagy (autophagy)<br />

(Woldt et al., Nat Med, 2013), a crucial lysosomal degrading<br />

system that likely provides substrates for HSC activation. Here,<br />

we examined whether pharmacological activation of Rev-erb<br />

with SR9009 or treatment with the pro-fibrotic cytokine, TGFβ,<br />

each differentially modulates autophagy in activated rat<br />

primary HSCs and in HSC or fibroblast cell lines. Methods:<br />

We treated activated human HSC cell lines, LX2 and TWNT-4,<br />

primary rat hepatic stellate cells (rHSCs) or mouse embryonic<br />

fibroblast cells (3T3) with TGF-β (5 ng/ml) or Rev-erb ligand,<br />

SR9009 (10 mM), for 24 hours. Autophagy was quantified<br />

by protein markers and autophagosome (AV) flux by Western<br />

blot analyses (WB) and immunohistochemistry (IHC). Results:<br />

IHC demonstrated decreased AV following TGF-β exposure<br />

in all cell types. Rev-erb activation with SR9009 decreased<br />

AV in TWNT-4, 3T3 and rHSCs but not in LX2 cells. WBs<br />

confirmed lower levels of AV marker protein, LC3II, in all TGFβ-<br />

treated cells even after 4hr of cytokine exposure. In LX2,<br />

TWNT-4 and rHSCs, TGF-β simultaneously decreased levels of<br />

P62, an adaptor protein whose levels reflect the degree of AV<br />

degradation. SR9009-treated LX2, TWNT-4 and 3T3 cells had<br />

decreased LC3II protein levels and an insignificant decrease<br />

in rHSCs. P62 levels were unaffected by SR9009 exposure<br />

in all cell types. TGF-β treatment of rHSC was associated with<br />

higher lysosome (Lys) (LAMP1) numbers and a higher co-localization<br />

frequency of AVs with Lys. In contrast, SR9009 treatment<br />

caused reduced Lys and AV-Lys co-localization. In rHSCs,<br />

SR9009 but not TGF-β, decreased lysosomal cathepsin B activity.<br />

When cells were co-treated with TGF-β and bafilomycin<br />

(blocks AV degradation), rHSCs exhibited numerically higher<br />

LC3II levels than cells treated with bafilomycin alone. Rat HSCs<br />

treated with SR9009 and bafilomycin exhibited significantly<br />

lower LC3II levels than cells treated with bafilomycin alone.<br />

In rHSC, TGF-β and the autophagy inducers, rapamycin and<br />

starvation decreased Rev-erbα nuclear staining. SR9009 treatment<br />

decreased Rev-erbα cytoplasmic staining. Conclusion:<br />

Our findings indicate that TGF-β accelerates AV degradation<br />

and SR9009 retards AV synthesis and degradation in activated<br />

HSCs. We propose that Rev-erbα is a transcriptional regulator<br />

of autophagy in HSCs and that its ligand, SR9009, slows autophagic<br />

degradation of the macromolecular substrates that fuel<br />

fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Paul G. Thomes, Nicole A. Feilen,<br />

Jennifer H. Benbow, Elizabeth Brandon-Warner, Cathy Culberson, Terrence M.<br />

Donohue, Laura W. Schrum<br />

1390<br />

Molecular mechanism responsible for tissue stiffness in<br />

advanced chronic liver fibrogenesis<br />

Takao Sakai; Department of Molecular and Clinical Pharmacology,<br />

Institute of Translational Medicine, University of Liverpool,<br />

Liverpool, United Kingdom<br />

Tissue fibrosis is a part of wound-healing response that maintains<br />

organ structure and integrity following tissue damage<br />

and characterized by extracellular matrix (ECM) remodeling<br />

and stiffening. However, functional contribution of tissue stiffening<br />

to non-cancer pathogenesis remains largely unknown. In<br />

particular, how remodeling of ECM by myofibroblasts results<br />

in changes in mechanical tension that support the activation<br />

of pathogenic signaling pathways remain to be elucidated.<br />

Fibronectin is an ECM glycoprotein substantially expressed<br />

during adult tissue repair. We have addressed the molecular<br />

mechanism responsible for tissue stiffness in advanced chronic<br />

liver fibrogenesis induced by carbon tetrachloride (16 weeks)<br />

using a mouse model lacking fibronectin in the adult liver.<br />

Fibronectin-null livers exhibited constitutively elevated local<br />

TGF-β activity, induced more myofibroblast phenotypes, and<br />

accumulated highly disorganized/diffuse collagenous ECM<br />

networks composed of thinner and significantly increased number<br />

of collagen fibrils during advanced chronic liver damage.<br />

Consequently, fibronectin-null livers lead to more extensive liver


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 891A<br />

cirrhosis, which was accompanied by significantly increased<br />

liver tissue stiffness and deteriorated hepatic functions. Mechanistically,<br />

mutant livers showed elevated lysyl oxidase expressions,<br />

and a significant amount of active lysyl oxidase was<br />

released in fibronectin-null hepatic stellate cells in response<br />

to TGF-β1 through both canonical Smad- and non-canonical<br />

such as PI3 kinase-mediated pathways. Furthermore, inhibition<br />

of lysyl oxidase resulted in the disruption of TGF-β-induced<br />

fine collagen fibril networks and significantly reduced<br />

formed collagen fibril stiffness, and treatment of fibronectin in<br />

fibronectin-null stellate cells recovered collagen fibril stiffness to<br />

wild-type levels. Thus, our novel findings demonstrate an indispensable<br />

role for fibronectin in chronic liver fibrosis/cirrhosis<br />

in negatively regulating TGF-β bioavailability, which in turn<br />

modulates ECM remodeling and stiffening, and consequently<br />

preserves adult tissue functions.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takao Sakai<br />

1391<br />

Antifibrogenic properties of monoacylglycerol lipase<br />

inhibitors<br />

Aida Habib 1 , Jinghong Wan 1 , Pushpa Hegde 1 , Emmanuel Weiss 1 ,<br />

Arthur Brouillet 2 , Richard Moreau 1 , Sophie Lotersztajn 1 ; 1 Inserm<br />

U1149, Center for Research on inflammation, Paris, France;<br />

2 Inserm U955, Paris, France<br />

Background and aims Monoacylglycerol lipase (MAGL) is the<br />

rate-limiting enzyme in the degradation of monoacylglycerols.<br />

In addition to its role in lipid metabolism, MAGL is a pivotal<br />

component of the endocannabinoid system, since this enzyme<br />

metabolizes 2-arachydonoyl-glycerol, an endogenous cannabinoid<br />

receptor ligand, into arachidonic acid. Recent <strong>studies</strong><br />

have identified inhibition of MAGL as an interesting anti-inflammatory<br />

strategy in several experimental models of chronic<br />

inflammatory diseases. In the present study, we investigated<br />

the impact of MAGL inhibition on inflammation and fibrosis<br />

regression. Methods: Liver fibrosis regression was studied in<br />

C57BL/6 mice exposed to chronic administration of carbon<br />

tertrachloride (CCl 4<br />

) for 6 weeks. Injections were discontinued<br />

and mice were daily administered with the MAGL inhibitor<br />

JZL184 (15 mg/k, i.p.) or vehicle for up to 4 days. In vitro<br />

<strong>studies</strong> were performed on isolated murine peritoneal macrophages<br />

and hepatic myofibroblasts, and on PBMC from alcoholic<br />

cirrhotic and healthy subjects. Sirius red staining was<br />

quantified by morphometric analysis, fibrogenic and inflammatory<br />

gene expression by RT-PCR analysis, and the identity<br />

of intrahepatic leucocytes was analyzed by FACS. Results:<br />

JZL184 accelerated fibrosis regression 4 days after cessation<br />

of CCl 4<br />

administration, as shown by a reduction of sirius red<br />

staining and smooth muscle alpha actin immunolabeling. JZL<br />

184 also decreased the expression of fibrogenic genes in<br />

treated vs untreated animals. Moreover, JZL184 significantly<br />

decreased the influx of Ly6C + infiltrating monocytes into the<br />

liver of CCl4-exposed mice. Accordingly, JZL 184 reduced the<br />

hepatic expression of Ly6C mRNA and decreased the hepatic<br />

mRNA expression of inflammatory genes including IL1-b, CCl2,<br />

and CCL4. In vitro <strong>studies</strong> demonstrated that JZL184 reduced<br />

hepatic myofibroblast proliferation and down-regulated<br />

LPS-stimulation of inflammatory gene expression in peritoneal<br />

macrophages. Finally, MAGL mRNA expression was increased<br />

in PBMC from patients with alcoholic cirrhosis as compared to<br />

healthy subjects and further enhanced in LPS-exposed PBMC.<br />

Conclusions: Pharmacological inhibition of MAGL accelerates<br />

liver fibrosis regression in a model of toxin-induced liver injury,<br />

most probably by a mechanism involving reduction of infiltrating<br />

pro-inflammatory monocytes into the liver and a decrease<br />

in inflammatory mediators produced by macrophages. These<br />

results unravel MAGL inhibition as a novel promising antifibrogenic<br />

approach.<br />

Disclosures:<br />

The following authors have nothing to disclose: Aida Habib, Jinghong Wan,<br />

Pushpa Hegde, Emmanuel Weiss, Arthur Brouillet, Richard Moreau, Sophie Lotersztajn<br />

1392<br />

Induction of XBP1 Leads to Hepatic Stellate Cell Activation<br />

Rosa S. Kim, Daisuke Hasegawa, Xiaochen Sun, Takuma<br />

Tsuchida, Dipankar Bhattacharya, Hsin I Chou, David Y. Zhang,<br />

Yujin Hoshida, Scott L. Friedman; Liver Diseases, Icahn School of<br />

Medicine at Mount Sinai, New York, NY<br />

BACKGROUND AND AIM: We have previously established<br />

that ER stress leading to the Unfolded Protein Response (UPR)<br />

promotes hepatic stellate cell (HSC) activation in liver injury<br />

(Hernandez-Gea, V., et al., J Hepatol, 2013. and Hasegawa,<br />

D., et al., PLoS ONE 2010.). UPR has three effector pathways,<br />

PERK, ATF6 and IRE1-XBP1. IRE1 splices uXBP1 into its active<br />

transcriptional effector, sXBP1. Our aim was to define the<br />

relative contribution of XBP1 to HSC activation. METHODS:<br />

We overexpressed XBP1 splice isoforms<br />

by lentiviral infection for<br />

96 hours in primary human HSCs and immortalized human<br />

HSC lines (TWNT4, LX2). Fibrogenic targets were assessed by<br />

qPCR and immunoblot (collagen I, αSMA, βPDGFR, MMP2,<br />

TIMP1). Genome-wide transcriptome profiling was performed<br />

using HumanHT-12 beadarrays (Illumina) to define changes in<br />

global gene expression and pathway engagement downstream<br />

of XBP1 splice isoforms<br />

. Stringent criteria (FDR


892A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1393<br />

Role of CREBH activation induced by HCV infection in<br />

up-regulation of TGF-β2 expression and fibrogenesis<br />

Takeshi Chida 1,2 , Masahiko Ito 1 , Kazuhito Kawata 2 , Yoshimasa<br />

Kobayashi 2 , Tetsuro Suzuki 1 ; 1 Infectious disease, Hamamatsu university<br />

school of medicine, Hamamatsu, Japan; 2 2nd department<br />

of internal medicine, Hamamatsu university school of medicine,<br />

Hamamatsu, Japan<br />

Background and Aim: The mechanisms by which infection<br />

with hepatitis C virus (HCV) modulates the process of liver<br />

fibrosis still remains poorly defined. Increased expression of<br />

pro-fibrogenic growth factors including transforming growth<br />

factor-β (TGF-β) and extra cellular matrix proteins has been<br />

correlated with the degree of fibrosis in the liver. The aim of<br />

this study was to elucidate the mechanism of increased expression<br />

of TGF-β mediated by HCV infection. Methods: Fibrogenic<br />

responses caused by HCV (J6/JFH-1 isolate) infection was<br />

assessed using co-culture system. Hepatocellular carcinoma<br />

Huh7 cells infected with HCV were co-cultured with hepatic stellate<br />

TWNT4 cells. TGF-β promoter activities were analyzed by<br />

the luciferase reporter assay. Expression of mRNAs of fibrotic<br />

markers was analyzed by real-time RT-PCR. Protein expression<br />

was assessed by Western blotting and immunofluorescence.<br />

Knockdown and knockout for targeted genes were performed<br />

using siRNA and CRISPR-Cas9 systems, respectively. Results:<br />

Expression of TGF-β1 and TGF-β2 was significantly higher<br />

in HCV-infected cells than in non-infected cells; in particular,<br />

gene expression and promoter activity of TGF-β2 were markedly<br />

increased in cells with HCV infection or expression of the<br />

viral polyproteins such as Core through NS2. Compared to<br />

HCV-infected Huh7 alone or TWNT4 alone, co-culture of the<br />

infected Huh7 and TWNT4 cells showed significantly higher<br />

expression of collagen type1A1 (COL1A1). The upregulation<br />

of COL1A1 was suppressed by TGF-β2 knockout in Huh7 cells.<br />

From mutagenesis together with search for transcription factor<br />

binding sites, two regions; previously-identified CRE and<br />

temporal “CRE binding protein, hepatocyte-specific (CRE-<br />

BH)-response” element (CREBHRE) within the proximal TGF-β2<br />

promoter are important for its transcription regulation. Gel shift<br />

assay showed CREBH binding to the regions. TGF-β2 promoter<br />

activity was decreased by CREBH knockdown or knockout, and<br />

was increased by expression of the active form of CREBH. ChIP<br />

assay showed that binding of CREBH and ATF2 to TGF-β2 promoter<br />

was increased when the HCV Core-NS2 was expressed,<br />

under which precursor CREBH was processed into its active<br />

form. Discussion: CREBH, an endoplasmic reticulum (ER)-localized,<br />

liver-enriched transcriptional factor, is well-known to be<br />

activated by ER-stress. Our findings suggest that the expression<br />

of HCV proteins in the virus-infected cells potentially induces<br />

ER-stress and CREBH activation. Increased expression of TGFβ2<br />

by CREBH activation and its release from the infected cells<br />

contributes to the induction of fibrogenic responses via paracrine<br />

signaling to nearby hepatic stellate cells.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takeshi Chida, Masahiko Ito,<br />

Kazuhito Kawata, Yoshimasa Kobayashi, Tetsuro Suzuki<br />

1394<br />

Expression of ENTPD1/CD39 is protective in a mouse<br />

model of biliary fibrosis<br />

Linda Feldbrügge 1,3 , Shuji Mitsuhashi 1 , Eva Csizmadia 1 , Xiaofeng<br />

Sun 1 , Moritz Schmelzle 2 , Simon C. Robson 1 ; 1 Gastroenterology,<br />

BIDMC, Boston, MA; 2 Department for Surgery, Charité, Berlin,<br />

Germany; 3 Department for Surgery, University Hospital Leipzig,<br />

Leipzig, Germany<br />

Introduction. Ecto-nucleoside triphosphate diphosphohydrolases<br />

(ENTPD) comprise a family of cell surface located<br />

transmembrane proteins that regulate purinergic signaling by<br />

catalyzing extracellular nucleotides, such as ATP, ADP, to ultimately<br />

generate adenosine. These mediators are important signaling<br />

molecules in hepatic injury and inflammation. ENTPD1/<br />

CD39 is expressed on endothelium, and sinusoidal immune<br />

cells, and closely regulates liver regeneration by generating<br />

AMP from ATP and ADP. ENTPD2/CD39L1 is expressed on<br />

portal fibroblasts and perivascular cells and is a preferential<br />

ecto-ATPase, generating ADP. The role of ENTPD-mediated<br />

catalysis in modulating liver and biliary fibrosis is currently<br />

unclear as both ATP and adenosine could promote fibrogenic<br />

signals in portal myofibroblast and stellate cells and ADP is a<br />

potent platelet agonist. Methods. C57BL6 wild type, ENTPD1<br />

null mice and ENTPD2 null mice were subjected to bile duct<br />

ligation (BDL). Blood and liver tissue were harvested at 2 and<br />

4 weeks after BDL. Liver function, tissue injury and extent of<br />

cholestasis were determined by serum liver function tests and<br />

histopathology. To assess the extent of biliary fibrosis, we also<br />

analyzed tissue hydroxyproline content and performed Masson’s<br />

trichrome staining. Results. We observe a trend towards<br />

decreased survival in the ENTPD1 null mice after BDL. ENTPD1<br />

null mice further show significantly more weight loss, develop<br />

more pronounced cholestasis, and more severe liver fibrosis<br />

than wild type mice. No major profibrogenic effects are noted<br />

in ENTPD2 null mice, as determined by hydroxyproline content<br />

and morphometric analysis of peribiliary collagen deposition.<br />

Conclusions. Purinergic signaling regulated by ENTPD1 limits<br />

biliary fibrosis. In contrast, ENTPD2 seems to have less pronounced<br />

effects. These opposing roles are most likely due to<br />

differential catalytic function and localization of both enzymes<br />

with ENTPD1 as the dominant, vascular endothelial ecto-enzyme.<br />

Our findings further suggest a dominant role of ATP-mediated<br />

purinergic receptor signal transduction during biliary<br />

fibrogenesis.<br />

Disclosures:<br />

Moritz Schmelzle - Grant/Research Support: Novartis GmbH<br />

Simon C. Robson - Grant/Research Support: Pfizer, NIH, Dainippon; Independent<br />

Contractor: Biolegend, EMD Millipore, Mersana; Management Position:<br />

eBioscience; Speaking and Teaching: ACP, Elsevier, ATC; Stock Shareholder:<br />

Nanopharma, Puretech<br />

The following authors have nothing to disclose: Linda Feldbrügge, Shuji Mitsuhashi,<br />

Eva Csizmadia, Xiaofeng Sun


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 893A<br />

1395<br />

Inhibition of microRNA-214 ameliorates hepatic fibrosis<br />

and tumor incidence in platelet-derived growth factor<br />

C transgenic mice by rescuing Mig-6, EGFR/MET signal<br />

inhibitor.<br />

Hikari Okada 1 , Masao Honda 1 , Jean S. Campbell 2 , Kai Takegoshi<br />

1 , Yoshio Sakai 1 , Taro Yamashita 1 , Takayoshi Shirasaki 1 ,<br />

Riuta Takabatake 1 , Mikiko Nakamura 1 , Takuji Tanaka 3 , Shuichi<br />

Kaneko 1 ; 1 Gastroenterology, Kanazawa University Graduate<br />

School of Medical Science, Kanazawa, Japan; 2 Department of<br />

Pathology, University of Washington School of Medicine, Seattle,<br />

WA; 3 Cancer Research and Prevention, The Tohkai Cytopathology<br />

Institute, Gifu, Japan<br />

Objective Overexpression of platelet-derived growth factor-C<br />

(PDGF-C) in the liver of mice (Pdgf-c Tg) induces hepatic fibrosis,<br />

followed by the development of hepatocellular carcinoma.<br />

We identified differentially expressed miRNAs in Pdgf-c Tg<br />

mice and evaluated their functional relevance in the progression<br />

of hepatic fibrosis and the development of HCC. Materials<br />

and Method Pdgf-c Tg mice at 9 weeks and 32 months<br />

of age were injected with LNA-antimiR-214 via the tail vein<br />

twice at an interval of 7 days and six times (50 μg each) with<br />

3-week intervals by using Invivofectamine® 2.0. The degree<br />

of hepatic fibrosis, tumor incidence, tumor number, and liver<br />

weight were calculated. The expression of collagens I and IV,<br />

alpha-smooth muscle actin (α-SMA), Mig-6 (Errfi1), p-SMAD3,<br />

p-EGFR, p-MET, p-AKT and p-ERK was evaluated by immunohistochemical<br />

staining, real-time PCR, and Western blotting. To<br />

determine the target of miR-214, we used human HCC cell lines<br />

Huh-7 and HLF and immortalized human stellate cells Lx-2 in<br />

vitro. Results MiR-214 was the most significantly up-regulated<br />

miRNA in the liver of Pdgf-c Tg mice. Moreover, we showed the<br />

expression of miR-214 was up-regulated with the progression<br />

of hepatic fibrosis in a diet-induced (Ath+HF) NASH mouse<br />

model and in patients with chronic hepatitis B or C. Experiments<br />

of co-culture of hepatocytes (Huh-7 cells) and stellate cells<br />

(Lx-2 cells) revealed that miR-214 is transferred to hepatocytes<br />

by exosomes, which are released from activated stellate cells.<br />

Pdgf-c Tg mice treated with LNA-antimiR-214 showed a marked<br />

reduction in fibrosis and tumor incidence compared with salineor<br />

LNA-miR-control-injected control mice. We found mir-214<br />

targeted a negative regulator of EGFR signaling, Mig-6. Mimic-miR-214<br />

decreased the expression of Mig-6 and increased<br />

the levels of EGF-mediated p-EGFR (Y1173 and Y845) and<br />

p-Met (Tyr1234/1235), then increased growth of Huh-7 cells.<br />

Conversely, LNA-antimiR-214 repressed the expression of these<br />

genes and decreased cell growth. Similarly, Mig-6 could be a<br />

target of miR-214 in Lx-2 cells. The expression of Mig-6 was<br />

induced by LNA- antimiR-214 and over expression of Mig-6<br />

repressed the expression of COLA2, COL4A1, PDGFRβ, p-AKT<br />

and p-ERK in Lx-2 cells. Conclusion These results demonstrate<br />

that miR-214 participates in the development of hepatic fibrosis<br />

and tumor by targeting EGFR/MET signal inhibitor, Mig6.<br />

LNA-antimiR-214 is therefore potentially useful in the prevention<br />

of hepatic fibrosis and HCC.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Jean S. Campbell - Employment: OncoSec Medical<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Masao Honda, Kai Takegoshi,<br />

Yoshio Sakai, Taro Yamashita, Takayoshi Shirasaki, Riuta Takabatake, Mikiko<br />

Nakamura, Takuji Tanaka<br />

1396<br />

Essential Roles of RNA-binding Protein HuR in Activation<br />

of Hepatic Stellate Cells Induced by Transforming<br />

Growth Factor-β1<br />

Jingjing Ge, Na Chang, Zhongxin Zhao, Lei Tian, Lin Yang, Liying<br />

Li; Department of Cell Biology, Capital Medical University, Beijing,<br />

China<br />

Sphingosine kinase 1 (SphK1) is involved in transforming<br />

growth factor-β1 (TGF-β1)-induced activation of hepatic stellate<br />

cells (HSCs), which is characterized by increased expression<br />

of α-smooth muscle actin (α-SMA) and Collagen α1(I) in<br />

liver fibrogenesis. RNA-binding protein HuR up-regulates TGFβ1<br />

expression and mediates TGF-β1-induced profibrogenic<br />

actions. However, the molecular mechanism by which TGFβ1<br />

regulates SphK1 remains undefined. Here we investigated<br />

the role of HuR in SphK1 expression. Mice liver fibrosis were<br />

induced by bile duct ligation or carbon tetrachloride treatment.<br />

Expressions of HuR, SphK1, α-SMA and Collagen α1(I) in the<br />

fibrotic liver or human HSCs cell line (LX-2) were measured by<br />

real-time RT-PCR or Western blot. RNA immunoprecipitation<br />

was performed to detect the association between HuR and<br />

SphK1 mRNA. We found that HuR mRNA was increased in<br />

the damaged liver and had positive correlation with mRNA<br />

of TGF-β1, SphK1, α-SMA and Collagen α1(I), respectively.<br />

Importantly, up-regulation of SphK1 and activation of HSCs<br />

stimulated by TGF-β1 depended on HuR cytoplasmic accumulation,<br />

as HuR cytoplasmic translocation was blocked by Leptomycin<br />

B (LMB, a specific nuclear export inhibitor), the effects<br />

of TGF-β1 were diminished. In addition, silencing HuR significantly<br />

abolished the up-regulation of SphK1 and activation of<br />

HSCs evoked by TGF-β1, and overexpressing HuR mimicked<br />

the effects of TGF-β1. Furthermore, TGF-β1 prolonged half-life<br />

of SphK1 mRNA by promoting its binding with HuR. Pharmacological<br />

or siRNA-induced inhibition of SphK1 abrogated<br />

HuR-mediated activation of HSCs. In conclusion, these data<br />

suggest that HuR binds with SphK1 mRNA and plays a crucial<br />

role in TGF-β1-induced activation of HSCs.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jingjing Ge, Na Chang, Zhongxin<br />

Zhao, Lei Tian, Lin Yang, Liying Li<br />

1397<br />

HIV infected Kupffer cells mediate pro-fibrogenic effects<br />

on stellate cells through secretion of TGFβ1<br />

Ankur Panchal, Arevik Mosoian, Lumin Zhang, M. Isabel Fiel,<br />

Sander S. Florman, Myron E. Schwartz, Andrea D. Branch, Meena<br />

B. Bansal; Icahn School of Medicine at Mount Sinai, New York,<br />

NY<br />

Background: Liver disease is the most common cause of<br />

non-AIDS related mortality in HIV-infected patients receiving<br />

effective anti-retroviral therapies. TGFβ1 is an important profibrogenic<br />

cytokine in chronic liver disease and both serum<br />

and intrahepatic levels of TGFβ1 are increased in both HIV<br />

monoinfected and HIV/HCV coinfected patients, respectively.<br />

As the resident liver macrophage, the Kupffer cell, has been<br />

shown to be susceptible to HIV infection in vitro and in vivo<br />

and is the predominant cellular source of intrahepatic TGFβ1,<br />

we hypothesized that HIV infection of Kupffer cells (KCs) results<br />

in increased TGFβ1 secretion with subsequent paracrine pro-fibrogenic<br />

effects on hepatic stellate cells (HSCs), the primary<br />

collagen producing cell in liver fibrosis. Methods: Primary KCs<br />

and HSCs were isolated by density centrifugation from normal<br />

livers in patients undergoing hepatic resection for primary<br />

benign tumors or a single metastasis from colon cancer. KC


894A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

preps were enriched by rapid adherence. Purity of KCs was<br />

assessed by RT-PCR and FACS for CD68, CD14, TLR4, CD3,<br />

and CD31. Primary KCs were infected with HIV-BaL (MOI=0.1)<br />

or mock infected and supernatants were collected at serial<br />

time points for p24, a marker of HIV replication, and TGFβ1<br />

ELISA. Pro-fibrogenic effects of conditioned media from KCs on<br />

HSCs was assessed by qRT-PCR and specific effects of TGFβ1<br />

were assessed using TGFβ1 blocking/neutralizing antibody.<br />

Recombinant TGFβ1 (10ng/mL) was used as a positive control.<br />

Results: 1) HIV Infection of KCs resulted in an 1.5 fold increase<br />

in TGFβ1 secretion (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 895A<br />

sustained to be increased in TAA+Vector group at week 3<br />

whereas TAA+Relaxin rats had suppressed TIMP-2 expression<br />

by then. Liver α-smooth muscle actin (α-SMA) mRNA in<br />

both TAA+Vector and TAA+Relaxin rats was increased after<br />

3 days but reversed to the normal level at week 3. Analysis<br />

of TGFβ and MMP9 mRNA demonstrated the similar pattern.<br />

Conclusions: A single adenoviral delivery of relaxin in the liver<br />

attenuated established hepatic fibrosis by suppressing collagen<br />

cross-links and enhancing collagen degradation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jung Il Lee, Ja Kyung Kim, Hye<br />

Young Chang, Kwan Sik Lee<br />

1400<br />

Galectin-1 promotes the activation of hepatic stellate<br />

cell and liver fibrosis through glycosylation-dependent<br />

interactions<br />

Ming-Heng Wu; Graduate Institute of Translational Medicine, College<br />

of Medical Sciences and Technology, Taipei Medical University,<br />

Taipei, Taiwan<br />

The activation of hepatic stellate cell (HSC) is a rate-limiting<br />

step in hepatic fibrosis. Therefore, developing approaches<br />

to destroy or normalize activated HSCs is important for liver<br />

fibrosis therapy. The coevolution of cell-surface glycans and<br />

glycan-binding proteins regulate the progression of physiologic<br />

and pathologic processes but this issue has not been<br />

well addressed in HSC activation and liver fibrosis. Here, we<br />

reported that galectin-1 (Gal-1), a β-galactoside binding protein,<br />

was highly expressed in the fibrotic livers. Gal-1 binding<br />

glycans [β1,6 N-acetylglucosamine-branched N-glycans and<br />

poly-N-acetylactosamine (LacNAc)] were highly presented in<br />

activated HSCs and facilitated Gal-1 binding. In contrast, the<br />

amounts of α2,6 sialylation glycan (which masks the Gal1<br />

binding glycan) is lower in activated HSCs than in quiescent<br />

HSCs. Two glycosyltransferases, Mgat5 (N-acetylglucosaminyltransferase<br />

5) and GCNT1 (core 2 N-acetylglucosaminyltransferase<br />

1), mediated N- and O-glycans were required for<br />

Gal-1 stimulated migration and activation of HSCs. In addition,<br />

Gal-1 regulated PDGF and TGF-β1 induced signaling and<br />

HSC activation through interacting with neuropilin-1 (NRP-1).<br />

Using a mouse model, we found that thioacetamide (TAA), an<br />

organosulfur compounds, induced liver fibrosis was attenuated<br />

in Gal-1 null mice. In summary, our results demonstrated that<br />

the crucial role of Gal-1 in HSC activation and liver fibrosis.<br />

Strategies that disrupt the Gal-1/NRP-1 interactions could be<br />

developed for liver fibrosis therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ming-Heng Wu<br />

1401<br />

TRPC6 Channels Contribute to Hepatic Stellate Cell Activation<br />

Leading to Liver Fibrosis<br />

Kyu-Hee Hwang, Ji-Hee Kim, Soo-Jin Kim, Kyu-Sang Park, Seung-<br />

Kuy Cha; Physiology, Yonsei University Wonju College of Medicine,<br />

Wonju, Korea (the Republic of)<br />

Activation of hepatic stellate cells (HSCs) is a main cause of<br />

liver cirrhosis and portal hypertension. In response to injury,<br />

HSCs are activated by diverse hormones and growth factors<br />

such as angiotensin II and endothelin-1 whose receptors are<br />

linked to phospholipase activation leading to Ca 2+ influx<br />

termed receptor-operated Ca 2+ entry (ROCE). The Ca 2+ -mediated<br />

signaling pathways are implicated either directly or<br />

indirectly in activation of HSCs causing de novo expression of<br />

a-smooth muscle actin (αSMA) and/or profibrotic ligand TGFβ.<br />

However, the molecular identity and underlying mechanism of<br />

ROCE involving HSC activation are ill-defined. Here, we report<br />

that TRPC6 channel is a predominant molecular component<br />

of ROCE mediating HSC activation. Among TRPC sub-family,<br />

the TRPC6 expression was significantly increased in the bile<br />

duct ligation- and TAA-induced liver fibrosis animal models.<br />

Functionally, TRPC6-stimulated Ca 2+ influx was confirmed by<br />

specific blockade of TRPC6 with the siRNA against Trpc6 and<br />

pharmacological inhibitor SKF96365 in cultured human HSCs.<br />

Furthermore, overexpression of TRPC6 by gene delivery in<br />

mouse liver induced de novo expression of αSMA and TGFβ<br />

suggesting that TRPC6-mediated Ca 2+ influx may involve in<br />

HSC activation and liver fibrosis. These results provide a new<br />

perspective on the pathogenesis of liver fibrosis and may provide<br />

clues to treatment of the liver cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kyu-Hee Hwang, Ji-Hee Kim, Soo-<br />

Jin Kim, Kyu-Sang Park, Seung-Kuy Cha<br />

1402<br />

Targeting Activated Stromal Fibroblasts in Primary Sclerosing<br />

Cholangitis<br />

Anja Moncsek 1 , Achim Weber 2 , Beat Müllhaupt 1 , Joachim C.<br />

Mertens 1 ; 1 Gastroenterology and Hepatology, University Hospital<br />

Zürich, Zürich, Switzerland; 2 Surgical Pathology, University Hospital<br />

Zürich, Zürich, Switzerland<br />

Background: Primary Sclerosing Cholangitis (PSC) is a chronic<br />

cholestatic condition with diffuse biliary inflammation and fibrosis<br />

that can affect the whole biliary system. The disease has<br />

a progressive course and results in biliary cirrhosis with all<br />

its sequelae. PSC is also one of the main known risk factors<br />

for the development of cholangiocarcinoma (CCA). We have<br />

previously shown that activated stromal fibroblasts (ASF) in<br />

cholangiocarcinoma, which have been implicated in tumor<br />

progression and metastasis display an activated phenotype<br />

with increased sensitivity to apoptotic stimuli and can be selectively<br />

targeted with pro-apoptotic BH3 mimetics (Mertens et.al.<br />

Cancer Res 2013). However, the role of ASF in PSC and progression<br />

to CCA is not well understood. Thus, our aim was<br />

to characterize ASF in PSC and explore selective pro-apoptotic<br />

targeting of this cell population. Methods: Human PSC<br />

specimens were examined by histology, immunofluorescence,<br />

Western blot and qRT-PCR. The MDR2 -/- mouse model was<br />

employed for in vivo <strong>studies</strong>. Results: The PSC tissue specimens<br />

we examined contain an abundance of activated stromal fibroblasts.<br />

By immunofluorescence the stromal compartment stains<br />

positive for Tenascin C and Collagen I, while the stromal fibroblasts<br />

are positive for alpha smooth muscle actin (aSMA). This<br />

phenotype resembles ASF in CCA. We were able to confirm<br />

the phenotype of PSC-ASF by PCR. The MDR2 -/- mouse model<br />

of PSC develops a periductal and bridging fibrosis within 8 to<br />

12 weeks. In analogy to previous experiments in a rat model<br />

of CCA, MDR2 -/- mice were treated with the pro-apoptotic BH3<br />

mimetic navitoclax for 2 weeks. Assessment of liver fibrosis by<br />

Hydroxyproline assay and Sirius red staining demonstrated a<br />

reduction in fibrosis. Preliminary results revealed a reduction of<br />

activated stromal fibroblasts. Conclusions: ASF in PSC resemble<br />

stromal fibroblasts in cholangiocarcinoma and display apoptotic<br />

sensitization that could be targeted therapeutically.<br />

Disclosures:<br />

The following authors have nothing to disclose: Anja Moncsek, Achim Weber,<br />

Beat Müllhaupt, Joachim C. Mertens


896A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1403<br />

Modulation of Extracellular Matrix Remodeling by Activation<br />

of the Aryl Hydrocarbon Receptor<br />

Cheri L. Lamb 1,2 , Kristen A. Mitchell 1,2 ; 1 Biological Sciences, Boise<br />

State University, Boise, ID; 2 Biomolecular Sciences Ph.D. Program,<br />

Boise State University, Boise, ID<br />

The aryl hydrocarbon receptor (AhR) is a soluble, ligand-activated<br />

transcription factor involved in developmental processes,<br />

xenobiotic metabolism, and adaptation to environmental stress.<br />

Recent evidence indicates that this receptor may also contribute<br />

to wound healing processes. We recently found that exposure<br />

to the potent AhR agonist 2,3,7,8 tetrachlorodibenzo-p-dioxin<br />

(TCDD) increases activation of hepatic stellate cells, which<br />

contribute to hepatic wound healing and fibrosis through the<br />

deposition of extracellular matrix material, namely collagen<br />

type I. The goal of the present study was to determine how<br />

TCDD treatment impacts matrix deposition and remodeling<br />

activities during experimental liver fibrosis elicited by carbon<br />

tetrachloride (CCl 4<br />

). Male C57Bl/6 mice were treated with<br />

CCl 4<br />

(5 ml/kg diluted 1:10 in corn oil) twice a week for eight<br />

weeks. During the final two weeks, mice were treated with<br />

TCDD (20 μg/kg by gavage) or vehicle (peanut oil). Collagen<br />

deposition was measured by cytohistological staining and<br />

evaluated by polarized microscopy. Levels of matrix enzymes<br />

and associated regulatory proteins were measured by qPCR,<br />

and activity was assessed by in situ zymography. Results indicate<br />

that TCDD treatment increased Col1a1 mRNA levels and<br />

altered the pattern of collagen deposition in the fibrotic septa.<br />

Exposure to TCDD also increased in situ collagenase activity<br />

in the fibrotic liver with a concomitant 4-fold increase in matrix<br />

metalloproteinase-13 mRNA levels. Inhibitors of matrix metalloproteinases,<br />

namely PAI-1 and TIMP-1, were increased in<br />

TCDD-treated mice, which could reflect direct AhR transcriptional<br />

activity or else a compensatory response to elevated<br />

collagenase activity. Collectively, these results indicate that<br />

TCDD-induced AhR activation not only increases fibrogenesis,<br />

but also perturbs extracellular matrix remodeling activities in<br />

the fibrotic liver. Hence, these findings implicate a previously<br />

unidentified role for the AhR in modulating wound healing<br />

responses in the liver and may provide a basis for designing<br />

therapeutic strategies to limit fibrosis by modulating AhR activation.<br />

Disclosures:<br />

The following authors have nothing to disclose: Cheri L. Lamb, Kristen A. Mitchell<br />

1404<br />

MicroRNA profiling of circulating exosomes in HBV cirrhosis<br />

patients following anti-HBV therapy<br />

Min Cong 1 , Li Chen 2 , Tianhui Liu 1 , Ping Wang 1 , Jidong Jia 1 , David<br />

Brigstock 2 , Hong You 1 ; 1 Liver Research Center, Beijing Friendship<br />

Hospital, Beijing, China; 2 The research Institute at Nationwide<br />

Children’s Hospital, Columbus, OH<br />

Aim: The objective of our study is to establish serum exosomal<br />

microRNA (miR) signatures to predict progression and reversion<br />

of fibrosis in hepatitis B (HBV)–induced cirrhosis undergoing<br />

anti-viral therapy. Methods: Twelve HBV patients with<br />

compensated cirrhosis were selected based on their cirrhosis<br />

regression (decreased Fibroscan values of > 5kPa) or progression<br />

(increased or stable Fibroscan values) after treatment for 6<br />

months with Entecavir (0.5mg/day oral), respectively. Twelve<br />

matched healthy blood donors were used as a reference control<br />

group. Serum (0.5ml) from the HBV patients, before or<br />

after therapy, or from the controls was individually processed<br />

to purify exosomes from which small RNA was then isolated.<br />

Samples in each group were pooled and analyzed using a<br />

Human miRNome PCR array to profile the1066 most abundantly<br />

expressed human miR sequences using SYBR Green real<br />

time PCR. Circulating exosomalmiRs were isolated from 8 other<br />

compensated cirrhosis HBV patients who presented with cirrhosis<br />

regression or progression to validate the change of selected<br />

exosomalmiRs by qRT-PCR. Results: Compared with control<br />

subjects, 59 out of 1066 miRs from 12 regressed cirrhosis<br />

HBV patients were up- or down-regulated at least 15-fold in the<br />

cirrhotic patients, with 46 out of 59 miRs showing correction<br />

(to more healthy) after therapy. Among the miRs identified in<br />

this unbiased analysis were miR-7, -15b, -16, -22, -26a, -27b,<br />

-29, -92b, -101, -183, -191, -196a, -200a, -370, -378a and<br />

-936 and let-7a-g which have been reported in the literature to<br />

be fibrosis-related in the liver or other organ systems. Analysis<br />

of exosomalmiRs from 12 progressed cirrhosis HBV patients<br />

showed that these miRs were less corrected as compared to<br />

the regressed patients and that different miRs served to discriminate<br />

these patients from the healthy controls. qRT-PCR analysis<br />

of exosomalmiRs from 8 cirrhosis HBV patients, before or after<br />

therapy, showed that miR-16,-20a,-25,-148a,-223,-320d were<br />

significantly decreased while miR-126,-191 were significantly<br />

increased in progressed patients. Conclusions: These data<br />

reveal that the circulating exosomalmiR content varies according<br />

to the stage of HBV fibrosis and that exosomalmiRs show<br />

quantitative and qualitative changes that reflect the patients’<br />

clinical course of fibrosis regression or progression after anti-viral<br />

treatment. Since circulating exosomalmiRs change dynamically<br />

with disease status, these findings highlight the possibility<br />

of identifying discriminatory miR signatures that are diagnostic<br />

or predictive for fibrosis and that can be obtained longitudinally<br />

and non-invasively in individual patients.<br />

Disclosures:<br />

Jidong Jia - Consulting: BMS, MSD, Roche<br />

David Brigstock - Stock Shareholder: FibroGen<br />

The following authors have nothing to disclose: Min Cong, Li Chen, Tianhui Liu,<br />

Ping Wang, Hong You<br />

1405<br />

Identification of Axon guidance signaling pathway<br />

mediators in hepatic stellate cells and liver fibrogenesis<br />

Jinsheng Guo 1 , David Zhang 2 , Yujing Wu 1 , Daisuke Hasegawa 2 ,<br />

Scott L. Friedman 1,2 ; 1 Division of Digestive Diseases, Zhong Shan<br />

Hospital, Fu Dan University, Shanghai, China; 2 Division of Liver<br />

Diseases, Icahn School of Medicine at Mount Sinai, New York, NY<br />

Background&Aims: Hepatic stellate cells (HSC) have many<br />

features of neural cells including expression of glial fibrillary<br />

protein, nestin and neurotrophin receptors. We recently used<br />

an informatics based approach to identify novel transcripts<br />

expressed by stellate cells encoding cell surface and secreted<br />

proteins (Zhang DY et al, Gut, 2015, in press). Among these,<br />

Robo2 is a cell surface molecule implicated in axon guidance.<br />

The aim of our study was to identify the novel signaling pathway<br />

and key molecules in liver fibrogenesis by dynamic network<br />

analysis of liver transcriptomes. Methods&Results: Two<br />

experimental models of hepatic fibrosis were employed by<br />

intraperitoneally injection of diethylnitrosamine (DEN, 10mg/<br />

kg/week) in mice for 14 weeks, or by intraperitoneal injections<br />

of carbon tetrachloride (CCl4, 0.9ml/kg/biw) for four weeks.<br />

In the DMN model, the liver samples were collected at 0, 6, 8<br />

and 14 weeks post DEN treatment for RNA sequencing. The<br />

transcriptomic data associated with development of liver fibrosis<br />

(i.e., injury, early fibrotic, late fibrotic or cirrhotic stages)<br />

were applied to time-series analysis, followed by network<br />

and pathway analyses. Among markedly up-regulated genes


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 897A<br />

during fibrosis and cirrhosis, 48 were components of axon<br />

guidance signaling pathways (P=4.65E-07, Enrichment=2.53),<br />

which transcripts encoding ligands (slit2, Sema4D), membrane<br />

receptors (Robo1 and 2, EphA and B, CXCR4, Plexin A4), and<br />

downstream kinases (PAK1, 2,3 and 6, Gnai 1 and 2, Rock2,<br />

and Mapk3). By immunoflurescence in immortalized human<br />

and mouse hepatic stellate cell (HSC) lines LX-2 and JS1 in<br />

vitro, and in the CCl4 induced fibrotic liver samples in vivo,<br />

Robo2, a key receptor mediating axon guidance signaling,<br />

was localized to the HSC surface. The amount and localization<br />

of its protein expression was correlated well with fibrotic septa<br />

in fibrotic livers. Conclusion: There is significant upregulation<br />

and activation of axon guidance signaling pathway members<br />

in liver during fibrogenesis. This signaling may have regulate<br />

actin cytoskeleton, cell attraction and outgrowth of HSC. The<br />

study also revealed RoBo2 as a novel HSC surface marker and<br />

a potential therapeutic and diagnostic target of liver fibrosis.<br />

Disclosures:<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

The following authors have nothing to disclose: Jinsheng Guo, David Zhang,<br />

Yujing Wu, Daisuke Hasegawa<br />

1406<br />

PDGFRα ubiquitination leads to its p62 mediated autophagic<br />

degradation in a synectin dependent manner in<br />

HSC<br />

Haibin Yu, Mary Drinane, Usman Yaqoob, Vikas K. Verma, Thiago<br />

de Assuncao, Sheng Cao, Vijay Shah; Gastroenterology Research<br />

Unit, Mayo Clinic, Rochester, MN<br />

Background/Aim: The platelet derived growth factor receptor<br />

(PDGFR) signaling pathway is critical for hepatic stellate cell<br />

(HSC) activation and liver fibrosis. While the critical role of<br />

PDGFRβ is well studied, recent work suggests importance of<br />

PDGFRα as well, making its regulation of interest. Synectin is a<br />

scaffold protein that is a key regulator of endocytic trafficking<br />

of multiple tyrosine kinases. Prior <strong>studies</strong> showed that knockdown<br />

of synectin from HSC led to autophagic degradation<br />

of PDGFRα. p62 is a ubiquitin-binding autophagy receptor<br />

which selectively recognizes autophagic cargo and mediates<br />

engulfment into autophagosomes in coordination with LC3B.<br />

We hypothesized that synectin regulates selective autophagic<br />

degradation of PDGFRα by recognizing specific ubiquitin sites<br />

which in turn govern p62 binding with PDGFRα. Methods/<br />

Results: Co-immunoprecipitation (co-IP) of PDGFRα and PDG-<br />

FRβ was used to investigate p62 binding using HSC lysates.<br />

p62 co-precipitated with PDGFRα, but not PDGFRβ. In response<br />

to PDGF ligand, co-precipitation of PDGFRα with p62 was further<br />

increased by 4.4 fold (p < 0.05). Synectin knockdown led<br />

to a 2-fold increase in binding of p62 with PDGFRα (p < 0.05)<br />

and increased co-localization of p62 and PDGFRα by confocal<br />

microscopy. Sequence analysis of PDGFRα and PDGFRβ was<br />

performed which identified two ubiquitin sites present within<br />

α but not β subunits. Site directed mutagenesis of PDGFRβ<br />

was performed to mutate the two corresponding residues to<br />

lysine (Q613K and R979K) and generate a mutant PDGFRβ<br />

construct (PDGFRβ- Q613K, R979K) that could be used to test<br />

the hypothesis that these two lysine residues may be critical for<br />

ubiquitin dependent PDGFR autophagic degradation. PDGFRβ-<br />

Q613K, R979K underwent autophagic degradation, akin to<br />

PDGFRα, while the wild-type PDGFRβ did not (p < 0.05). Conclusion:<br />

The specificity of interaction of the autophagic receptor<br />

p62 with PDGFR is conferred by specific ubiquitin sites and is<br />

mediated by synectin.<br />

Disclosures:<br />

The following authors have nothing to disclose: Haibin Yu, Mary Drinane, Usman<br />

Yaqoob, Vikas K. Verma, Thiago de Assuncao, Sheng Cao, Vijay Shah<br />

1407<br />

Reactive gamma-ketoaldehydes induce a proinflammatory<br />

phenotype, oxidative stress, and activation of autophagy<br />

in primary human hepatic stellate cells<br />

Lisa Longato 1 , Fausto Andreola 1 , Sean S. Davies 2 , Jackson L. Roberts<br />

2 , Giuseppe Fusai 1 , Massimo Pinzani 1 , Kevin Moore 1 , Krista<br />

Rombouts 1 ; 1 Medicine, University College London, London, United<br />

Kingdom; 2 Vanderbilt University, Nashville, TN<br />

Background and aims. Products of oxidative stress such as<br />

4-hydroxynonenal are key activators of hepatic stellate cells<br />

(HSCs). γ-Ketoaldehydes (γ-KAs), including levuglandins and<br />

their isomers, are a family of acyclic aldehydes, which are<br />

formed during oxidation of arachidonic acid, or as a by-product<br />

of the cyclooxygenase pathway. γ-KAs are highly reactive<br />

and form protein adducts and cross-links at a rate > 100-fold<br />

compared to 4-HNE. Increased serum antibodies against proteins<br />

cross-linked to γ-KAs are present in patients with alcoholic<br />

liver disease. Since the contribution of γ-KAs to liver injury has<br />

not been studied, we sought to investigate the effects of γ-KA<br />

levuglandin E 2<br />

(LGE 2<br />

) on the cell biology of primary human<br />

HSCs. Methods. Primary human HSCs were exposed to LGE 2<br />

(0.5 pM- 5 mM) for up to 48 hours and analyzed for proliferation,<br />

cytotoxicity (lactate dehydrogenase, MTS, and Neutral<br />

Red assays), RNA/protein expression, reactive oxygen<br />

species (ROS) production, and ultrastructure. Results. Exposure<br />

to a 5 μM dose of LGE 2<br />

was profoundly cytotoxic and<br />

resulted in apoptotic cell death, as indicated by LDH leakage,<br />

reduced MTS and Neutral Red incorporation, increased levels<br />

of cleaved PARP, CHOP, and JNK phosphorylation. Lower,<br />

non-cytotoxic doses (50 nM-500 nM) of LGE 2<br />

induced selected<br />

indices of HSCs activation, such as increased expression of<br />

a-smooth muscle actin (α-SMA), and activation of signaling<br />

pathways (ERK1/2). In contrast, LGE 2<br />

had no effect on DNA<br />

synthesis or pro-fibrogenic markers, particularly collagen type<br />

I, and lysyl oxidase (LOX). In contrast, HSCs exposed to LGE 2<br />

displayed a marked increase in expression for various cytokines/chemokines<br />

including interleukin-8, -6, -1β, and MCP-1.<br />

This was accompanied by increased secretion of bioactive<br />

IL1β/IL-18, as measured by HEK-Blue reporter cells. Pre-incubation<br />

of cells with the either chemicals inhibitors of JNK<br />

(SP600125), or NFkB (PDTC), but not p38MAPK (SB203580),<br />

partially reduced these increases in cytokines and chemokines.<br />

Ultrastructural analyses revealed that LGE 2<br />

-treated HSCs had<br />

prominent formation of autophagic vescicles, suggesting ongoing<br />

autophagy, which was confirmed by immunofluorescence<br />

for LC3B, as well as by accumulation of LC3II. Lastly, shortterm<br />

exposure to LGE 2<br />

promoted a dose-dependent formation<br />

of ROS in the cells, as measured by the carboxy-H 2<br />

DCFDA<br />

fluorescent probe. Conclusions. γ-KAs represent a newly identified<br />

class of activators of HSCs in vitro, which are biologically<br />

active at concentrations as low as 50 nM, and are particularly<br />

effective at promoting a pro-inflammatory response, as well as<br />

oxidative stress, and autophagy.<br />

Disclosures:<br />

Massimo Pinzani - Advisory Committees or Review Panels: Intercept Pharmaceutical,<br />

Silence Therapeutic, Abbot; Consulting: UCB; Speaking and Teaching:<br />

Gilead, BMS


898A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Kevin Moore - Advisory Committees or Review Panels: Servier<br />

The following authors have nothing to disclose: Lisa Longato, Fausto Andreola,<br />

Sean S. Davies, Jackson L. Roberts, Giuseppe Fusai, Krista Rombouts<br />

1408<br />

Sequence Profiling of miRNAs Reveals mir-150-5p as a<br />

Regulator of Transdifferentiation<br />

Luigi Locatelli, Pier P. Paoli, Timothy Hardy, Rachel M. Howarth,<br />

Fiona Oakley, Derek Mann; Institute of Cellular Medicine, Newcastle<br />

University, Newcastle upon Tyne, United Kingdom<br />

Cellular transdifferentiation is important in many physiological<br />

processes (e.g. wound healing), and is implicated in pathologies<br />

such as cancer and tissue fibrosis. To-date the role of<br />

miRNAs, a class of small noncoding RNAs that regulate gene<br />

expression by binding to target mRNAs suppressing its translation<br />

or initiating its degradation, in transdifferentiation has not<br />

been fully addressed. We use a cell culture model for transdifferentiation<br />

of human and rat hepatic stellate cells (HSC) into<br />

myofibroblasts to determine functional changes in miRNAs. The<br />

aims of this study were: a) perform Next Generation Sequencing<br />

(NGS) to generate a miRNA profile in human and rat<br />

activated HSCs to identify potential miRNA-mRNA regulatory<br />

networks in common with both; b) replicate NGS results based<br />

on gene profile and biological validation of specific miRNA<br />

target genes. The most critical step in miRNA biology lies in the<br />

definition of the rules for miRNA target recognition. Data generated<br />

by sequencing were analysed via the Seq-Imp pipeline<br />

(http://www.ebi.ac.uk/research/enright/software/kraken), to<br />

identify individual miRNA and an initial statistical analysis.<br />

Further analysis was performed with a custom R script to generate<br />

data correlation and heatmap matrices. Finally, targets for<br />

each miRNA were identified from TargetScan, microCOSM,<br />

and miRTar, by picking genes that were in at least two out of<br />

three databases, revealing 11 significantly upregulated and 11<br />

downregulated miRNAs that were identical in both activated<br />

human and rat HSC. From the list of differentially regulated<br />

miRNAs, we focused our attention on mir-150-5p, which regulates<br />

a series of target proteins such as the Zinc finger E-box<br />

binding homeobox (ZEB1) involved in the cell cycle, cellular<br />

development, and apoptosis. Particularly mir-150-5p has been<br />

found to be downregulated in both human and rat activated<br />

HSCs. A significant inverse correlation between ZEB1 and mir-<br />

150-5p was found at the RNA and protein level. In vitro we<br />

showed that the overexpression of mir-150-5p in human and<br />

rat HSCs by a mir-150-5p mimic was able to downregulate<br />

the expression of ZEB1 with an observed reduction of extracellular<br />

matrix proteins collagen type I (COL1A1) and α-smooth<br />

muscle actin (α-SMA). Moreover, the level of mir-150-5p in the<br />

serum of 10 Alcoholic Liver Disease (ALD) patients has been<br />

found to be significantly downregulated compared to healthy<br />

controls. In conclusion, HSC transdifferentiation is associated<br />

with discrete alterations in the expression of miRNAs, and in<br />

particular we have identified mir-150-5p as a regulator of the<br />

myofibroblast phenotype and potential biomarker of fibrosis in<br />

ALD patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Luigi Locatelli, Pier P. Paoli, Timothy<br />

Hardy, Rachel M. Howarth, Fiona Oakley, Derek Mann<br />

1409<br />

Human precision Cut Liver Slices model for testing<br />

anti-fibrotic drugs<br />

Lynda Aoudjehane 1,2 , Margaux Legrand 1 , Grégoire Bisch 1 , Rolland<br />

Delelo 2 , Chantal Housset 2 , Pierre Balladur 3 , Claire Goumard<br />

4 , Jérôme Becquart 1 , Yvon Calmus 2,5 , Filomena Conti 2,5 ;<br />

1 Human HepCell, Faculté de Médecine Pierre et Marie Curie,<br />

Paris, France; 2 UMPC UNIV Paris 6 & INSERM, UMR_S 938,<br />

CDR Saint-Antoine, Paris, France; 3 AP-HP, Hôpital Saint Antoine,<br />

Service de Chirurgie Digestive, Paris, France; 4 AP-HP, Hôpital Pitié-<br />

Salpêtrière, Service de Chirurgie Digestive, Paris, France; 5 AP-<br />

HP, Hôpital Pitié-Salpêtrière, Unité de Transplantation Hépatique,<br />

Paris, France<br />

Background: Fibrosis is the common end stage of all chronic<br />

liver diseases. Understanding the molecular mechanisms underlying<br />

liver fibrogenesis is mandatory to develop new antifibrotic<br />

therapies. Current methods for evaluating the efficacy<br />

of anti-fibrotic drugs rely mostly on immortalized cell lines and<br />

animal models. Human precision-cut liver slices (PCLS) preserve<br />

the cell types and proportions, and particularly the cell-cell<br />

interactions. Several <strong>studies</strong> have recently suggested that PCLS<br />

are potentially useful as in vitro model to study liver fibrosis.<br />

The current study aimed to test the efficacy of human PCLS to<br />

evaluate the effect of anti-fibrotic drugs. Methods: Human PCLS<br />

of 250 μm thickness were obtained from normal or cirrhotic<br />

liver samples. PCLS were incubated for 1-48h at 37°C under<br />

85% O2 and, after 24h, were treated with different anti-fibrotic<br />

drugs: ciclosporin A (CsA), pirfenidone (PFD) or SMAD3<br />

inhibitor (SIS3). The viability of PCLS was assessed by ATP production.<br />

Hepatic metabolism and fibrosis markers (α-SMA and<br />

Coll1, HSP47, TIMP-1, MMP-2 and TGF-beta) were determined<br />

by RT-PCR. Results: PCLS from normal and cirrhotic human liver<br />

samples could be cultured with an excellent viability over 48h<br />

and this culture induced a spontaneous fibrosis. The expression<br />

of fibrotic markers (α-SMA and Coll1, HSP47, TIMP-1,<br />

MMP-2 and TGF-beta) significantly increased at 24 and 48h<br />

when compared to 1h. Fibrosis markers significant decreased<br />

(α-SMA and Coll1 and TIMP1 mRNA expression) when normal<br />

or cirrhotic human PCLS were treated for 24h with PFD<br />

and SIS3. However, CsA had no significant effect on fibrosis<br />

marker expression on normal or cirrhotic PCLS. Conclusion:<br />

These preliminary results suggest that human PCLS can be<br />

maintained in culture for 48 hours without significant modification<br />

of viability. The culture of PCLS induces a spontaneous<br />

activation of the fibrotic process. This feature can be used to<br />

study the effect of antifibrotic molecules and thus constitutes<br />

an interesting physiological human model for the screening of<br />

antifibrotic drugs.<br />

Disclosures:<br />

Jérôme Becquart - Management Position: Human Hepcell<br />

The following authors have nothing to disclose: Lynda Aoudjehane, Margaux Legrand,<br />

Grégoire Bisch, Rolland Delelo, Chantal Housset, Pierre Balladur, Claire<br />

Goumard, Yvon Calmus, Filomena Conti<br />

1410<br />

The AMPK-related kinase NUAK2 interacts with TGF-β<br />

and regulates the activation of hepatic stellate cells<br />

Cristina Tosti Guerra 1 , Alessandra Caligiuri 1 , Angela Provenzano 1 ,<br />

Krista Rombouts 2 , Massimo Pinzani 2 , Fabio Marra 1 ; 1 University<br />

of Florence, Florence, Italy; 2 University College London, London,<br />

United Kingdom<br />

Background and Aims: Nuak2 is a member of AMPK related<br />

kinases (ARKs), that act as energy sensors and controllers of<br />

cellular structure, with different effects on cell motility and cytoskeletal<br />

organization, depending on the cell types. AMPK pos-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 899A<br />

sesses anti-fibrogenic activity. A recent study showed a link<br />

between Nuak2 and TGF-β, a major driver of hepatic fibrogenesis.<br />

However, little information is available on the role<br />

of Nuak2 in liver fibrosis, and in particular on HSC trans-differentiation.<br />

Aims of this study was to elucidate the involvement<br />

of Nuak2 in HSC transactivation, and to investigate the<br />

possible interaction between Nuak2 and TGF-β signaling.<br />

Methods: HSC were isolated from normal rat and human livers<br />

and activated by culture on plastic. Knockdown of Nuak2 was<br />

achieved by siRNA. Cell migration was evaluated in modified<br />

Boyden Chambers. Protein expression and signaling pathways<br />

were analyzed by Western blotting. Results: In fully activated<br />

HSC, down-regulation of Nuak2 positively modulated cell<br />

migration and induced changes in the expression of molecules<br />

involved in cytoskeletal organization. Knockdown of Nuak2<br />

increased α-SMA expression and phosphorylation of SMAD3,<br />

in HSC exposed to TGFβ. Moreover, Nuak2 expression was<br />

up-regulated in HSC following TGFβ treatment. Conclusions:<br />

The AMPK related kinase, Nuak2, is modulated during the activation<br />

process of HSC, regulates cytoskeletal organization and<br />

cell motility and interacts with TGF-β in regulating the pro-fibrogenic<br />

properties of HSC.<br />

Disclosures:<br />

Massimo Pinzani - Advisory Committees or Review Panels: Intercept Pharmaceutical,<br />

Silence Therapeutic, Abbot; Consulting: UCB; Speaking and Teaching:<br />

Gilead, BMS<br />

Fabio Marra - Advisory Committees or Review Panels: Abbvie; Consulting: Bayer<br />

Healthcare; Grant/Research Support: ViiV<br />

The following authors have nothing to disclose: Cristina Tosti Guerra, Alessandra<br />

Caligiuri, Angela Provenzano, Krista Rombouts<br />

1411<br />

The bile acid-phospholipid conjugate Ursodeoxycholyl<br />

Lysophosphatidylethanolamide (UDCA-LPE) disturbs<br />

pro-fibrogenic Integrin and TGFβ signaling<br />

Jie Su, Hongying Gan-Schreier, Walee Chamulitrat, Wolfgang<br />

Stremmel, Anita Pathil; Department of Internal Medicine IV, University<br />

of Heidelberg, Heidelberg, Germany<br />

BACKGROUND: Integrin receptors, which are involved in cellcell<br />

and cell-matrix interaction emerge as crucial mediators of<br />

TGFβ1 activation in liver fibrosis. Ursodeoxycholyl Lysophosphatidylethanolamide<br />

(UDCA-LPE) is a synthetic bile acid-phospholipid<br />

conjugate with hepatoprotective and anti-fibrogenic<br />

functions in vitro and in vivo. In this study we aim to elucidate<br />

signaling pathways, which mediate anti-fibrogenic action of<br />

UDCA-LPE. RESULTS: In order to promote pro-fibrogenic signaling<br />

upon extracellular matrix binding integrins recruit focal<br />

adhesion kinase (FAK) and SRC kinase, which are phosphorylated<br />

in response to integrin engagement. Incubation of CL48<br />

liver cells with UDCA-LPE altered the localization of integrin α2,<br />

α3, α5, αv, β1, β4, β5 and β6 as observed by immunofluorescence.<br />

In contrast, incubation with the integrin blocking peptide<br />

RGD reduced integrin protein levels, but did not change its<br />

localization. After UDCA-LPE treatment integrins were transported<br />

to the ER and the nuclear envelop, which led to a loss of<br />

co-localization of integrins with SRC resulting in dephosphorylation<br />

of FAK (Tyr 925 and Tyr 576/577) and SRC (Tyr416).<br />

This translocalization of integrins was not induced by UDCA<br />

and/or LPE treatment, but was exclusively achieved by UDCA-<br />

LPE. To further dissect whether UDCA-LPE would mediate a<br />

shift of integrins to certain membrane microdomains such as<br />

lipid raft and non-raft regions lipid fractioning was performed.<br />

Integrin α2, α3, α5, αv, β1, β4 and β6, but not SRC, was<br />

significantly reduced in transitional fractions and increased in<br />

lipid raft fractions. Taken together, UDCA-LPE disturbed integrin-FAK<br />

signaling after short activation by modifying the localization<br />

of integrins. UDCA-LPE further led to a shift of TGFβ1<br />

receptor I and II to the ER and the nuclear envelope resulting<br />

in dephosphorylation of Smad2/3. According to lipid fractionation<br />

TGFβ1 receptor I and II were also reduced in transitional<br />

fractions and increased in lipid raft fractions after UDCA-LPE<br />

treatment. Analysis by HPLC-MS revealed that UDCA-LPE shows<br />

an integrated fractionation of UDCA and LPE, suggesting that<br />

the localization of UDCA-LPE depends on both UDCA and LPE<br />

end. CONCLUSIONS: UDCA-LPE mediated translocation of<br />

integrins and TGFβ1 receptors into lipid rafts leads to a loss of<br />

colocalization with their down-stream signaling proteins SRC<br />

and Smad2/3. By inhibiting crucial pro-fibrogenic signaling<br />

pathways UDCA-LPE emerges as a promising experimental<br />

drug-candidate for the treatment of liver fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jie Su, Hongying Gan-Schreier,<br />

Walee Chamulitrat, Wolfgang Stremmel, Anita Pathil<br />

1412<br />

Curcumin blocks the loss of lipid droplets in activated<br />

hepatic stellate cells by increasing perilipin 5 gene<br />

expression in vitro<br />

DingYu Zhang; Hepatology, Wuhan Medical Treatment Center,<br />

Wuhan, China<br />

BACKGROUND: Activation of hepatic stellate cells(HSCs) are<br />

the major players during liver fibrogenesis, and featured by the<br />

loss of intracellular lipid droplets(LD). Currently, accumulating<br />

evidence showed that blockade of LD loss could inhibit HSC<br />

activation including decreasing cell proliferation and extracellular<br />

matrix production. Curcumin, a phytochemical from<br />

turmeric, has been demonstrated to inhibit HSC activation and<br />

protect the liver from fibrogenesis in vitro and in vivo. However,<br />

whether curcumin affects intracellular LD formation in<br />

HSCs and underlying mechanism remain largely undefined.<br />

AIM AND HYPOTHESIS: The aims of this study are to evaluate<br />

roles of curcumin in the formation of LD in HSC, and to further<br />

explore the underlying mechanisms. It is hypothesized that curcumin<br />

might eliminate the loss of LD in activated HSCs, rather<br />

than hepatocytes, by increasing perilipin 5(plin5) gene expression<br />

and inducing lipogenesis. RESULTS: Curcumin restored<br />

intracellular LD formation, and increased triglyceride (TG) and<br />

free fatty acid(FFA) levels as well as induced expression of<br />

genes closely relevant to lipogenesis including SREBP-1c, fatty<br />

acid synthase(FAS), PPAR-coactivator (PCA), in passaged mice<br />

HSCs. In contrast, under the same condition, the LD formation<br />

and lipid contents have been unaffected by curcumin in<br />

mice hepatocytes. Furthermore, curcumin dose-dependently<br />

increased plin5 gene expression which has been suppressed<br />

in activated state in HSCs. Meanwhile, transduction with lentiplin5-YFP<br />

in passaged HSCs had the similar effect as curcumin<br />

treatment. On the contrary, in HSCs transfected with plin5-<br />

siRNA prior to curcumin treatment, LD contents and lipid levels<br />

were not changed significantly compared with the normal cells.<br />

CONCLUSIONS: Our results demonstrate that curcumin restores<br />

the LD formation in activated HSCs in vitro, which is carried<br />

out by inducing lipid synthesis via increasing lipogenesis-related<br />

gene expression. The process is mediated by activating<br />

plin5 gene expression. Our results provide novel insight into<br />

the mechanism of the role of curcumin in the inhibition of HSC<br />

activation and hepatic fibrogenesis and potential therapeutic<br />

strategies for treatment of hepatic fibrogenesis without increasing<br />

lipotoxicity in hepatocytes.<br />

Disclosures:<br />

The following authors have nothing to disclose: DingYu Zhang


900A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1413<br />

Versican: A Novel Modulator of Hepatic Fibrosis<br />

Terence N. Bukong 1 , Sean B. Maurice 1,2 , Barinder Chahal 2,1 ,<br />

David Schaeffer 2 , Paul J. Winwood 2,1 ; 1 Northern Medical Program,<br />

University of Northern BC, Prince George, BC, Canada;<br />

2 University of British Columbia, Vancouver, BC, Canada<br />

Little is known about the deposition and turnover of proteoglycans<br />

in liver fibrosis, despite their abundance in the extracellular<br />

matrix. Versican plays diverse roles in modulating cell<br />

behaviour in other fibroproliferative diseases, but remains<br />

poorly described in the liver. Hepatic fibrosis was induced by<br />

carbon tetrachloride (CCl 4<br />

) treatment of C57BL/6 mice over<br />

4 weeks followed by recovery over a 28 day period. Primary<br />

mouse hepatic stellate cells (HSCs) were activated in culture<br />

and versican was transiently knocked down in human (LX2)<br />

and mouse HSCs. Expression of versican, A Disintegrin-like and<br />

Metalloproteinase with Thrombospondin-1 motifs (ADAMTS)-1,<br />

-4, -5, -8, -9, -15 and -20, and markers of fibrogenesis were<br />

studied using qRT-PCR and Western blotting. CCl 4<br />

treatment<br />

led to significant increases in versican expression and the proteoglycanases<br />

ADAMTS-5, -9, -15 and -20, alongside TNF-α,<br />

alpha-smooth muscle actin (α-SMA), collagen-1, and TGF-β<br />

expression. During recovery, expression of many of these genes<br />

returned to control levels. However, expression of ADAMTS-5,<br />

-8, -9 and -15 showed delayed increases in expression at 28<br />

days of recovery which corresponded with decreases in versican<br />

V0 and V1 cleavage products (G1-DPEAAE 1401 and<br />

G1-DPEAAE 441 ). Activation of primary HSCs in-vitro significantly<br />

increased versican, α-SMA, and collagen-1 expression.<br />

Transient knockdown of versican in HSCs led to decreases in<br />

markers of fibrogenesis and reduced cell proliferation, without<br />

inducing apoptosis. Versican expression increases during<br />

HSC activation and liver fibrosis, and proteolytic processing<br />

occurs during the resolution of fibrosis. Knockdown <strong>studies</strong> in<br />

vitro suggest a possible role of versican in modulating hepatic<br />

fibrogenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Terence N. Bukong, Sean B.<br />

Maurice, Barinder Chahal, David Schaeffer, Paul J. Winwood<br />

1414<br />

Hepatic stellate cell-derived retinoid enhances functional<br />

maturity of human embryonic stem cell-derived hepatocytes<br />

on drug metabolism<br />

Hyuk-Soo Eun, Wonhyo Seo, Won-IL Jeong; KAIST, Daejeon,<br />

Korea (the Republic of)<br />

Background: Human embryonic stem cell-derived hepatocyte<br />

(hES-Hep) could be used for the screening tool of drug hepatotoxicity.<br />

However, drug metabolic functions of hES-Hep<br />

are still poor. Retinoids including retinol and its metabolites<br />

play important roles not only in organ development but also<br />

metabolic regulation and they are enriched in hepatic stellate<br />

cells (HSCs) of liver. Therefore, in this study, we evaluate the<br />

effects of mouse HSCs (mHSCs) on the expression and activity<br />

of drug metabolic enzymes in hES-Hep. Methods: Analyses<br />

of microarray, co-culturing with mHSCs or human HSC cell<br />

lines (LX-2 and hTERT), retinoid treatment, flow cytometry analysis,<br />

protein and gene expression, and drug metabolic activity<br />

were performed in hES-Hep. Results: In microarray analyses,<br />

gene expression of drug metabolic enzymes such as CYP1A2,<br />

CYP2C9, CYP2D6, CYP2E1 and CYP3A4 in hES-Hep was<br />

specifically down-regulated compared with normal hepatocyte.<br />

However, that expression was significantly increased in<br />

co-cultured hES-Hep with retinol-storing mHSCs compared to<br />

control hES-Hep. In co-cultured mHSCs, expression of LRAT was<br />

down-regulated while expression of retinaldehyde dehydrogenase-1<br />

(Raldh1) and cellular retinoic acid binding protein-1<br />

was significantly increased, indicating decreased retinol storage<br />

but increased production and delivery of retinoic acids in<br />

HSCs. Similarly, direct treatments of retinoic acids increased<br />

expression of metabolic enzymes in hES-Hep compared with<br />

control hES-Hep. In contrast, retinoic acid deficiency in LX-2,<br />

hTERT and Raldh1-depleted HSCs did not affect gene expression<br />

of metabolic enzymes in hES-Hep, suggesting that retinoic<br />

acid might be a critical factor for functional maturation of hES-<br />

Hep. In flow cytometry, two biggest populations of hES-Hep<br />

depending on cell size and granularity expressed major types<br />

of drug metabolic enzymes and one smallest population still<br />

highly expressed mesodermal markers such as SOX17, FOXA2<br />

and GATA4. In western blotting, protein levels of CYP1A2,<br />

CYP2E1, and CYP3A4 were increased in co-cultured hES-Hep<br />

with retinol-storing HSCs. Consistently, metabolic activities of<br />

CYP1A2 (phenacetin), CYP3A4 (testosterone) and CYP2E1<br />

(chlorozoxazone) were significantly improved in co-cultured or<br />

retinoic acid-treated hES-Hep compared with controls. Conclusions:<br />

Based on our data, retinol-storing mHSCs could enhance<br />

functional maturation of hES-Hep through retinoic acid-mediated<br />

increased expression and activity of drug metabolic<br />

enzymes. Thus, co-culturing with mHSCs or treatments of retinoic<br />

acids could be useful ways to adapt immature hES-Hep for<br />

the application of drug toxicity screening system.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hyuk-Soo Eun, Wonhyo Seo,<br />

Won-IL Jeong<br />

1415<br />

Calcium Mobilization through L-type Channels in<br />

Hepatic Stellate Cell is Essential for TGF-b-mediated<br />

CTGF Induction in Pyk2-dependent Manner<br />

Jonghwa Kim 1 , SoHee Kang 1 , Soohyun Park 1 , Ju-Yeon Cho 1 , Won<br />

Sohn 1 , David A. Brenner 2 , Yong-Han Paik 1 ; 1 Samsung medical<br />

center, Seoul, Korea (the Republic of); 2 UC San Diego, La Jolla,<br />

CA<br />

Background: In hepatic fibrogenesis hepatic stellate cell (HSC)<br />

is a major cell type responsible for producing a major profibrogenic<br />

cytokine TGF-b, and connective tissue growth factor<br />

(CTGF), a major fibrogenic mediator in several organs. The<br />

multi-functional nature of TGF-b signaling in hepatic fibrogenesis<br />

is still elusive. At the previous AASLD (2014) we reported<br />

that Pyk2 is essential for TGF-b-mediated, Smad-independent<br />

CTGF induction. Pyk2 is known to be calcium-sensitive, and<br />

TGF-b was reported to increase intracellular calcium level.<br />

Therefore, we investigated if TGF-b activates Pyk2 by increasing<br />

intracellular calcium levels through L-type voltage-gated<br />

calcium channel in hepatic stellate cell. Methods: Immortalized<br />

human stellate cell line, LX-2, has been cultured. After TGF-b<br />

treatment, expression of CTGF and a-SMA were assessed with<br />

RT-PCR and western blot. Pharmacological inhibitor and siR-<br />

NA-mediated knockdown were used to modulate the activities<br />

and expression levels of protein. Intracelllular calcium mobilization<br />

was measured with Fura-2/AM. Activation of Pyk2<br />

was addressed in western blot using different phosphorylation<br />

site-specific antibodies. Results: CTGF expression was up-regulated<br />

within 1hr in TGF-b stimulated LX-2. This up-regulation<br />

was greatly suppressed by siRNA-mediated knockdown and<br />

pharmacological inhibitor of Pyk2. TGF-b treatment increased<br />

phosphorylation of Pyk2 on tyrosine 402, 579/580, and 881.<br />

Consistent with the previous reports, TGF-b increased intracellular<br />

calcium concentration in Fura-2-preloaded LX-2. CTGF


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 901A<br />

induction by TGF-b was blocked in dose-dependent manner<br />

by pre-treatment with BAPTA-AM, an intracellular calcium chelator.<br />

Treatment of LX-2 with A23187, a calcium ionopore,<br />

increased CTGF induction in the absence of TGF-b, suggesting<br />

that increase of intracellular calcium level is enough to induce<br />

CTGF expression. In addition, A23187 increased activation of<br />

Pyk2 in western. The CTGF induction by A23187 is also greatly<br />

reduced by siRNA of Pyk2. Pre-treament of LX-2 with Nifedipine<br />

(an L-type calcium channel blocker) suppressed CTGF induction<br />

by TGF-b in dose-dependent manner, while FPL64176 (L-type<br />

channel activator) increased CTGF expression without TGF-b.<br />

The CTGF up-regulations by TGF-b, A23187, and FPL64176<br />

were all suppressed by siRNA-mediated knockdown and pharmacological<br />

inhibitors of Pyk2. Conclusions: In hepatic stellate<br />

cell, TGF-b increases the intracellular calcium level through<br />

L-type calcium channel, leading to downstream Pyk2 signaling<br />

for CTGF induction.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jonghwa Kim, SoHee Kang,<br />

Soohyun Park, Ju-Yeon Cho, Won Sohn, David A. Brenner, Yong-Han Paik<br />

1416<br />

Paracrine suppression of proliferation and activation of<br />

hepatic myofibroblasts by adipose-derived stem cells<br />

Danial Afsharzadeh 2 , Shiva Shajari 1 , Linda Brouwer 2 , Han<br />

Moshage 1 , Martin C. Harmsen 2 , Klaas Nico Faber 1 ; 1 MDL, University<br />

medical center Groningen, Groningen, Netherlands; 2 Medical<br />

Biology, University Medical Center Groningen, Groningen,<br />

Netherlands<br />

BACKGROUND AND AIM: Chronic liver diseases lead to liver<br />

fibrosis that may progress to cirrhosis and predisposes for liver<br />

cancer. Liver fibrosis is the result of excessive extracellular<br />

matrix production by hepatic myofibroblasts that may originate<br />

from hepatic stellate cells (HSC) and portal myofibroblasts<br />

(PMF). Though fibrosis is reversible, no drug-based therapy is<br />

available to cure fibrosis and liver transplantation is the only<br />

life-saving option for patients with advanced cirrhosis. Transplantation<br />

of mesenchymal stem cells has been shown to significantly<br />

improve liver function in end-stage liver cirrhosis. Among<br />

mesenchymal stem cells, adipose-derived stem cells (ADSC)<br />

are especially attractive in the context of future clinical applications.<br />

Indeed, recent <strong>studies</strong> demonstrate that ADSC ameliorate<br />

chemical- and diet-induced liver fibrosis, but the mechanistic<br />

understanding is limited. It remains to be determined whether<br />

ADSC produced factors that act in a paracrine fashion or that<br />

the ADSCs engraft and differentiate to functional liver cells.<br />

Here, we investigated the interaction between human ADSC<br />

and primary rat HSC and PMF. METHODS: Human ADSC were<br />

isolated from Human subcutaneous adipose tissue and primary<br />

rat HSC and PMF from Wistar rats. ADSC and HSC or PMF<br />

were cocultured in a transwell system. Moreover, HSC and<br />

PMF were cultured in conditioned medium from ADSC (24<br />

h culture). Cell proliferation and migration was assessed by<br />

real time cell analysis (xCELLigence). Expression of markers<br />

of fibrosis, e.g. Col1a1 and Acta2, was analyzed by Q-PCR.<br />

Chemokine receptor antagonists (SB225002, NBI74330 and<br />

AMD 3465) were used to block CXCR2, CXCR3 and CXR4,<br />

respectively. RESULTS: ADSC showed a strong migratory<br />

potential towards activated HSC and PMF, but not to quiescent<br />

HSC. None of the individual chemokine receptor antagonists<br />

affected ADSC migration towards HSC/PMF. Only the combination<br />

of all 3 antagonists showed a partial inhibition of the<br />

migratory phenotype of ADSC. Proliferation of activated HSC<br />

and PMF, as well as the expression of Col1a1 and Acta2, was<br />

suppressed when cocultured with ADSC. A similar effect was<br />

observed when monocultures of activated HSC and PMF were<br />

exposed to conditioned medium of ADSC. CONCLUSIONS:<br />

ADSC are highly migratory towards hepatic myofibroblasts<br />

through a combination of chemokine signaling pathways.<br />

ADSC produce factors that suppress proliferation and activation<br />

of hepatic myofibroblasts in a paracrine manner. Identifying<br />

the antifibrotic factor(s) in the ADSC secretome holds<br />

promise for novel antifibrotic therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Danial Afsharzadeh, Shiva Shajari,<br />

Linda Brouwer, Han Moshage, Martin C. Harmsen, Klaas Nico Faber<br />

1417<br />

WITHDRAWN<br />

1418<br />

Deactivation of human and rat cirrhotic hepatic stellate<br />

cells: a novel alternative use for the glucagon-like 1<br />

receptor agonist Liraglutide.<br />

Fernanda C. de Mesquita 1,2 , Sergi Guixé-Muntet 1 , Jose Luis<br />

Rosa 3 , Jaime Bosch 1 , Jarbas R. de Oliveira 2 , Jordi Gracia-Sancho<br />

1 ; 1 Barcelona Hepatic Hemodynamic Lab, IDIBAPS - Hospital<br />

Clinic de Barcelona - CIBEREHD, Barcelona, Spain; 2 Laboratório<br />

de Biofísica Celular e Inflamação, PUCRS, Porto Alegre-RS, Brazil;<br />

3 Departament de Ciències Fisiològiques II, Institut d’Investigació<br />

Biomèdica de Bellvitge (IDIBELL), University of Barcelona, L’Hospitalet<br />

de Llobregat, Spain<br />

Background and Aims: Activation of hepatic stellate cells<br />

(HSC) plays a key role in hepatic fibrogenesis, thus improving<br />

HSC phenotype may help to promote fibrosis regression<br />

and ameliorate chronic liver disease. Glucagon-like peptide 1<br />

(GLP-1) receptor agonists, like liraglutide, are well established<br />

in the control of type 2 diabetes. Considering recent <strong>studies</strong><br />

demonstrating anti-inflammatory and anti-oxidant properties<br />

to this kind of drugs, our study aimed to evaluate the effects of<br />

liraglutide on activated HSC phenotype. Methods: Liraglutide<br />

(10-50mM), or its vehicle, was administered to human and rat<br />

cirrhotic HSC (24h and 72h). HSC activation phenotype was<br />

assessed by mRNA and protein expression of alpha smooth<br />

muscle actin (α-SMA) and type I collagen. Effects of liraglutide<br />

on HSC viability and proliferation were evaluated by cell counting<br />

and platelet-derived growth factor receptor (PDGFR) expression.<br />

Results: Liraglutide markedly improved HSC phenotype as<br />

demonstrated by the down-regulation of a-sma (-40% mRNA,<br />

-40% protein expression vs. vehicle) and collagen (-60% mRNA<br />

expression) after 72 hours of treatment. Liraglutide did not<br />

affect HSC viability but significantly diminished cell proliferation<br />

(-32% in cell proliferation, -50% in PDGFR expression vs.<br />

vehicle-treated cells). All p


902A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1419<br />

Serine Protease Omi/HtrA2 Reduces Oxidative Stress<br />

and Preserves Mitochondrial Function in Liver Fibrosis<br />

Seung Kew Yoon 1,2 , Wonhee Hur 1 , Joon Ho Lee 1 , Sung Woo<br />

Hong 1 , Jung-Hee Kim 1 , Jung Eun Choi 1 , Sung Min Kim 1 , Eun Byul<br />

Lee 1 ; 1 The Catholic University Liver Research Center & WHO Collaborating<br />

Center of Viral Hepatitis, Seoul, Korea (the Republic<br />

of); 2 Department of Internal Medicine, St. Mary’s Hospital, The<br />

Catholic University of Korea, Seoul, Korea (the Republic of)<br />

Background: Liver fibrosis is one of chronic liver diseases<br />

caused by various causes. A mitochondrial serine protease<br />

Omi/high temperature requirement protein A2 (HtrA2) is<br />

involved in several disease, but has been poorly understood<br />

in relation to liver fibrosis. The aim of this study is to investigate<br />

the role of Omi/HtrA2 in the liver fibrosis. Methods: Protein<br />

expression of Omi/HtrA2 was analyzed with liver tissues<br />

from CCl - 4<br />

induced mice and liver cirrhosis patients. Reactive<br />

oxygen species (ROS) damage was measured by H 2<br />

DCF-DA<br />

and mito-Sox staining in isolated hepatocytes from mnd2 mice<br />

which have the S276C mutation in the protease domain of<br />

Omi/HtrA2. The Omi/HtrA2 gene was delivered by hydrodynamic<br />

injection in the CCl 4<br />

induced-liver fibrosis mice model<br />

which had lower Omi/HtrA2 expression. The change of collagen<br />

accumulation was measured by Masson’s trichrome (MT)<br />

staining and hydorxyproline assay. Morphological alteration<br />

of mitochondria was analyzed by using transmission electron<br />

microscopy (TEM). Result: Protein expression of Omi/HtrA2<br />

was decreased in liver tissues from CCl 4<br />

induced mice and<br />

liver cirrhosis patients. In mnd2 mice, oxidative stress increased<br />

compared with wild type mice and abnormal morphology of<br />

mitochondria was observed under TEM. Mitochondrial damage<br />

induced with CCl 4<br />

was protected by hydrodynamic based<br />

gene delivery of Omi/HtrA2. In addition, the overexpression<br />

of Omi/HtrA2 decreased collagen accumulation in the liver<br />

tissue. Conclusion: Our findings suggest that Omi/HtrA2<br />

overexpression modulates mitochondrial ROS generation in<br />

liver fibrosis. Omi/HtrA2 gene delivery in liver fibrosis mice<br />

model showed anti-oxidative effect through preserving liver<br />

mitochondrial functions. Therefore, the modulation of Omi/<br />

HtrA2 through the mitochondrial antioxidant defense mechanism<br />

might be promising in anti-fibrotic therapeutic approach.<br />

Disclosures:<br />

The following authors have nothing to disclose: Seung Kew Yoon, Wonhee Hur,<br />

Joon Ho Lee, Sung Woo Hong, Jung-Hee Kim, Jung Eun Choi, Sung Min Kim,<br />

Eun Byul Lee<br />

1420<br />

Cancer associated fibroblasts in intrahepatic cholangiocarcinoma<br />

is derived from portal fibroblasts, but<br />

not hepatic stellate cells; immunohistochemical study in<br />

human tissues<br />

Naoki Uyama 1 , Rei A. Ito 1 , Ami Kurimoto 1 , Tomohiro Okamoto 1 ,<br />

Akito Yada 1 , Hideaki Sueoka 1 , Seikan Hai 1 , Hisashi Kosaka 1 ,<br />

Yuichi Kondo 1 , Ikuo Nakamura 1 , Kazuhiro Suzumura 1 , Yasukane<br />

Asano 1 , Toshihiro Okada 1 , Tadamichi Hirano 1 , Junichi<br />

Yamanaka 1 , Norifumi Kawada 2 , Jiro Fujimoto 1 ; 1 Hepato-biliary-pancreas<br />

surgery, Hyogo College of Medicine, Nishinomiya,<br />

Japan; 2 Department of Hepatology, Osaka City University Graduate<br />

School of Medicine,, Osaka, Japan<br />

(Background and aim) Cancer associated fibroblasts (CAFs),<br />

a key player in the extracellular matrix production of cancer<br />

tissue, is known to be involved in the regulation of cancer<br />

behavior. Although origin of CAFs in intrahepatic cholangiocarcinoma<br />

(ICC) is suspected to be composed of hepatic<br />

stellate cells (HSCs), portal fibroblasts (PF), or bone marrow<br />

derived cells such as fibrocytes, detailed analysis has lacked<br />

so far. Until now, it has been reported that fascin is a specific<br />

marker for HSC and fibrin-2 and Thy-1 are specific markers<br />

for PFs, while α-SMA and PDGFRβ are expressed by both<br />

cells. However, there has been no report on the expression<br />

of PDGFRα in hepatic mesenchymal cells. In this study, we<br />

performed immunohistochemistry in order to characterize<br />

CAFs in human ICC. (Materials and Methods) Ten pairs of ICC<br />

cancerous and noncancerous tissues were prepared. Tissue<br />

sections were Zinc-fixed and paraffin-embedded. Immunohistochemistry<br />

of α-SMA, PDGFRα, PDGFRβ, fascin, fibrin-2 and<br />

Thy-1 in cancerous and noncancerous area was performed.<br />

As for noncancerous area, peri-portal area and sinusoidal<br />

area were intensively observed. To investigate the expression<br />

of individual markers in PFs, HSCs and CAF, double-labeled<br />

immunofluorescent staining with α-SMA was performed under<br />

confocal microscope. (RESULTS) In immunohistochemistry of<br />

noncancerous area of ICC, PDGFRα expression was detected<br />

at periportal area, but not in sinusoidal area. Double immunofluorescent<br />

staining for PDGFRα and α-SMA revealed that both<br />

protein expressions were overlapped in PFs, but not in HSCs,<br />

indicating PDGFRα as a novel marker for PFs. Further immunohistochemical<br />

analyses showed PFs are characterized as<br />

α-SMA + , PDGFRβ + , fascin - , fibrin-2 + and Thy-1 + and HSCs are<br />

as α-SMA + , PDGFRβ + , fascin + , fibrin-2 - and Thy-1 - . In cancerous<br />

area, immunohistochemistry revealed that cancer stroma<br />

was positive for all of α-SMA, PDGFRα, PDGFRβ, fibrin- and<br />

Thy-1. As for fascin, 3 samples were positive, but the others<br />

were negative. α-SMA expression was overlapped with any of<br />

PDGFRα, PDGFRβ, fibrin-2 or Thy-1 in CAFs. (Conclusion) Our<br />

human study revealed that PDGFRα is a specific marker for<br />

PFs. HSCs are α-SMA + , PDGFRβ + , fascin + , fibrin-2 - , Thy-1 - and<br />

PDGFRα - , while PFs are α-SMA + , PDGFRβ + , fascin - , fibrin-2 + ,<br />

Thy-1 + and PDGFRα + . CAFs associated with ICC are positive<br />

for PF markers (fibrin-2 + , Thy-1 + and PDGFRα + ) in addition to<br />

α-SMA and PDGFRβ. Taken together, it is plausible that CAFs<br />

are derived from PFs rather than HSCs.<br />

Disclosures:<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

The following authors have nothing to disclose: Naoki Uyama, Rei A. Ito, Ami<br />

Kurimoto, Tomohiro Okamoto, Akito Yada, Hideaki Sueoka, Seikan Hai, Hisashi<br />

Kosaka, Yuichi Kondo, Ikuo Nakamura, Kazuhiro Suzumura, Yasukane Asano,<br />

Toshihiro Okada, Tadamichi Hirano, Junichi Yamanaka, Jiro Fujimoto<br />

1421<br />

Human neutrophil peptide-1 enhances alcohol-induced<br />

hepatocyte apoptosis in vivo and in vitro.<br />

Rie Ibusuki 2,1 , Hirofumi Uto 2,3 , Masaya Kozono 2 , Kohei Oda 2 ,<br />

Akihiko Oshige 2 , Kotaro Kumagai 2 , Seiichi Mawatari 2 , Tsutomu<br />

Tamai 2 , Akihiro Moriuchi 2 , Hirohito Tsubouchi 4,5 , Takezaki<br />

Toshiro 1 , Akio Ido 2,4 ; 1 Department of International Island and<br />

Community Medicine, Kagoshima University Graduate School of<br />

Medical and Dental Sciences, Kagoshima, Japan; 2 Digestive and<br />

Lifestyle Diseases, Department of Human and Environmental Science,<br />

Kagoshima University Graduate School of Medical and Dental<br />

Sciences, Kagoshima, Japan; 3 Center for Digestive and Liver<br />

Diseases, Miyazaki Medical Center Hospital,, Miyazaki, Japan;<br />

4 Department of HGF Tissue Repair and Regenerative Medicine,<br />

Kagoshima University Graduate School of Medical and Dental Sciences,<br />

Kagoshima, Kagoshima, Japan; 5 Kagoshima City Hospital,,<br />

Kagoshima, Japan<br />

Background Neutrophil infiltration of the liver is a typical<br />

feature of alcoholic liver injury and nonalcoholic steatohepatitis.<br />

Human neutrophil peptide (HNP)-1 is an antimicrobial<br />

peptide secreted by neutrophils. We previously reported that


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 903A<br />

transgenic expression of HNP-1 enhances hepatic fibrosis<br />

and hepatocyte apoptosis in a murine steatohepatitis model<br />

established by ethanol administration or a choline-deficient,<br />

L-amino acid–defined diet. We also showed that HNP-1 exerts<br />

an additive effect on ethanol-induced hepatocyte apoptosis in<br />

SK-Hep1 hepatocellular carcinoma cells in vitro. The aim of<br />

this study was to elucidate the mechanism of hepatocyte apoptosis<br />

induced by HNP-1. Materials and Methods Transgenic<br />

(Tg) mice expressing HNP-1 under the control of a β-actin–<br />

based promoter were established. Ethanol was orally administered<br />

to HNP-1 Tg or wild-type C57BL/6N (WT) mice for 24<br />

weeks. Gene expression and protein levels in liver tissue were<br />

evaluated by RT-PCR and Western blot analysis, respectively.<br />

Apoptosis-related microRNAs were analyzed using miScript<br />

miRNA PCR Array and real-time PCR. SK-Hep1 hepatocellular<br />

carcinoma cells were used to investigate the effect of HNP-1<br />

on hepatocytes in vitro. Results After 24 weeks of ethanol<br />

administration, hepatocyte apoptosis, as assessed by TUNEL<br />

assay, was significantly more severe in TG mice than in WT<br />

mice. Apoptosis was accompanied by higher protein levels of<br />

caspase-3, caspase-8, cleaved PARP, and Bax in liver tissue.<br />

Levels of CD14, TLR4, NFκB-p65, and IL-6 were higher in TG<br />

mice than in WT mice. In addition, p-ASK1, ASK1, p-JNK,<br />

JNK1, JNK2, and CHOP were all more abundant in TG mice<br />

than in WT mice. By contrast, the level of anti-apoptotic Bcl-2<br />

protein in the liver was significantly lower in TG mice than<br />

in WT mice. Analysis of apoptosis-related microRNAs in liver<br />

tissue revealed that miR-34a-5p expression was significantly<br />

higher in TG mice than in WT mice. Furthermore, in the presence<br />

of ethanol, HNP-1 increased the levels of caspase 3 and<br />

Bax and decreased the level of Bcl-2 in a concentration-dependent<br />

manner in vitro. Conclusion HNP-1 secreted by neutrophils<br />

may exacerbate hepatocyte apoptosis following alcoholic liver<br />

injury via elevated microRNA expression.<br />

Disclosures:<br />

The following authors have nothing to disclose: Rie Ibusuki, Hirofumi Uto, Masaya<br />

Kozono, Kohei Oda, Akihiko Oshige, Kotaro Kumagai, Seiichi Mawatari,<br />

Tsutomu Tamai, Akihiro Moriuchi, Hirohito Tsubouchi, Takezaki Toshiro, Akio Ido<br />

1422<br />

Liver Fibrosis in a Mouse Model of Non-alcoholic Fatty<br />

Liver Disease is Mediated by Hepatocyte Hypoxia Inducible<br />

Factor-1<br />

Omar Mesarwi, Mi-Kyung Shin, Shannon Bevans-Fonti, Vsevolod<br />

Polotsky; Johns Hopkins University, Baltimore, MD<br />

Introduction: Obstructive sleep apnea (OSA) is associated with<br />

the progression of non-alcoholic fatty liver disease (NAFLD) to<br />

steatohepatitis and fibrosis. This progression correlates with the<br />

severity of OSA and, specifically, with the burden of hypoxia.<br />

In mice with diet induced obesity, hepatic steatosis leads to<br />

liver tissue hypoxia, which is worse with exposure to intermittent<br />

hypoxia. Intermittent hypoxia in mice also results in more<br />

severe liver fibrosis. We hypothesized that hepatocyte specific<br />

knockout of the oxygen sensing α subunit of hypoxia inducible<br />

factor-1 (HIF-1), a master regulator of the global response to<br />

hypoxia, may be protective against the development of liver<br />

fibrosis. Methods: Hepatocytes were isolated from wild-type<br />

C57BL6/J mice and were exposed to sustained hypoxia (1%<br />

O 2<br />

) or normoxia (16% O 2<br />

) for 24 hours. The culture media<br />

was used to reconstitute type I collagen and the resulting<br />

matrices were examined using confocal microscopy. This was<br />

repeated using hepatocytes from mice with hepatocyte-specific<br />

HIF-1α knockout (Hif1a -/- hep mice). Additionally, wild-type<br />

(n=7) and Hif1a -/- hep (n=10) mice were fed a high trans-fat<br />

diet for six months, to induce hepatic steatosis. Hepatic fibrosis<br />

was evaluated by Sirius red stain and collagen quantification<br />

by hydroxyproline assay. Liver enzymes, fasting insulin,<br />

and hepatic triglyceride content were also assessed. Results:<br />

Culture media from wild-type mouse hepatocytes exposed to<br />

hypoxia allowed for avid collagen cross-linking, but very little<br />

cross-linking was seen when hepatocytes were exposed to normoxia,<br />

or when hepatocytes from Hif1a -/- hep mice were used<br />

in hypoxia or normoxia. Wild-type mice on a high trans-fat diet<br />

had 80% more hepatic collagen than Hif1a -/- hep mice (2.21<br />

mg collagen/mg liver tissue, versus 1.23 mg collagen/mg liver<br />

tissue, p=0.03), which was confirmed by Sirius red staining.<br />

Liver enzymes were similarly elevated in both mouse groups.<br />

Body weight, liver weight, mean hepatic triglyceride content,<br />

and fasting insulin were similar between groups. Conclusions:<br />

HIF-1 is necessary for collagen cross-linking in an in vitro model<br />

of fibrosis. HIF-1 also mediates an increase in liver fibrosis<br />

in mice with diet induced obesity, even in the absence of IH<br />

exposure, perhaps due to liver tissue hypoxia in hepatic steatosis.<br />

This suggests that hepatocyte HIF-1 may play a role in the<br />

progression of liver fibrosis in patients with NAFLD and OSA,<br />

and ongoing work will further delineate the precise mechanism<br />

for this relationship.<br />

Disclosures:<br />

The following authors have nothing to disclose: Omar Mesarwi, Mi-Kyung Shin,<br />

Shannon Bevans-Fonti, Vsevolod Polotsky<br />

1423<br />

Treatment of Hepatic Fibrosis using amniotic fluid<br />

derived mesenchymal stem cells overexpressing interleukin-17<br />

receptor like molecule<br />

Peipei Wang 1,2 , Qiyi Zhao 1,2 , Liang Peng 1,2 , Zhanlian Huang 1,2 ,<br />

Shicheng Su 3 , Zhiliang Gao 1,2 ; 1 Department of Infectious Diseases,<br />

Third Affiliated Hospital of Sun Yat-sen University, Guangzhou,<br />

China; 2 GuangDong Provincial Key Laboratory of Liver Disease,<br />

Third Affiliated Hospital of Sun Yat-sen University, Guangzhou,<br />

China; 3 Sun Yat-sen University, Guangzhou, China<br />

Background: Stem cell transplantation has been shown to have<br />

the capacity to partly reverse liver fibrosis based on its repairing<br />

and reconstructing effects. However, therapeutic effects of liver<br />

fibrosis with genetically modified stem cells which have both<br />

repairing effects and immunotherapy remain uncertain. Our<br />

study aimed to investigated the therapeutic effects and immunological<br />

changes in hepatic fibrosis rats after transplantation<br />

of amniotic fluid derived mesenchymal stem cells over-expressing<br />

interleukin-17 receptor like molecule (IL-17RLM). Materials<br />

and methods: Amniotic fluid derived-mesenchymal stem cells<br />

(AF-MSC) were isolated, and then AF-MSC cell line stably<br />

over-expressing IL-17RLM (AFMSC-IL-17RLM) was established<br />

via lentiviral transfection. Liver fibrosis model was established<br />

in Wista rats after subcutaneous injection of CCl4 twice weekly<br />

for 9 weeks. AFMSC- IL-17RLM were transplanted twice weekly<br />

for 3 weeks via tail vein injection. Results: Genetically modified<br />

stem cells transplantation was safety for liver fibrosis rats; it<br />

could home to the damaged liver after tail vein injection, with<br />

no damage to other organs (e.g. brain, lung, spleen). Lower<br />

hepatic fibrosis score and inflammation grade were observed<br />

in pathological section of AFMSC- IL-17RLM group, compared<br />

with the control groups. IL-17, MCP-1, IL-1α, TNF-α showed<br />

an obvious decrease(p


904A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

among the three groups(p=NS).Th17 cells were obviously<br />

reduced in AFMSC- IL-17RLM group as compared with blank<br />

groups(p6.5) relative to NAFL or NASH subjects<br />

without diabetes (ANOVA p=0.056), and correlated with<br />

HbA1c (r 2 =0.57). FSR of circulating plasma lumican, a proteoglycan<br />

involved in collagen fibril assembly, correlated with<br />

liver collagen FSR (r 2 =0.57) and with liver stiffness (r 2 =0.73).<br />

Conclusion: Hepatic collagen remodeling and synthesis rates<br />

are higher in NASH than NAFL, increase with fibrosis severity<br />

and track with diabetes, a risk factor for fibrosis progression.<br />

Plasma lumican FSR correlates closely with liver collagen FSR<br />

and liver stiffness. Non-invasive kinetic measurements may be<br />

helpful as early read outs of treatment response in anti-fibrotic<br />

trials. The elevated hepatic fibrosis rates in advanced disease<br />

suggest that hepatic scar is dynamic and potentially reversible.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 905A<br />

Disclosures:<br />

Claire Emson - Employment: KineMed<br />

Martin Decaris - Employment: KineMed Inc.<br />

Scott M. Turner - Employment: KineMed, Inc.<br />

Carine Beysen - Employment: KineMed, Inc; Stock Shareholder: KineMed, Inc<br />

Kelvin Li - Employment: KineMed, Inc; Stock Shareholder: KineMed, Inc<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

The following authors have nothing to disclose: Carolyn Hernandez, Jeffrey Y.<br />

Cui, Leander Lazaro, David A. Brenner, Marc Hellerstein<br />

1426<br />

In vivo cell specific gene silencing in the liver using<br />

novel siRNA-loaded nanohydrogel particles<br />

Leonard Kaps; Institute of Translational Immunology and Research<br />

Center for Immunotherapy (FZI), Mainz, Germany<br />

Background: Efficient in vivo transport of active siRNA to specific<br />

liver cells remains challenging. Compared to lipoplex or<br />

polyplex formulations, cationic nanohydrogel particles (NHP)<br />

can serve as superior oligonucleotide carriers [Siegwart DJ et<br />

al, PNAS 2011]. We have designed NHP [Nuhn L et al, ACS<br />

Nano 2012] that serve as stable carriers for in vivo siRNA<br />

delivery [Nuhn L et al, Biomacromolecules 2014]. Results: NHP<br />

composed of block copolymers loaded with negative control<br />

siRNA up to 600 nM siRNA and of a size of 30 to 40 nm did<br />

not show cytotoxicity for murine 3T3 fibroblasts, Raw or MHS<br />

macrophages, HepG2 human liver cells or AML-12 murine<br />

benign hepatocytes. In full media FITC-labelled NHP were<br />

taken up efficiently reaching up to 90% after 1 h incubation as<br />

determined by FACS analysis. NHP loaded with procollagen<br />

a1(I) or CD68 targeted siRNA yielded a robust knockdown<br />

in 3T3 and Raw cells, respectively, after 48 h of incubation<br />

as determined by qPCR. In vivo <strong>studies</strong> using near infrared<br />

labelled NHP loaded with Cy5 labelled procollagen a1(I)<br />

siRNA were performed in lCCl4 treated mice. After i.v. injection<br />

NHP accumulated in the liver (80%), as determined by in<br />

vivo and ex vivo near infrared imaging. IHC and ex vivo FACS<br />

analysis revealed a preferential colocalization with myofibroblasts<br />

(α-SMA+) and macrophages (CD45+ F4/80+ CD11b+).<br />

NHP loaded with procollagen a1(I) siRNA generated an up to<br />

80% in vivo knockdown in CCl4-induced liver fibrosis. Efficient<br />

knockdown was further confirmed by decreased liver collagen<br />

(Sirius red staining and hydroxyproline content). Conclusions:<br />

siRNA loaded NHP accumulate in the liver and especially in<br />

(myo-) fibroblasts and macrophages/Kupffer cells. Due to their<br />

intrinsic specificity for nonparenchymal cells, an apparent lack<br />

of toxicity and a high loading capacity, they are attractive<br />

carriers for siRNA-based antifibrotic or immune modulatory<br />

therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Leonard Kaps<br />

1427<br />

Host microbiota dictates the profibrotic impact of PAMPs<br />

in the liver<br />

Suriguga Suriguga 1 , Floris Fransen 2 , Dorenda Oosterhuis 1 , Geny<br />

M. Groothuis 3 , Paul D. Vos 2 , Henricus A. Mutsaers 1 , Peter Olinga 1 ;<br />

1 Division of Pharmaceutical Technology and Biopharmacy, Department<br />

of Pharmacy, University of Groningen, Groningen, Netherlands;<br />

2 Division of Medical Biology, Department of Pathology and<br />

Medical Biology, University Medical Center Groningen, University<br />

of Groningen, Groningen, Netherlands; 3 Division of Pharmacokinetics,<br />

Toxicology and Targeting, Department of Pharmacy, University<br />

of Groningen, Groningen, Netherlands<br />

Background Pathogen-associated molecular patterns (PAMPs),<br />

e.g. lipopolysaccharide (LPS) and flagellin, might promote liver<br />

fibrosis via the gut-liver axis. However, how these bacterial<br />

components initiate fibrogenesis needs further clarification.<br />

In germ-free (GF) mice, the liver has no history of interaction<br />

with PAMPs, while in colonized (specific pathogen free, SPF)<br />

mice the liver is exposed to PAMPs. Therefore, this study was<br />

designed to investigate the possible fibrogenic effect of PAMPs<br />

in SPF and GF mice, using precision-cut liver slices (PCLS) as<br />

fibrosis model. Methods PCLS from SPF and GF mice were<br />

incubated with or without 10 mg/ml LPS or 50 ng/ml flagellin<br />

for 48h. Viability of PCLS was assessed by measuring their<br />

ATP content. Gene expression of key fibrosis markers Procollagen1α1<br />

(Col1α1), α-Smooth Muscle Actin (α-SMA), Heat<br />

Shock Protein 47 (Hsp47), and Plasminogen Activator Inhibitor-1<br />

(PAI-1) were determined by real-time qPCR. Results Both<br />

SPF and GF mice liver slices were viable up to 48h. Solely in<br />

SPF PCLS, we observed spontaneous onset of fibrosis during<br />

culture up to 48h. Yet, treatment with LPS or flagellin did not<br />

augment the fibrotic response. In contrast, in GF PCLS, Hsp47<br />

expression was significantly up-regulated in the presence of<br />

LPS: 1.4-fold, while Col1α1 and α-SMA tended to increase<br />

compared to control slices incubated for 48h. PAI-1, a marker<br />

for transforming growth factor-β (TGF-β) pathway activation,<br />

was significantly elevated up to 81.8-fold in SPF and 37.2-fold<br />

in GF PCLS after culture (48h) compared to directly after slicing.<br />

Exposure to LPS further increased PAI-1 gene expression in<br />

GF PCLS (2.9-fold), but not in SPF mice. Moreover, in GF liver<br />

slices treatment with flagellin significantly increased expression<br />

of Col1α1 (1.8-fold), α-SMA (1.8-fold), and Hsp47 (1.2-fold),<br />

while it did not effect PAI-1. Conclusion Both LPS and flagellin<br />

could exacerbate fibrogenesis in GF mouse liver slices but not


906A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

in SPF mouse liver slices. These results indicate that PAMPs are<br />

profibrotic in the absence of microbiota, whereas in SPF mice<br />

the microbiota might confer hepatoprotection, which attenuates<br />

the onset of fibrosis by PAMPs in SPF PCLS. Moreover, LPS<br />

may rely on the TGF-β signaling pathway to elicit a fibrotic<br />

response, while flagellin might depend on other pathways.<br />

Disclosures:<br />

The following authors have nothing to disclose: Suriguga Suriguga, Floris Fransen,<br />

Dorenda Oosterhuis, Geny M. Groothuis, Paul D. Vos, Henricus A. Mutsaers,<br />

Peter Olinga<br />

1428<br />

Correlations between hepatic morphometric collagen<br />

content, histologic fibrosis staging, and serum markers<br />

in patients with advanced fibrosis due to nonalcoholic<br />

steatohepatitis (NASH)<br />

Zachary D. Goodman 1 , Lakshmi Alaparthi 1 , Fanny Monge 1 , Keyur<br />

Patel 2 , Rohit Loomba 3 , Stephen H. Caldwell 4 , Eric Lawitz 5 , Stephen<br />

A. Harrison 6 , Mitchell L. Shiffman 7 , Manal F. Abdelmalek 2 ,<br />

Robert P. Myers 8 , Raul E. Aguilar Schall 8 , Mani Subramanian 8 ,<br />

John G. McHutchison 8 , Vlad Ratziu 9 , Arun J. Sanyal 10 , Nezam<br />

H. Afdhal 11 ; 1 Inova Fairfax Hospital, Falls Church, VA; 2 Duke<br />

Clinical Research Institute, Durham, NC; 3 University of California<br />

at San Diego, San Diego, CA; 4 University of Virginia, Charlottesville,<br />

VA; 5 Texas Liver Institute, University of Texas Health Science<br />

Center, San Antonio, TX; 6 San Antonio Military Medical Center,<br />

Fort Sam Houston, TX; 7 Liver Institute of Virginia, Richmond, VA;<br />

8 Gilead Sciences, Inc., Foster City, CA; 9 Hopital Pitie-Salpretriere,<br />

Paris, France; 10 Virginia Commonwealth University, Richmond,<br />

VA; 11 Beth Israel Deaconess Medical Center and Harvard Medical<br />

School, Boston, MA<br />

Background: Our aim was to determine the relationships<br />

between hepatic collagen assessed by morphometry with serum<br />

fibrosis markers, Ishak fibrosis staging, and other histologic<br />

features of NASH in patients with advanced fibrosis. Methods:<br />

The study included adults with NASH and bridging fibrosis<br />

(Ishak stage 3 or 4) or cirrhosis (stage 5 or 6) enrolled in two<br />

phase 2b trials of simtuzumab, a monoclonal antibody against<br />

lysyl oxidase-like-2 (LOXL2). Liver biopsies were graded centrally<br />

according to the NAFLD Activity Score (NAS) and hepatic<br />

collagen in sirius red-stained biopsies was quantified via computer-assisted<br />

morphometry. Serum LOXL2 (sLOXL2) was measured<br />

using an immunoassay (VIDAS® LOXL2; bioMérieux,<br />

Marcy L’Etoile, France). The associations between hepatic collagen<br />

and Ishak fibrosis stage, noninvasive fibrosis markers<br />

(sLOXL2, FibroTest, ELF, APRI, FIB-4, and NAFLD Fibrosis Score<br />

[NFS]), and the NAS and its components were determined.<br />

Results: 429 of 477 randomized patients (89.9%) with biopsies<br />

acceptable for computerized morphometry were included.<br />

The median age was 56 years (IQR 50-60), 62% were female,<br />

52% had cirrhosis, and the median hepatic collagen content<br />

was 8.4% (IQR 5.2-14.7%). Hepatic collagen was moderately<br />

correlated with fibrosis stage (Spearman ρ=0.59; P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 907A<br />

1429<br />

Concerted effects of the two major genetic risk factors<br />

for cholestatic (ABCB4) and fatty liver disease (PNPLA3)<br />

on liver fibrosis: elastography-based study in chronic<br />

liver disease<br />

Marcin Krawczyk 2,1 , Frank Grünhage 2 , Frank Lammert 2 ; 1 Laboratory<br />

of Metabolic Liver Diseases, Department of General, Transplantation<br />

and Liver Surgery, Medical University of Warsaw,<br />

Warsaw, Poland; 2 Department of Medicine II, Saarland University<br />

Medical Center, Homburg, Germany<br />

Background: The liver responds uniformly with liver fibrosis<br />

to any chronic injury. Genome-wide association <strong>studies</strong><br />

have established the common polymorphism p.I148M of the<br />

PNPLA3 gene as the major genetic determinant of non-alcoholic<br />

and alcoholic fatty liver disease. Whereas rare mutations<br />

of the hepatobiliary phosphatidylcholine transporter ABCB4<br />

are known to cause biliary cirrhosis, large-scale whole-genome<br />

sequencing lead to the identification of the common ABCB4<br />

variant c.711A>T as general risk factor for cholestasis and cirrhosis<br />

(Nat Genet 2015). Hence, our hypothesis now was that<br />

the combination of these two most relevant genetic risk factors<br />

should aggravate fibrosis across chronic liver diseases (CLD).<br />

Patients and methods: To address this issue, we studied 713<br />

patients (age 50±12 years, 434 men) recruited for our non-invasive<br />

elastography-based study (J Hepatol 2011), which confirmed<br />

that the PNPLA3 mutation represents a common fibrosis<br />

risk gene. Liver stiffness was determined non-invasively using<br />

transient elastography (TE, Fibroscan). The PNPLA3 c.444C>G<br />

(p.I148M) and ABCB4 c.711A>T (splice region variant) polymorphisms<br />

were genotyped by PCR-based assays. Association<br />

with transient elastography results was tested in contingency<br />

tables, and TE values in carriers of distinct PNPLA3 and ABCB4<br />

genotypes were compared by non-parametric tests. Results:<br />

The ABCB4 c.711 ([TT] n = 23, [AT] n = 204, [AA] n = 486)<br />

and the PNPLA3 p.I148M ([CC] = 382, [CG] = 280 and [GG]<br />

= 50) genotype distributions were in Hardy-Weinberg equilibrium.<br />

The median liver stiffness in the entire cohort was 6.7<br />

kPa and ranged from 2.2 to 75.0 kPa; 156 patients presented<br />

with TE ≥ 13.0 kPa. Of note, the procholestatic allele ABCB4<br />

c.711A was present in the majority of the patients and did not<br />

significantly influence TE per se. The combination of ABCB4<br />

and PNPLA3 risk alleles lead to incremental increases of mean<br />

TE from 5.6 kPa in carriers of 0 to 7.8 kPa in carriers of 4 risk<br />

alleles. In heterozygous carriers of the PNPLA3 risk variant<br />

p.I148M, the presence of the procholestatic ABCB4 geneotype<br />

[AA] was significantly (P = 0.04) associated with increased<br />

TE; in contrast, this effect was absent in PNPLA3 homozygotes.<br />

Conclusions: This is the first study to assess simultaneously the<br />

effects of the currently known most common mutations in two<br />

distinct core pathways simultaneously in patients with CLD.<br />

Our results suggest that the procholestatic ABCB4 variant<br />

might boost the harmful effects of the major PNPLA3 mutation<br />

and vice versa. Prospective <strong>studies</strong> are warranted to evaluate<br />

ABCB4-PNPLA3 compound genotypes as genetic determinants<br />

of liver vulnerability.<br />

Disclosures:<br />

Frank Grünhage - Grant/Research Support: Gilead; Speaking and Teaching:<br />

Falk, Abbvie<br />

The following authors have nothing to disclose: Marcin Krawczyk, Frank Lammert<br />

1430<br />

Innate Lymphoid Cell (ILC) Subset Composition of<br />

Human Liver: Enrichment for ILC1s in Hepatic Cirrhosis<br />

Lucy Golden-Mason 1 , Silvia Giugliano 1 , Linling Cheng 1 , Hugo R.<br />

Rosen 1,2 ; 1 GI/Hepatology, Univ Colorado Denver, Aurora, CO;<br />

2 Veteran’s Affairs Medical Center, Denver, CO<br />

Introduction: Innate lymphoid cells (ILCs) are emerging as<br />

important immune effectors that, in addition to stimulating proand<br />

anti-inflammatory responses, are involved in tissue remodeling.<br />

Three groups of ILCs have been identified in humans.<br />

These include pro-inflammatory Group 1 (ILC1) populations<br />

which produce IFN-γ and ILC3s which produce IL-17 and or<br />

IL-22. ILC2s are characterized by the production of anti-inflammatory<br />

cytokines, predominantly IL-13. In an acute liver injury<br />

model, natural cytotoxicity receptor (NCR) negative ILC3s are<br />

tissue protective. The ILC composition of human liver is unexplored.<br />

We hypothesized that the human liver, an established<br />

site for innate immunity, would harbor pro-inflammatory ILCs<br />

(ILC1/ILC3) which would be increased in hepatic inflammation<br />

and cirrhosis. Methods: Single cell suspensions were prepared<br />

from liver tissues using mechanical disruption and enzymatic<br />

digestion. Our study group consisted of normal (n=12), alcoholic<br />

liver disease (ALD, n=6), cholestatic liver disease (PBC/<br />

PSC, n=6) and HCV (n=18). The majority of subjects (88%)<br />

were Caucasian and 68% were male. Multi-parameter flow<br />

cytometric analysis was used to characterize human hepatic<br />

ILCs. Total ILCs were identified by their lymphoid morphology,<br />

absence of markers for myeloid, T, NK, B and stem cells<br />

and by co-expression of CD45 and the IL-7 receptor (CD127).<br />

ILC subsets were identified by the expression pattern of c-kit<br />

(CD117), NCR NKp44 and CRTH2. Results: In normal liver<br />

ILCs represent 4.45% (1.39-9.1) of CD45-positive lineage<br />

negative lymphoid cells. ILC1 are the predominant subset<br />

detected in normal liver followed by NCR neg ILC3s. Anti-inflammatory<br />

ILC2s comprise 2.39-16.8% of total ILCs in normal<br />

liver. In cirrhotic liver tissues levels of total ILCs are stable in<br />

HCV-induced and cholestatic liver disease-related cirrhosis.<br />

However, in alcohol-induced cirrhosis a greater than 3-fold<br />

increase in ILC levels was observed (p


908A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1431<br />

Risk of cirrhosis in first degree relatives of patients with<br />

NAFLD cirrhosis: A prospective study<br />

Rohit Loomba 1 , Jeffrey Y. Cui 2 , Ricki Bettencourt 3 , Nicholas<br />

Schork 4 , Chi-Hua Chen 5 , Mahdi Al Ikhwan 2 , Archana Bhatt 2 ,<br />

Lisa M. Richards 2 , Claude B. Sirlin 5 , David A. Brenner 2 ; 1 Division<br />

of Gastroenterology and Epidemiology, University of California<br />

at San Diego, La Jolla, CA; 2 Gastroenterology, UCSD School of<br />

Medicine, La Jolla, CA; 3 Family and Preventive Medicine, UCSD<br />

School of Medicine, La Jolla, CA; 4 J Craig Venter Institute, La Jolla,<br />

CA; 5 Radiology, UCSD School of Medicine, La Jolla, CA<br />

Background: The risk of cirrhosis in the first-degree relative of<br />

patients (pts) with NAFLD-cirrhosis is unknown & needs to be<br />

determined. This determination would help develop guidelines<br />

for cirrhosis screening in first-degree relatives of patients with<br />

NAFLD-cirrhosis. Aim: To determine the risk of NAFLD-cirrhosis<br />

in first-degree relatives of pts with known NAFLD cirrhosis.<br />

Methods: This is a cross-sectional analysis of two prospectively<br />

recruited familial cohorts: the familial NAFLD-cirrhosis cohort &<br />

the control-families cohort. The familial cirrhosis cohort included<br />

pts with known NAFLD-cirrhosis (probands) recruited from the<br />

UCSD NAFLD Clinic & their first-degree relatives. The controls<br />

were derived from a community-dwelling cohort of normal<br />

twin/sib-sib/parent-offspring pair with one random participant<br />

as a control & the respective pair as the first-degree relative<br />

recruited from the general population of Southern California.<br />

Probands had documented evidence of NAFLD (based upon<br />

AASLD Guidelines) & either biopsy-proven or imaging-proven<br />

cirrhosis. The first-degree relatives of pts with NAFLD-cirrhosis<br />

& controls underwent a research visit & advanced MRI +MRE<br />

to detect advanced fibrosis (MRE liver stiffness >3.63 kPa has<br />

an AUROC 0f 0.93). Multivariable-adjusted logistic regression<br />

(odds ratio [OR]) was done to examine the risk of cirrhosis in<br />

first-degree relatives of pts with cirrhosis. Results: This study<br />

included 19 probands (mean age 62.5 yrs & BMI 32.3 kg/<br />

m 2 ) with NAFLD-cirrhosis, & 36 first-degree relatives (mean<br />

age 47.4 yrs & BMI 30.9 kg/m 2 ) of probands. The controls<br />

(mean age 40.9 yrs & BMI 25.5 kg/m 2 ) included 61 pairs<br />

(total n=122) of either twin/sib-sib/parent-offspring pair;<br />

with 61 without any evidence of NAFLD or cirrhosis, and 61<br />

first-degree relatives of these randomly ascertained unaffected<br />

first-degree relative pairs. The prevalence of NAFLD-cirrhosis in<br />

the first degree relatives of pts with NAFLD-cirrhosis was significantly<br />

higher than the controls (13.9 % vs. 1.6%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 909A<br />

1433<br />

Safety, Pharmacokinetics, and Biologic Activity of<br />

ND-L02-s0201, a Novel Targeted Lipid-Nanoparticle to<br />

Deliver HSP47 siRNA for the Treatment of Patients with<br />

Advanced Liver Fibrosis: Interim Results from Clinical<br />

Phase 1b/2 Studies<br />

Eric Lawitz 2 , Yasunobu Tanaka 3 , Fred Poordad 2 , Julio A. Gutierrez<br />

2 , Kathy Carr 4 , Wenbin Ying 1 , Yoshiro Niitsu 5 , Kageshi<br />

Maruyama 3 ; 1 NDT, Oceanside, CA; 2 Texas Liver Institute, University<br />

of Texas Health Science Center, San Antonio, TX; 3 Nitto Denko<br />

Corporation, Tokyo, Japan; 4 RRD International, RRD International,<br />

LLC, Rockville, MD; 5 Sapporo Medical University, Sapporo, Japan<br />

Objectives: Therapy for liver fibrosis is a major unmet medical<br />

need. We have reported a novel therapeutic modality that<br />

inhibits expression of the collagen-specific chaperone, heat<br />

shock protein 47 (HSP47), with stellate cell (SC)-specific drug<br />

delivery (Sato et al. 2008 Nature Biotech), with a goal of<br />

reversal of collagen deposition. NDL02s0201 is an injectable<br />

lipid nanoparticle formulation with a novel vitamin A analogue<br />

targeting agent for SC and a chemically modified small interfering<br />

ribonucleic acid (siRNA) active ingredient that inhibits<br />

HSP47. We report safety, PK and preliminary biologic activity<br />

results from a Phase 1a study and an ongoing Phase 1b/2<br />

study. Methods: The Phase 1a study was a randomized, double-blind,<br />

placebo-controlled study to evaluate the safety and<br />

PK of escalating single doses of ND-L02-s0201 from 0.03 to<br />

0.8 mg/kg in normal volunteers. The Phase 1b/2 clinical trial<br />

is an ongoing, open label, multiple-dose, escalation study to<br />

evaluate safety, PK, and biological activity in patients with<br />

METAVIR F3-4. NDL02s0201 is intravenously infused once or<br />

twice weekly for 5 weeks at 0.2 to 0.6 mg/kg/week. Results:<br />

The Phase 1a study included 56 subjects (42 ND-L02-s0201,<br />

14 placebos). Seven subjects treated with ND-L02-s0201 at<br />

dose levels of 0.03 to 0.6 mg/kg had treatment-related mild/<br />

moderate adverse events characteristic of hypersensitivity infusion<br />

reactions. After premedication with H1 and H2 blockers<br />

and a slower infusion rate, no infusion reactions occurred<br />

at 0.8 mg/kg. PK results of the siRNA were linear over the<br />

range of doses studied. The mean elimination half-life across<br />

the 7 doses was 27.3 h with no apparent dose-related trends.<br />

Available data from the Phase 1b/2 study includes 8 subjects<br />

from the 1 st dose cohort: 4 with NASH and 4 with HCV. The<br />

baseline (within 12 months before enroll) METAVIR score was<br />

F3 for 3 subjects and F4 for 5 subjects. Cohort 1 (0.2 mg/kg/<br />

week) of the Phase 1b/2 study was complete. Two possibly<br />

treatment-related gastrointestinal AEs were reported (epigastric<br />

discomfort and abdominal pain). At the lowest planned dose 6<br />

out of 8 subjects had at least 1 stage reduction in Ishak score<br />

(mean reduction 1.25, range 0-4, N=8), and 1 subject showed<br />

reduction of METAVIR score from F-3 to F-0 right after 5 weeks<br />

dosing. Conclusion: ND-L02-s0201 was well tolerated in both<br />

<strong>studies</strong> with no dose limiting toxicities up to 0.8 mg/kg (single)<br />

and of 0.2 mg/kg/week (multiple doses). The PK results were<br />

consistent between healthy subjects and patients as well as for<br />

single and multiple doses with no drug accumulation. Histologic<br />

improvement was seen in the majority of patients at the<br />

lowest dose studied.<br />

Disclosures:<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Wenbin Ying - Management Position: Nitto Denko Technical Corporation<br />

The following authors have nothing to disclose: Yasunobu Tanaka, Julio A. Gutierrez,<br />

Kathy Carr, Yoshiro Niitsu, Kageshi Maruyama<br />

1434<br />

Antifibrotic efficacy of a TGF-β kinase inhibitor on<br />

early-onset and end-stage of fibrosis in precision-cut<br />

human liver slices<br />

Theerut Luangmonkong 1,2 , Suriguga Suriguga 1 , Emilia Bigaeva 1 ,<br />

Dorenda Oosterhuis 1 , Koert P. de Jong 3 , Detlef Schuppan 4,5 , Henricus<br />

A. Mutsaers 1 , Peter Olinga 1 ; 1 Pharmaceutical Technology<br />

and Biopharmacy, University of Groningen, Groningen, Netherlands;<br />

2 Pharmacy, Mahidol University, Bangkok, Thailand; 3 Hepato-Pancreato-Biliary<br />

Surgery and Liver Transplantation, University<br />

Medical Center Groningen, Groningen, Netherlands; 4 Medicine,<br />

Translational Immunology and Research Center for Immunotherapy,<br />

University of Mainz Medical Center, Mainz, Germany; 5 Gastroenterology,<br />

Beth Israel Deaconess Medical Center, Harvard<br />

Medical School, Boston, MA<br />

Background Advanced liver fibrosis is a major cause of liver<br />

related morbidity and mortality, but to date, preclinically effective<br />

antifibrotic compounds have failed to show efficacy in<br />

clinical trials. Galunisertib, a TGF-β receptor kinase inhibitor, is<br />

currently studied for the treatment of cancers, including hepatocellular<br />

carcinoma. Due to its mode of action, galunisertib is<br />

also a potential candidate for the treatment of liver fibrosis. Precision-cut<br />

liver slices (PCLS) are an ideal human model to test<br />

the antifibrotic efficacy of compounds. Therefore, we aimed<br />

to assess the antifibrotic potency of galunisertib in both the<br />

early-onset and end-stage of fibrosis in human PCLS. Methods<br />

Early-onset fibrosis was studied in PCLS prepared from surgical<br />

excess material of donor livers for reduced-size liver transplantations.<br />

End-stage fibrosis was studied in PCLS from explanted<br />

cirrhotic livers of patients undergoing liver transplantation.<br />

PCLS were incubated with galunisertib (0.625-10 μM) for<br />

48h. Viability was assessed by ATP content of the slices, and<br />

gene expression of the key fibrosis markers Procollagenα1(I)<br />

(COL1α1), α-smooth muscle actin (α-SMA), Heat shock protein<br />

47 (HSP47), and Plasminogen activator inhibitor-1 (PAI-1),<br />

was assessed by real-time qPCR. Results Liver slices remained<br />

viable for 48h. In the early-onset of fibrosis, COL1α1 gene<br />

expression increased significantly 3.5-fold, while expression<br />

of α-SMA, HSP47, and PAI-1 remained unaltered. Following<br />

treatment with galunisertib, only the highest concentration lowered<br />

the ATP content by about 20 %, while the fibrosis markers<br />

decreased in a concentration-dependent manner up to 82 ± 3<br />

%, 55 ± 7 %, and 71 ± 4 %, for COL1α1, HSP47, and PAI-1,<br />

respectively. α-SMA gene expression did not show a concentration-dependent<br />

inhibition, and only the highest concentration<br />

of galunisertib significantly decreased gene expression by 22<br />

± 9 %. In the cirrhotic PCLS, COL1α1 gene expression significantly<br />

increased 6.5-fold after 48h of incubation, while the<br />

other genes remained unchanged. The highest concentration<br />

of galunisertib in the early-onset of fibrosis study (10 μM) significantly<br />

inhibited the expression of COL1α1 by 81 ± 11 %<br />

and PAI-1 by 78 ± 4 % in cirrhotic PCLS, without toxicity, while<br />

HSP47 and α-SMA expression remained unchanged. Conclusion<br />

Human PCLS that reflect early and late stage fibrosis,<br />

appear to be a viable tool to predict antifibrotic effects prior to<br />

clinical <strong>studies</strong>. Galunisertib is an attractive potential drug for<br />

the treatment of human liver fibrosis.


910A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Theerut Luangmonkong, Suriguga<br />

Suriguga, Emilia Bigaeva, Dorenda Oosterhuis, Koert P. de Jong, Detlef Schuppan,<br />

Henricus A. Mutsaers, Peter Olinga<br />

1435<br />

Serum lysyl oxidase-like-2 (sLOXL2) levels correlate with<br />

fibrosis stage in patients with nonalcoholic steatohepatitis<br />

(NASH)<br />

Stephen A. Harrison 1 , Zachary D. Goodman 2 , Vlad Ratziu 3 , Rohit<br />

Loomba 4 , Anna Mae Diehl 5 , Eric Lawitz 6 , Holger Hinrichsen 7 ,<br />

Kiran Bambha 8 , Manal F. Abdelmalek 5 , Robert P. Myers 9 , Raul<br />

E. Aguilar Schall 9 , Mani Subramanian 9 , John G. McHutchison 9 ,<br />

Nezam H. Afdhal 10 , Andrew J. Muir 5 ; 1 San Antonio Military<br />

Medical Center, Fort Sam Houston, TX; 2 Inova Fairfax Hospital,<br />

Falls Church, VA; 3 Hôpital Pitié-Salpêtrière, Paris, France; 4 University<br />

of California at San Diego, San Diego, CA; 5 Duke Clinical<br />

Research Institute, Durham, NC; 6 Texas Liver Institute, University<br />

of Texas Health Science Center, San Antonio, TX; 7 Gastroenterologisch-Hepatologisches<br />

Zentrum, Kiel, Germany; 8 University of<br />

Colorado Denver, Aurora, CO; 9 Gilead Sciences, Inc., Foster City,<br />

CA; 10 Beth Israel Deaconess Medical Center and Harvard Medical<br />

School, Boston, MA<br />

Background: LOXL2 plays a central role in liver fibrosis by<br />

catalyzing collagen cross-linkage. Our aim was to examine<br />

the association between sLOXL2 and disease severity in<br />

patients with advanced fibrosis due to NASH. Methods: The<br />

study included adults with NASH and bridging fibrosis (Ishak<br />

stage 3 or 4) or cirrhosis (stage 5 or 6) enrolled in two phase<br />

2b trials of simtuzumab, a monoclonal antibody against lysyl<br />

oxidase-like-2 (LOXL2). sLOXL2 was measured using an immunoassay<br />

(VIDAS® LOXL2; bioMérieux, Marcy L’Etoile, France)<br />

in NASH patients and 50 healthy controls. Liver biopsies were<br />

graded centrally according to the NAFLD Activity Score (NAS)<br />

and fibrosis was staged according to the Ishak classification.<br />

The associations between sLOXL2 and fibrosis stage, NAS,<br />

demographics, body mass index (BMI), diabetes, liver biochemistry,<br />

MELD, and fibrosis markers (FibroTest, ELF, APRI,<br />

FIB-4, NAFLD Fibrosis Score [NFS]) were determined. Results:<br />

474 of 477 randomized patients (99.4%) with available<br />

sLOXL2 and stage 3-6 fibrosis were included. The median age<br />

was 56 years (IQR 50-60), 63% were female, 68% had diabetes,<br />

the median BMI was 33.6 kg/m 2 (IQR 29.8-38.2), and the<br />

median serum ALT was 39 U/L (IQR 28-62). Median sLOXL2<br />

was higher in patients with NASH than controls (107 pg/mL<br />

[IQR 82-147] vs. 81 pg/mL [IQR 71-108]; P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 911A<br />

no change after treatment. All patients were divided equally<br />

into training and testing cohorts. The fibrosis scoring index for<br />

CHC was analyzed against Ishak staging system as the gold<br />

standard. Results: 27 out of 100 parameters showed high-level<br />

correlation with Ishak staging, mainly in the portal and total<br />

collagen compartments. Systematic area under curve (AUC)<br />

optimization analysis further identified 6 parameters with the<br />

highest correlation out of the 27. These 6 parameters were as<br />

efficient (AUC: 0.86–0.94) in identifying differences between<br />

all Ishak brosis stages as using all 27 parameters (AUC: 0.85–<br />

0.94). Systematic analysis performed for the 39 paired biopsies<br />

identified 8, 1 and 2 parameters for the respective Ishak<br />

staging decrease, increase, and no change groups. These 11<br />

parameters from the paired biopsies comprised 7 which were<br />

already identified in the 27 parameters, and 4 new ones. Taking<br />

the best correlating parameter from these 4 new ones with<br />

Ishak staging, and adding it to the 6 highest correlation parameters,<br />

we developed a 7-parameter indexing model (comprising<br />

Strlength, Portal, NoShortStrP, NoStrPA, NoLongStrPA,<br />

StrLengthPA, NoXlinkP). This model can faithfully and reliably<br />

recapitulate Ishak brosis scores in CHC patients, including the<br />

paired pre- and post-treatment biopsies. Conclusions: A modified<br />

qFibrosis model was developed and validated for CHC<br />

patients and for treatment response assessment. An indexing<br />

system with 7 parameters was shown to be highly accurate<br />

and reproducible for the objective and quantitative analysis of<br />

fibrosis progression and/or regression in CHC patients.<br />

Disclosures:<br />

Lai Wei - Advisory Committees or Review Panels: Gilead, AbbVie; Grant/<br />

Research Support: BMS<br />

The following authors have nothing to disclose: Huiying Rao, Feng Liu, Aileen -.<br />

Wee, Bo Feng, Ming Yang<br />

1437<br />

Collagen and tissue turnover as function of age – implications<br />

for fibrosis<br />

Morten A. Karsdal 1 , Federica Genovese 1 , Emilie A. Madsen 1 ,<br />

Tina Manon-Jensen 1 , Detlef Schuppan 2,3 ; 1 Nordic Bioscience, Herlev,<br />

Denmark; 2 Institute of Translational Immunology and Research<br />

Center for Immunotherapy, University Medical Center, Mainz,<br />

Germany; 3 Division of Gastroenterology, Beth Israel Deaconess<br />

Medical Center, Harvard Medical School, Boston, MA<br />

Background: The extracellular matrix (ECM) is the backbone of<br />

all tissues. It is a complex grid consisting of multiple structural<br />

proteins which each play a vital role for the function and maintenance<br />

of normal tissue function. In development and growth,<br />

tissue is being formed and elaborated (tissue modeling), while<br />

in adult life, tissues are being maintained and remodeled.<br />

These processes involve likely different mechanisms. During<br />

tissue modeling and remodeling, small fragments of proteins<br />

are released into the circulation, where they may be used as<br />

biomarkers of tissue turnover. Aim: The aim of the study was<br />

to investigate ECM turnover in rodents as a function of age.<br />

Materials and Methods: Serum of rats of 1, 2, 3, 4, 5, 6,<br />

10 and 12 months of age was profiled for 15 markers of<br />

ECM turnover, including: fragments of type I, II, III, IV, V and<br />

VI collagen formation (P1NP, Pro-C3, P4NP_7S, Pro-C5, Pro-<br />

C6) and degradation (C1M, C2M, C2M-beta, C3M, C4M,<br />

C5M, C6M); biglycan (BGM) and elastin (ELM7) degradation;<br />

and the type I and II collagen telopeptides CTX-I and CTX-II.<br />

Results: Type I and II collagen turnover was up to 93% and<br />

97% downregulated in old (one year) compared to young (one<br />

month) old animals (p


912A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1439<br />

Comparison of 16 blood and/or elastometric fibrosis<br />

tests in 5 causes of chronic liver diseases: too much?<br />

Towards a simplification<br />

Jerome Boursier 1,2 , Alexandra Ducancelle 3,2 , Vincent Leroy 4 , Julien<br />

Vergniol 5 , Nathalie Sturm 6 , Brigitte Le Bail 7 , Jean-Pierre H. Zarski 4 ,<br />

Victor de Ledinghen 5 ; 1 Hepato-Gastroenterology Departement,<br />

University Hospital, Angers, France; 2 HIFIH Laboratory EA3859,<br />

University, Angers, France; 3 Virology Department, University Hospital,<br />

Angers, France; 4 Hepato-Gastroenterology Department,<br />

University Hospital, Grenoble, France; 5 Hepatology Department,<br />

University Hospital, Bordeaux, France; 6 Pathology Department,<br />

University Hospital, Grenoble, France; 7 Pathology Department,<br />

University Hospital, Bordeaux, France<br />

Introduction: Most non-invasive fibrosis tests were developed<br />

in chronic hepatitis C (CHC), but are frequently used in other<br />

causes. A lot of tests are available. Therefore, we compared<br />

the accuracy of several usual tests between main etiologies to<br />

try to determine whether a test could be the most performant<br />

in several etiologies. Methods: Populations included 1660<br />

patients: 698 with CHC, 178 with hepatitis B (CHB), 444<br />

with HIV/CHC, 225 with NAFLD, and 115 with alcoholic liver<br />

disease (ALD). 16 tests (13 blood tests, 1 elastometry by VCTE<br />

-Fibroscan- and 2 combining blood markers and elastometry)<br />

were evaluated. Reference was Metavir fibrosis (F) stage by<br />

liver biopsy. Results: Accuracy was sensitive to biopsy length.<br />

Obuchowski indices, reflecting accuracy for all F stages, were<br />

by decreasing order, CHC: Elasto-FibroMeter VCTE2G/3G : 0.812,<br />

FibroMeter VIRUS2G : 0.797, FibroMeter VIRUS3G : 0.785, CirrhoMeter<br />

VIRUS2G : 0.771, Fibrotest: 0.762, CirrhoMeter VIRUS3G :<br />

0.756, Fibroscan: 0.754, Hepascore: 0.752, FibroMeter ALD :<br />

0.750, APRI: 0.742, FIB4: 0.741. CHB: Obuchowski indices<br />

were roughly the same as in CHC. HIV/CHC: Obuchowski<br />

indices were moderately decreased compared to CHC except<br />

for Fibrotest (p=0.03). NAFLD: 7 tests had a substantial accuracy<br />

loss compared to CHC, especially Fibrotest (p=0.006),<br />

while 5 tests had a substantial gain especially FibroMeter ALD<br />

(p=0.039) and Zeng score (p=0.030). ALD: 2 tests had a<br />

dramatic accuracy loss compared to CHC: APRI (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 913A<br />

1441<br />

Comparing the liver stiffness by transient elastography<br />

in cirrhosis patients with PBC and HBV to predict the<br />

risk of esophageal variceal bleeding<br />

Ying Huang 1 , Haiyun Zhang 1 , Li Xiao 1 , Liqing Zhu 1 , Qingjing<br />

Zhu 2 , Xiaohua Hou 1 , Ling Yang 1 ; 1 Division of Gastroenterology,Department<br />

of Internal Medicine, Huazhong University of Science&Technology,<br />

Wuhan, China; 2 Wuhan Medical Treatment<br />

Center, Wuhan, China<br />

Background&Aim: Staging hepatic fibrosis is critical for the<br />

evaluation and deternination of prognosis in chronic liver diseases.<br />

Variceal bleeding is the servere complications of liver<br />

cirrhosis. Endosocopy as the only means to evaluate the varices<br />

and risk of bleeding is not an easy examination to be<br />

accepted by patients with expensive price and poor tolerance.<br />

Transient elastography (TE) is useful as non-invasive diagnositic<br />

method for predicting liver stiffness and portal hypertension.<br />

Our aim was to analyze the diagnostic accuracy of liver stiffness<br />

measurement(LSM) and the esophageal variceal grading<br />

and the risk of esophageal variceal bleeding in different etiological<br />

cirrhosis patients with HBV infection or primary billiary<br />

cirrhosis by transient elastography. Patients and Methods:<br />

Endoscopy and transient elastography were performed in 24<br />

primary biliary cirrhosis patients and 66 hepatitis B related<br />

cirrhosis. The relationships of the liver stiffness between esophageal<br />

variceal degree and the risk of esophageal variceal<br />

bleeding, as well as the hematological and biochemical laboratory<br />

parameters were analyzed. Results: The grade of<br />

esophageal varices is highly correlated with the stiffness of<br />

the liver(P


914A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Nezam H. Afdhal - Advisory Committees or Review Panels: Trio Helath Care;<br />

Board Membership: Journal Viral hepatitis; Consulting: Merck, EchoSens, BMS,<br />

Achillion, GlaxoSmithKline, Springbank, Gilead, AbbVie; Grant/Research Support:<br />

Gilead; Stock Shareholder: Springbank<br />

Christophe Hézode - Speaking and Teaching: Roche, BMS, MSD, Janssen<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

Steven L. Flamm - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, AbbVie, Janssen, Salix; Consulting: Merck, Janssen, Bristol Myers<br />

Squibb, AbbVie, Salix, Gilead; Grant/Research Support: Janssen, Bristol Myers<br />

Squibb, Gilead, AbbVie; Speaking and Teaching: Salix<br />

Robert P. Myers - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Raul E. Aguilar Schall - Employment: Gilead Sciences, Inc.<br />

Robert H. Hyland - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Michael R. Charlton - Grant/Research Support: GIlead Sciences, Merck, Janssen,<br />

AbbVie, Novartis<br />

Marc Bourlière - Advisory Committees or Review Panels: Schering-Plough,<br />

Bohringer inghelmein, Schering-Plough, Bohringer inghelmein, Transgene; Board<br />

Membership: Bristol-Myers Squibb, Gilead, Idenix; Consulting: Roche, Novartis,<br />

Tibotec, Abott, glaxo smith kline, Merck, Bristol-Myers Squibb, Novartis, Tibotec,<br />

Abott, glaxo smith kline; Speaking and Teaching: Gilead, Roche, Merck, Bristol-Myers<br />

Squibb<br />

1443<br />

How gold is the “gold standard” method for staging<br />

liver fibrosis?<br />

Ioan Sporea; University of Medicine and Pharmacy Timisoara,<br />

Timisoara, Romania<br />

Backgrounds and aim: Liver biopsy (LB) is considered the gold<br />

standard method for staging liver fibrosis. The aim of our study<br />

was to assess the quality of liver samples obtained by percutaneous<br />

LB in a large cohort and to determine the sensitivity of<br />

this method for staging liver fibrosis. Material and method: We<br />

performed a retrospective study on echo-assisted LB performed<br />

in our Department, using 1.4 mm and 1.6 mm Menghini modified<br />

needles, with 2 liver passages. 1012 LBs performed were<br />

analyzed (41.4% male, 58.5% female, mean age of 46±12<br />

years). Based on the length of the liver specimens obtained we<br />

divided the LBs into 4 groups (< 15 mm; 15- 24 mm; 25-39<br />

mm; > 40 mm). We calculated the length means for every<br />

group and usingBedossa’s (1) study we analyzed the percentage<br />

of expectedcorrectly classified biopsies according to the<br />

length of biopsy specimen. Results: The overall mean length of<br />

liver specimen obtained in our cohort was 33±9 mm in size,<br />

with mean number of portal tracts of 20±10. 1%(10) of the LBS<br />

were included in the first group (< 15 mm) with mean length of<br />

98±2 mm. 13% (135) LBS were included in the second group<br />

(15- 24 mm) obtaining a mean length of 20±1.8 mm.41%<br />

(418) of the LBs had between 25 and 39 mm with a mean<br />

length of 30±3 mm. 45% (449) of the LBs obtained specimens<br />

larger than 40 mm with mean length of 42±5 mm.Using<br />

Bedossa’s study (1) and diagram considering the sensitivity of<br />

LB for staging liver fibrosis according to the length of biopsy<br />

specimen we obtained the following sensitivities: Group 1 (<<br />

15 mm) 55%; Group 2 (15-24 mm) 70%; Group 3 (25-39<br />

mm) 75%; Group 4 (> 40 mm) 83% and an overall sensitivity<br />

of 80%. Conclusions:Despite the fact that good liver specimens<br />

are obtained using Menghini needles with 2 passages technique<br />

the overall sensitivity of liver biopsy is only around 80%<br />

using Bedossa criteria. References: 1. Bedossa P et al. Sampling<br />

variability of liver fibrosis in chronic hepatitis C. Hepatology<br />

2003;38:1449–1457.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ioan Sporea<br />

1444<br />

Reproductive Aging, Independent of Chronologic Age,<br />

Is Associated with Accelerated Fibrosis Progression in<br />

HIV/HCV Co-Infected Women<br />

Monika Sarkar 1 , Jennifer L. Dodge 1 , Ruth Greenblatt 1 , Mark H.<br />

Kuniholm 2 , Jack Dehovitz 3 , Michael Plankey 9 , Andrea Kovacs 4 ,<br />

Audrey French 7 , Eric C. Seaberg 8 , Igho Ofotokun 5 , Margaret Fischl<br />

6 , Edgar T. Overton 10 , Erin Kelly 1 , Peter Bacchetti 1 , Marion G.<br />

Peters 1 ; 1 University of California, San Francisco, San Francisco,<br />

CA; 2 Albert Einstein College of Medicine, Bronx, NY; 3 SUNY<br />

Downstate Medical Center, Brooklyn, NY; 4 University of Southern<br />

California, Alhambra, CA; 5 Emory University, Atlanta, GA; 6 University<br />

of Miami, Miami, FL; 7 Cook County Health and Hospitals<br />

System, Chicago, IL; 8 Johns Hopkins University, Baltimore, MD;<br />

9 Georgetown University Medical Center, Washington, DC; 10 University<br />

of Alabama, Birmingham, AL<br />

Cross sectional <strong>studies</strong> note less fibrosis in pre- versus post<br />

menopausal women, though data lack serial fibrosis measures<br />

and adequate adjustment for chronologic aging. Methods:<br />

This is a retrospective cohort study of HIV/HCV co-infected<br />

women, followed in the multicenter Women’s Interagency HIV<br />

Study. Eligible women were pre-menopausal at entry (n=411),<br />

confirmed by detectable anti-mullerian hormone (AMH). Subsequent<br />

AMH loss during follow-up indicated menopausal<br />

transition. We assessed fibrosis rate using the validated fibrosis<br />

marker, FIB4 (level < 1.45 indicates < stage 3 fibrosis).<br />

Fibrosis rate was evaluated in each woman as she transitioned<br />

through menopause using random-intercept-random-slopes,<br />

accounting for chronologic age with flexible linear splines.<br />

Reproductive age was defined as pre- (detectable AMH), peri-<br />

(≤ 5 yrs after AMH loss) and post-menopause (> 5 yrs after<br />

AMH loss). Covariates included race, metabolic factors, alcohol,<br />

and antiretroviral therapy (ART). Results: Median follow-up<br />

was 11.5 (IQR 6-18.5) years with mean age of 44.5 (+/- 7.0)<br />

and median entry FIB4 of 1.71 (IQR 1.2-2.6). Adjusted for<br />

chronologic aging, fibrosis was 0.126 FIB4 per year faster<br />

(95% CI 0.029-0.224, p=0.01) during peri-menopause and<br />

0.149 FIB4 per year faster (95% CI 0.003-0.295, p=0.046)<br />

during post- as compared to pre-menopause. After adjusting<br />

for race, alcohol, and ART, accelerated fibrosis persisted in<br />

peri- compared to pre-menopause (0.103 FIB4/year faster,<br />

95% CI 0.007-0.200, p=0.04), though not significant (0.117,<br />

95% CI -0.029-0.0262, p=0.12) for post- vs pre-menopausal<br />

comparison. Conclusion: In HIV/HCV co-infected women, liver<br />

fibrosis accelerates through peri-menopause, independent of<br />

chronologic aging. Peri-menopausal women may have more<br />

urgent need for HCV therapy. Longitudinal analyses of hepatic<br />

fibrosis across reproductive age in women with non HCV-related<br />

liver disease are warranted.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 915A<br />

when AZA 1nΜ was applied in SFM as well as in TGF induced<br />

HSCs: For COL1A1: SFM 1.45 vs AZA 0.47 (p


916A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1447<br />

A novel canine liver cirrhosis model to develop less<br />

invasive liver regeneration therapy using cultured bone<br />

marrow-derived cells<br />

Takashi Matsuda 1,2 , Taro Takami 1 , Yuki Aibe 1 , Toshihiko Matsumoto<br />

1,4 , Tsuyoshi Ishikawa 1 , Naoki Yamamoto 1,5 , Shuji Terai 3 ,<br />

Isao Sakaida 1,6 ; 1 Department of Gastroenterology and Hepatology,<br />

Yamaguchi University Graduate School of Medicine, Ube,<br />

Yamaguchi, Japan; 2 Sanyo-Onoda General Hospital, Sanyo-Onoda,<br />

Japan; 3 Division of Gastroenterology and Hepatology,<br />

Niigata University, Niigata, Japan; 4 Department of Oncology and<br />

Laboratory Medicine, Yamaguchi University Graduate School of<br />

Medicine, Ube, Yamaguchi, Japan; 5 Yamaguchi University Health<br />

Administration Center, Yamaguchi University, Ube, Yamaguchi,<br />

Japan; 6 Center for Reparative Medicine, Yamaguchi University<br />

Graduate School of Medicine, Ube, Yamaguchi, Japan<br />

Background/Aims: We developed autologous bone marrow<br />

cell infusion (ABMi) therapy for decompensated liver cirrhosis<br />

patients using non-cultured autologous whole bone marrow<br />

(BM) cells aspirated under general anesthesia, and then<br />

reported the safety and efficacy of this approach. We have<br />

also been developing a new, less invasive liver regeneration<br />

therapy using cultured autologous BM-derived mesenchymal<br />

stem cells (BMSCs) from small amounts of BM fluid aspirated<br />

under local anesthesia. Before human clinical trials, the safety<br />

and efficacy of cultured autologous BMSC infusion in mediumto-large<br />

animals must be confirmed; thus, we aimed to develop<br />

a canine liver cirrhosis model. Methods: Eight 1- to 2-year-old<br />

beagles were studied. A small amount of BM fluid was aspirated<br />

from the canine humerus to assess the characteristics of<br />

BMSCs. We implanted a human peripherally inserted central<br />

venous catheter (PICC) in the stomach and a subcutaneous<br />

infusion port in the back of the neck of each canine. Repeated<br />

injection of carbon tetrachloride (CCl 4<br />

) through the implanted<br />

catheter (high-dose period: 0.75 mL/kg body weight, twice a<br />

week for 6 weeks; low-dose period: 0.20 mL/kg body weight,<br />

twice a week for 4 weeks) was performed to induce liver cirrhosis.<br />

After 10 weeks of CCl 4<br />

injection, eight canines were<br />

equally divided into two groups: no cell infusion (control group)<br />

and autologous BMSC infusion via the peripheral vein (BMSC<br />

group). Low-dose CCl 4<br />

was continued after BMSC infusion, and<br />

blood examinations, ultrasonography-guided liver biopsies,<br />

and indocyanine green (ICG) tests were carried out before and<br />

at 4 weeks after the infusion. Results: Canine BMSCs cultured<br />

in standard DMEM with 10% FBS adhered to plastic and were<br />

CD44+, CD90+, and CD45−. They differentiated into adipocytes<br />

and osteocytes, consistent with known characteristics<br />

of BMSCs. In the BMSC group 4 weeks after BMSC infusion,<br />

the area of Sirius red-stained liver fibrosis was significantly<br />

reduced (fibrotic area (%); BMSCs: −1.62 ± 0.63; controls:<br />

+1.44 ± 1.01, p < 0.05), consistent with a significantly shortened<br />

half-life of ICG (half-life of ICG (min); BMSCs: −1.45 ±<br />

0.42; controls: +1.23 ± 0.87, p < 0.05). After BMSC infusion,<br />

no pro-coagulation activity and no thrombosis were seen in<br />

lung arteries using contrast-enhanced computed tomography.<br />

Conclusions: We established a useful canine liver cirrhosis<br />

model with repeated CCl 4<br />

administration through a PICC and<br />

confirmed that cultured autologous BMSC infusion improved<br />

liver cirrhosis and promoted liver regeneration without adverse<br />

effects.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takashi Matsuda, Taro Takami,<br />

Yuki Aibe, Toshihiko Matsumoto, Tsuyoshi Ishikawa, Naoki Yamamoto, Shuji<br />

Terai, Isao Sakaida<br />

1448<br />

A novel glycobiomarker, Wisteria floribunda agglutinin<br />

Macrophage Colony-Stimulating Factor Receptor can<br />

predict carcinogenesis and survival of liver cirrhosis<br />

with hepatitis C virus<br />

Etsuko Iio 1,3 , Makoto Ocho 2 , Akira Togayachi 2 , Masanori<br />

Nojima 4 , Atsushi Kuno 2 , Masashi Mizokami 5 , Hiroshi Yatsuhashi 6 ,<br />

Tatsuya Ide 7 , Shunsuke Nojiri 3 , Hisashi Narimatsu 2 , Yasuhito<br />

Tanaka 1 ; 1 Virology and Liver Unit, Nagoya City University Graduate<br />

School of Medical Sciences, Nagoya, Japan; 2 Research Center<br />

for Medical Glycoscience, National Institute of Advanced Industrial<br />

Science and Technology, Tsukuba, Japan; 3 Gastroenterology and<br />

Metabolism, Nagoya City University Graduate School of Medical<br />

Sciences, Nagoya, Japan; 4 Advanced Medicine Promotion,<br />

The Advanced Clinical Research Center, The Institute of Medical<br />

Science, University of Tokyo, Tokyo, Japan; 5 The Research Center<br />

of Japan, Hepatitis and Immunology, Kohnodai Hospital, International<br />

MedicalCenter, Ichikawa, Japan; 6 Clinical Research Center,<br />

National Nagasaki Medical Center, Omura, Japan; 7 Division of<br />

Gastroenterology, Department of Medicine, Kurume University,<br />

Kurume, Japan<br />

[Background/Aims] Recently, we identified a novel liver fibrosis<br />

glycobiomarker Wisteria floribunda agglutinin (WFA)-reactive<br />

colony stimulating factor 1 receptor (WFA + -CSF1R) using a glycoproteomics-based<br />

strategy. The aim of the present study was<br />

to assess the value of measuring WFA + -CSF1R levels for hepatocarcinogenesis<br />

and outcome in liver cirrhosis (LC) patients<br />

with hepatitis C virus (HCV). [Methods] WFA + -CSF1R and total<br />

CSF1R levels were measured by an antibody-lectin sandwich<br />

ELISA in serum samples from 214 consecutive HCV infected<br />

patients (99 CH and 115 LC) from January 1998 to April 2013<br />

in order to evaluate impact on carcinogenesis and survival of<br />

LC patients. Among 115 patients with LC, 56 (48.7%) patients<br />

without hepatocellular carcinoma (HCC). Ultimately, the 56 LC<br />

patients and independent 45 LC patients without HCC were<br />

assessed. [Results] LC patients without HCC of training cohort<br />

(n = 56) and validation cohort (n = 45) at baseline were analyzed<br />

for survival and carcinogenesis rates. The optimal cut-off<br />

of the WFA + -CSF1R were determined to be 310 ng/ml according<br />

to the smallest P value screened by Cox regression analysis<br />

in the training set. The AUC of WFA + -CSF1R for predicting<br />

overall survival, calculated by time-dependent ROC analysis,<br />

was 0.691 and the HR (per 1-SD increase) was 1.80 (95% CI,<br />

1.23-2.62, p < 0.001) of combined with the training set and<br />

validation set. The survival rate of LC patients with high WFA + -<br />

CSF1R levels (≥ 310 ng/ml) was significantly worse than those<br />

with lower levels (p < 0.01). Next, we analyzed the percentage<br />

of WFA + -CSF1R among total-CSF1R (WFA + -CSF1R%) for<br />

the potential of HCC. The optimal cut-off for the WFA + -CSF1R%<br />

for predicting the cumulative carcinogenesis rate was 35.0%.<br />

We further verified the utility of WFA + -CSF1R%, using the validation<br />

set. In the combined data the AUC of WFA + -CSF1R% for<br />

predicting the cumulative carcinogenesis rate was 0.760, with<br />

an HR of 1.66 (95% CI 1.26-2.20, p < 0.001). The AUC of<br />

WFA + -CSF1R% was superior to other fibrosis and tumor markers<br />

(i.e. Fib4, APRI, albumin, AFP, AFP-L3 and PIVKA-II) for predicting<br />

the cumulative carcinogenesis rate. LC patients in the<br />

training cohort with high WFA + -CSF1R% (≥ 35.0%, n = 6) had<br />

cumulative carcinogenesis rates of 33.3%, 50.0%, and 75.0%<br />

at 1, 3 and 5 years, respectively. In contrast, patients with<br />

low WFA + -CSF1R% (< 35.0%, n = 50) had cumulative rates<br />

of 4.6%, 18.0%, and 35.0%, respectively (p = 0.006). These<br />

data were supported in the validation cohort. [Conclusions]<br />

Assessing serum levels of WFA + -CSF1R has diagnostic value<br />

for predicting carcinogenesis and the survival of LC patients.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 917A<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Etsuko Iio, Makoto Ocho, Akira<br />

Togayachi, Masanori Nojima, Atsushi Kuno, Masashi Mizokami, Hiroshi Yatsuhashi,<br />

Tatsuya Ide, Shunsuke Nojiri, Hisashi Narimatsu<br />

1449<br />

Second harmonic generation microscopy based quantitative<br />

assessment for HBV-associated liver fibrosis<br />

Kai-Wen Huang; National Taiwan University Hospital, Taipei, Taiwan<br />

Background Liver fibrosis is associated with an abnormal<br />

increase in an extracellular matrix in chronic liver diseases.<br />

Quantitative characterization of collagen in intact tissue is<br />

essential for both fibrosis <strong>studies</strong> and clinical applications.<br />

Commonly used methods, histological staining followed by<br />

either semi-quantitative or computerized image analysis, have<br />

limited sensitivity, accuracy, and operator-dependent variations.<br />

The collagen in livers could be observed through second-harmonic<br />

generation (SHG) microscopy. The two photon<br />

excited fluorescence (TPEF) images, recorded simultaneously<br />

with SHG, clearly revealed the hepatocyte morphology and<br />

associated collagen progression dynamics. Purpose To access<br />

liver fibrosis using SHG/TPEF microscopy on human liver<br />

biopsy and resection samples, and to identify and quantitatively<br />

measure the morphological features, and to generate<br />

as an index for describing the severity of fibrosis. Methods A<br />

total of 95 specimens were sampled from patients with hepatitis<br />

B viral infections underwent liver biopsy or liver resection<br />

in National Taiwan University Hospital between 2010-2012.<br />

Liver fibrosis was graded by pathologist in NTUH using the<br />

Metavir scoring system. Quantitative analysis was also done<br />

on the same sample using SHG/TPEF microscopy to generate<br />

100 quantitative measurements on morphological features.<br />

These 100 quantitative measurements are used to develop the<br />

algorithm to fit the staging results from Metavir using non-linear<br />

support vector machine (SVM) model. Results Out of 100 quantitative<br />

measurements, 53 showed strong correlation with Metavir<br />

staging and could statistically distinguish patients of stage<br />

0 from stage 1, and 67 for distinguishing patients of stage 3<br />

from stage 4. Out of these quantitative measurements, 10 were<br />

selected to develop the algorithm using SVM model using half<br />

of the data for model training, and area under curve (AUC)<br />

generated from the remaining half samples showed 0.76 with<br />

these 10 selected features. Conclusion Quantitative assessment<br />

methodology combining non-stain SHG/TPEF imaging technology<br />

and non-linear fitting model showed great correlation<br />

with traditional pathology staging system. This process can be<br />

automated to minimize the potential inter and intra observer<br />

discrepancies and give highly quantitative and reproducible<br />

results for clinical diagnosis and treatment response assessment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kai-Wen Huang<br />

1450<br />

Hepatitis C treatment and Progression of Liver Stiffness<br />

in Australia, subsequent to the introduction of protease<br />

inhibitors<br />

Yang Wu 1 , Hank H. Chen 1 , Martin Weltman 1 , Guy D. Eslick 2 ,<br />

Bilel Jideh 1 , Izabella Pokorski 1 ; 1 Gastroenterology, Nepean Hospital,<br />

Kingswood, NSW, Australia; 2 Surgery, Nepean Hospital,<br />

Kingswood, NSW, Australia<br />

Background: Transient elastography (TE) is a noninvasive and<br />

well-validated method for measurement of liver stiffness in<br />

patients with chronic Hepatitis C. Aim: In the previous era of<br />

interferon-based therapy, <strong>studies</strong> have shown a strong association<br />

between Sustained Virological Response (SVR) and<br />

improvement in Liver stiffness (LS) 1 .Our current study aims<br />

to explore the kinetics of liver stiffness in Australian patients<br />

receiving hepatitis C treatment including protease inhibitors, an<br />

area not examined previously. Method: Consecutive patients at<br />

Nepean Hospital treated for Hepatitis C from 2011 onwards<br />

were included in the study. Patients were treated according to<br />

standard of care for their respective genotypes, which included<br />

protease inhibitors in patients with Genotype 1. The patient’s<br />

initial LS measurement was taken prior to their treatment and<br />

a second measurement was made at an interval at least 12<br />

weeks after the end of their treatment. Other patient factors<br />

such as gender, age, presence or absence of cirrhosis, alcoholism<br />

and blood tests were also collated. Results: Of the 36<br />

patients that were included in the study, 26 (72%) patients<br />

achieved SVR. Overall, 21(58%) patients were treated with<br />

protease inhibitors. The mean intra-patient change relative to<br />

baseline at the follow-up elastography was -4.97 kPa in the<br />

patient group who achieved SVR, vs +1.55 kPa in the patients<br />

who did not achieve an SVR (p = 0.07056). In uni-variate<br />

analysis (Mann-Whitney U test), other parameters such as protease<br />

inhibitor use (p=0.29), viral genotype, treatment experience<br />

(p=0.89), gender (p=0.09), age, alcohol intake (p=0.62)<br />

and presence of cirrhosis (p=0.94) were not predictive of LS<br />

improvement. Similarly, viral load and ALT measurement at<br />

the start of treatment didn’t reveal any significant relationship<br />

with post-treatment LS during Logistic regression. Conclusion:<br />

Measurement of liver stiffness with TE after hepatitis C treatment<br />

showed that although the SVR group had more improvement in<br />

the liver stiffness score compared to the non-SVR group, the difference<br />

was not statistically significant in this study. Our study<br />

is one of the first to include patients treated with protease inhibitors.<br />

References: Andersen ES, Moessner BK, Christensen PB,<br />

Kjær M, Krarup H, Lillevang S, Weis N. Lower liver stiffness<br />

in patients with sustained virological response 4 years after<br />

treatment for chronic hepatitis C. Eur J Gastroenterol Hepatol.<br />

2011 Jan;23(1):41-4.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yang Wu, Hank H. Chen, Martin<br />

Weltman, Guy D. Eslick, Bilel Jideh, Izabella Pokorski


918A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1451<br />

The Attention-to-Burden Index (ABI): a New Tool that<br />

Reveals Profound Disparities between Research Attention<br />

and Disease Burden among Liver Diseases.<br />

Nambi J. Ndugga 1 , Teisha G. Lightbourne 2 , Kavon Javaherian 2 ,<br />

Neha Verma 2 , Eric S. Orman 3 , David Pesci 2 , Ramon Bataller 4,2 ;<br />

1 Center for Gastrointestinal Biology and Disease, University of<br />

North Carolina at Chapel Hill School of Medicine, Chapel Hill,<br />

NC; 2 Nutrition, University of North Carolina at Chapel Hill, Chapel<br />

Hill, NC; 3 University of North Carolina at Chapel Hill, Chapel Hill,<br />

NC; 4 Medicine, University of North Carolina at Chapel Hill School<br />

of Medicine, Chapel Hill, NC<br />

Rationale: In recent years, there have been great advances<br />

in the management of HBV and HCV. Conversely, there are<br />

no approved therapies for fatty liver diseases (NAFLD and<br />

ALD). We hypothesize that these unequal advances are due<br />

to disparities in the research attention surrounding different<br />

liver diseases. Methods: To test this hypothesis, we developed<br />

the Attention-to-Burden Index (ABI), a comprehensive tool comprised<br />

of main research activities, and an estimate of disease<br />

burden for the main types of liver diseases: HCV, HBV, ALD,<br />

NAFLD. A systematic study of the research attention was carried<br />

out for the years 2010-2015 on these 4 liver diseases. The<br />

mean research attention for each liver disease was calculated<br />

from 5 different levels: reference in the two major annual liver<br />

meetings (AASLD and EASL), number of drugs in the pipeline of<br />

38 major pharmaceutical companies, funded grants (USA and<br />

EU), ongoing clinical trials (ClinicalTrials.gov) and scientific<br />

publications (PubMed). The mean disease burden was calculated<br />

from the following parameters: hospitalization costs in<br />

USA, etiology of liver transplantation in USA and Europe, etiology<br />

of cirrhosis, and etiology of liver fibrosis. The ABI for each<br />

liver disease was calculated as the mean attention divided by<br />

the mean disease burden. Results: The research attention in<br />

both AASLD and EASL meetings to the main liver diseases was<br />

similar: HCV (55 and 54% of total attention, respectively), HBV<br />

(26 and 26%), NAFLD (13 and 15%), and ALD (6 and 4%).<br />

The number of drugs in the pipeline with specific indications<br />

for liver diseases was: 13 for HCV, 4 for HBV, 3 for NAFLD,<br />

1 for ALD. The percentage of funded grants in the US and EU<br />

was: HCV (54 and 62% of total grants, respectively), HBV<br />

(24 and 33%), NAFLD (17 and 5%), and ALD (5 and 1%).<br />

A similar pattern of research attention was found in ongoing<br />

clinical trials and scientific publications. The calculated overall<br />

burden for the main liver disease was 34% for HCV, 4% for<br />

HBV, 37% for ALD and 24% for NAFLD. When comparing the<br />

ratio between the overall research attention and the disease<br />

burden of main liver diseases, we found striking differences.<br />

The calculated ABI for HCV and HBV revealed a 1 and 7-fold<br />

over-attention, respectively, while NAFLD and ALD received an<br />

infra-attention of 2 and 8-fold. Conclusions: The ABI reveals disparities<br />

between research attention and disease burden among<br />

liver diseases, with a clear over-attention to viral hepatitis compared<br />

to fatty liver diseases. Notably, the attention to ALD is<br />

minimal. There is a critical need to increase awareness of ALD<br />

in the liver research community.<br />

Disclosures:<br />

Ramon Bataller - Advisory Committees or Review Panels: Sandhill; Consulting:<br />

VTI, Oncozyme Pharma<br />

The following authors have nothing to disclose: Nambi J. Ndugga, Teisha G.<br />

Lightbourne, Kavon Javaherian, Neha Verma, Eric S. Orman, David Pesci<br />

1452<br />

Hospital Based Initiative to Decrease Readmission Rates<br />

in End-Stage Liver Disease<br />

Sheela S. Reddy 1 , Chad Gorn 2 , Stephania Dottin 2 , Ann E. Burke 1 ,<br />

Karen Newell 2 , Karen Pine 2 , Ursula Hobbs 2 , Maria Neff 2 , Jonathan<br />

M. Fenkel 1 , David A. Sass 1 , Steven K. Herrine 1 , Jesse M.<br />

Civan 1 , She-Yan Wong 1 , Dina Halegoua-De Marzio 1 ; 1 Medicine,<br />

Division of Gastroenterology and Hepatology, Thomas Jefferson<br />

University, Philadelphia, PA; 2 Transplant Services, Thomas Jefferson<br />

University Hospital, Philadelphia, PA<br />

Introduction: The Hospital Readmission Reduction Program, a<br />

part of the Affordable Care Act, requires reduced payments to<br />

hospitals with excess readmissions, or admission to a hospital<br />

within 30 days of discharge. This penalization is based on a<br />

hospital’s expected readmission rate, an adjusted value of the<br />

national average. Studies have shown a ~37% readmission<br />

rate for patient with end stage liver disease (ESLD). Due to the<br />

burden of ESLD readmissions at our institution, a care coordination<br />

(CC) program was created to improve transitions of<br />

care from the inpatient to outpatient setting. The aim of this<br />

study is to assess whether the implementation of the CC program<br />

reduced 30-day readmissions. Methods: Our institution<br />

is an urban quaternary care liver transplant center performing<br />

approximately 50 liver transplants yearly in highly competitive<br />

UNOS Region 2. Our CC program was created February<br />

2014, with a focus on patient education, short-term outpatient<br />

follow-up visits, and post-discharge follow-up calls by a transplant<br />

nurse coordinator. During inpatient admission, patients<br />

deemed high-risk by attending hepatologists were enrolled.<br />

High-risk was defined as cirrhotic patients admitted with fluid<br />

overload, hepatic encephalopathy and/or renal disease requiring<br />

hemodialysis. Once discharged, the patient was scheduled<br />

for an office visit within 7 days and received follow-up phone<br />

calls at 48 hours after discharge, and on days 7, 14, 21, and<br />

30. Patients who were transplanted, transferred, or expired<br />

before 30 days were excluded from final analysis. Results: Our<br />

hepatology service readmission rate was 33.67% in 2013.<br />

Since February 2014, 90 high-risk patients were identified and<br />

enrolled in the CC program. Of these 90 high-risk patients,<br />

30-day readmission was prevented in 37.9% of patients, but<br />

62.1% still had at least one 30-day readmission. CC protocol<br />

compliance was 87.5%. From October 2014 to March 2015,<br />

encephalopathy was the most likely reason for readmission<br />

(13 admissions), followed by bleeding (6), infection (6), and<br />

shortness of breath (6). There was no improvement seen in the<br />

hepatology readmission rate, with an overall rate of 36% in<br />

Q4 of 2014. Conclusion: Preventing readmission in patients<br />

with ESLD and high-risk decompensating events remains challenging,<br />

especially in a high MELD at transplant region. Liver<br />

transplantation is the only way to reduce readmissions in this<br />

population, as the natural history of ESLD lends itself to frequent<br />

hospitalization in the last stages of the disease. The target of a<br />

hospital-based initiative to prevent readmission should perhaps<br />

target those with a moderate risk for readmission in order to<br />

have an impact.<br />

Disclosures:<br />

Jonathan M. Fenkel - Advisory Committees or Review Panels: Merck; Consulting:<br />

Gilead Pharmaceuticals, Janssen Therapeutics, Bristol-Myers Squibb, Abbvie<br />

Steven K. Herrine - Board Membership: ABIM; Grant/Research Support: BMS,<br />

Gilead, Merck, Vertex<br />

Jesse M. Civan - Consulting: Merck<br />

The following authors have nothing to disclose: Sheela S. Reddy, Chad Gorn,<br />

Stephania Dottin, Ann E. Burke, Karen Newell, Karen Pine, Ursula Hobbs, Maria<br />

Neff, David A. Sass, She-Yan Wong, Dina Halegoua-De Marzio


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 919A<br />

1453<br />

Cost-effectiveness of Sofosbuvir plus Ledipasvir Antiviral<br />

Treatment for Elderly Patients with Chronic Hepatitis C<br />

Genotype 1<br />

Antonio Ciaccio 2 , Paolo Cortesi 3 , Giuseppe Bellelli 4 , Matteo<br />

Rota 5 , monica Rota 2 , Lorenzo G. Mantovani 3 , Giorgio Annoni 4 ,<br />

Mario Strazzabosco 1,2 ; 1 Yale University, Section of Digestive Diseases,<br />

New Haven, CT; 2 Department of Surgery and Translational<br />

Medicine, University of Milano-Bicocca, Milan, Italy; 3 Research<br />

Centre on Public Health (CESP), University of Milan-Bicocca,<br />

Milan, Italy; 4 Department of Health Science, University of Milan-Bicocca,<br />

Milan, Italy; 5 Department of Epidemiology, IRCCS-Istituto di<br />

Ricerche Farmacologiche Mario Negri, Milan, Italy<br />

A relevant proportion of patients affected by Chronic Hepatitis<br />

C (CHC) is older than 65 years. In the interferon era, comorbidities<br />

and a higher susceptibility to interferon/ribavirin adverse<br />

events have limited treatment in these patients. Recent approval<br />

of interferon-free regimens, with high efficacy and limited toxicity,<br />

provide unprecedented chances for these patients to be<br />

treated. However, the cost-effectiveness of all-oral Direct-Acting<br />

Antivirals (DAAs) has not been addressed in the elderly<br />

population. We have performed a cost-effectiveness analysis<br />

taking into account the severity of liver disease, the age of the<br />

patient and the geriatric (frailty) status. METHODS: A semi-Markov<br />

model of CHC natural history was built. The study focuses<br />

on CHC Genotype 1 patients older than 65 years, stratified<br />

according to liver fibrosis (METAVIR F3 and F4), age (65 to 90<br />

years old) and frailty phenotype defined by Fried’s (not frail,<br />

pre-frail and frail) for a total of 30 cohorts simulated. Treatment<br />

with sofosbuvir plus ledipasvir (SOF/LDV) versus no treatment<br />

was assessed for each cohort. The model estimated costs, Life<br />

Years and Quality-Adjusted Life Years (QALY) using a lifetime<br />

time horizon and the Health System perspective. Results are<br />

presented as incremental cost-effectiveness ratios (ICERs) per<br />

QALY gained. Cost-effectiveness was defined as an ICER under<br />

the willingness-to-pay threshold of 40,000 € per QALY gained.<br />

RESULTS: At each fibrosis score, ICER increased with age and<br />

frailty index. Among patients with F3 fibrosis, ICER ranged from<br />

€5,530/QALY in not-frail 65 years old and €149,355 in frail<br />

85 years old patients. Among F4 patients ICER ranged from<br />

€6,552/QALY in not frail 65 years old and €85,559 in frail<br />

85 years old patients. In F3 and F4 cohorts ICER was below<br />

€40,000/QALY up to age 84 and 87 years in non-frail pts, up<br />

to age 80 and 83 years in pre-frail patients, and up to age 76<br />

and 80 in frail patients, respectively. Adopting an alternative<br />

scenario with a 20% discount of SOF/LDV treatment, cost-effectiveness<br />

increases accordingly. CONCLUSION: This analysis<br />

shows that, assuming that efficacy and safety of SOF/LDV are<br />

not age-dependent, SOF/LDV treatment is cost-effective in most<br />

CHC patients older than 65 years, although a careful assessment<br />

of the patient status is mandatory. Furthermore, under<br />

conditions of restricted budget, it should be considered that<br />

ICER varies considerably between two non-frail F4 patients,<br />

one at age 65 and the other at 85 years (between 8,000 and<br />

33,000 EUR/QALY, respectively). This cost-effectiveness analysis<br />

should promote a prospective study to verify efficacy and<br />

side effects in elderly HCV patients.<br />

Disclosures:<br />

Antonio Ciaccio - Grant/Research Support: Gilead Sciences<br />

Lorenzo G. Mantovani - Advisory Committees or Review Panels: Bayer; Grant/<br />

Research Support: Jansen, Merck and Co; Speaking and Teaching: Bayer<br />

The following authors have nothing to disclose: Paolo Cortesi, Giuseppe Bellelli,<br />

Matteo Rota, monica Rota, Giorgio Annoni, Mario Strazzabosco<br />

1454<br />

Living Donor Liver Transplants (LDLT), compared to<br />

Deceased Donor Liver Transplants (DDLT), and MELD<br />

Score Increase Liver Transplant Resource Utilization at<br />

an Experienced Transplant Center from June 2008 to<br />

July 2013<br />

Grace Kim 1 , Yael J. Coppleson 2 , Robert S. Brown 3,2 , Thresiamma<br />

Lukose 1 , Jean C. Emond 1,2 , Benjamin Samstein 2 ; 1 Surgery, Columbia<br />

University, New York, NY; 2 NewYork Presbyterian, New York,<br />

NY; 3 Medicine, Columbia University, New York, NY<br />

Purpose: To examine liver transplant charges at an experienced<br />

transplant center. Methods: This retrospective analysis<br />

used data for liver transplant recipients at NewYork-Presbyterian<br />

from hospital billing charges, medical records, and<br />

UNOS. The patient population (n=526) included both adult<br />

(n=458) and pediatric patients (n=68) undergoing a first liver<br />

transplant, but excluded combined transplants and status 1<br />

listed patients. Total hospital charges were used as a proxy<br />

for resource utilization, and included all clinical services and<br />

resources for transplant day, and inpatient and outpatient preand<br />

post-transplant care. Transplant day charges included both<br />

donor and recipient operating and recovery services. Multivariate<br />

linear regressions were utilized to examine the impact of<br />

living donor transplantation and MELD on transplant day and<br />

total transplant charges, controlling for gender and age, and<br />

diagnosis of HCC. Results: The average overall MELD score<br />

of the cohort was 24.3, while the average MELD was 25.4<br />

for DDLT (n=474), and 18.1 for LDLT (n=87) (p=0.00). Age,<br />

gender, and etiology of liver disease were similar across donor<br />

type. For adult patients, transplant day charges were greater<br />

for LDLT patients (p=.04) and higher MELD patients (p=0.00).<br />

Total charges were greater but not statistically significant<br />

(p=.14) for LDLT patients, though still increased by each MELD<br />

point (p=0.00). For all patients, transplant day charges were<br />

higher for LDLT patients (p=.04) and higher MELD patients,<br />

($13K per point, p =0.00). For total costs, LDLT charges were<br />

higher (p=.16) and MELD point increased charges (p=.01).<br />

The difference between LDLT and DDLT in total transplant<br />

charges were equalized by an 8-point differential in MELD<br />

score. Conclusion: Living donor transplants are associated with<br />

greater transplant resource utilization, but the difference in total<br />

transplant charges is not statistically significant. The higher<br />

charges associated with the increased complexity of LDLT cases<br />

are mitigated by shorter waiting time and fewer post-operative<br />

days until discharge. Patients listed with higher MELD scores<br />

utilize significantly more resources for transplant care.<br />

Disclosures:<br />

Robert S. Brown - Consulting: Gilead, Janssen, Abbvie, Merck, BMS<br />

The following authors have nothing to disclose: Grace Kim, Yael J. Coppleson,<br />

Thresiamma Lukose, Jean C. Emond, Benjamin Samstein


920A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1455<br />

A probabilistic decision model using non-invasive fibrosis<br />

markers in Primary Care NAFLD pathways predicts<br />

increased cirrhosis detection rates and reduced overall<br />

healthcare expenditure<br />

Ankur Srivastava 1,2 , Simcha Jong 3 , Anna Gola 4 , Laura Fenlon 5 ,<br />

Petra Scantlebury 5 , Sudeep Tanwar 1,2 , Hannah Liu 7 , Ruth E.<br />

Gailer 6,7 , Sarah Morgan 6 , Alex Warner 6 , Karen Sennett 7 , Julie<br />

Parkes 1 , James O’Beirne 1,2 , Emmanuel Tsochatzis 1,2 , William M.<br />

Rosenberg 1,2 ; 1 Institute of Liver and Digestive Health, University<br />

College London (UCL), London, United Kingdom; 2 Hepatology,<br />

Royal Free London NHS Foundation Trust, London, United Kingdom;<br />

3 Management Science & Innovation, University College<br />

London, London, United Kingdom; 4 Health Economics, University<br />

College London, London, United Kingdom; 5 Department of<br />

Finance, Royal Free London NHS Foundation Trust, London, United<br />

Kingdom; 6 Primary Care, Camden Clinical Commissioning Group,<br />

London, United Kingdom; 7 Primary Care, Islington Clinical Commissioning<br />

Group, London, United Kingdom<br />

Background: Risk factors for Non-Alcoholic Fatty Liver Disease<br />

(NAFLD) have reached epidemic proportions and whilst only<br />

a minority of at-risk individuals develop significant liver disease<br />

warranting specialist referral, ensuring identification of<br />

this group is a primary care challenge. Non-invasive tests (NIT)<br />

for liver fibrosis enable community-based doctors to identify<br />

patients with NAFLD who have advanced fibrosis and would<br />

benefit from early referral for specialist care and interventions<br />

to limit the complications of cirrhosis. As part of an evaluation<br />

of a novel primary care NAFLD pathway in two London boroughs<br />

(Camden and Islington, C&I), which uses NIT to stratify<br />

patients for advanced liver fibrosis, we have constructed an<br />

analytical model to examine the potential cost implications of<br />

this strategy on an intention-to-treat basis. Methods: A probabilistic<br />

decision analytical model was constructed to explore<br />

the financial dimensions of two strategies in community based<br />

management of NAFLD from a healthcare payer perspective:<br />

1) Standard of care (SOC) comprising physicians’ clinical judgment<br />

based on clinical history, examination, blood tests and<br />

ultrasound 2) SOC plus Non-Invasive Testing for liver fibrosis<br />

using FIB-4, and the ELF test (to resolve indeterminate FIB-4<br />

tests) to guide referral. The model was populated from the literature,<br />

national UK data and expert opinion. A five year time<br />

horizon was applied. Results: The C&I population is 434,958.<br />

It was assumed that 30% of the NAFLD population consult their<br />

general practitioner (GP) annually. >F2 fibrosis prevalence in<br />

NAFLD was set at 5%. Local clinical audit estimated SOC sensitivities<br />

and specificities of 0.35 and 0.70. The model estimated<br />

24% needed ELF test. Utilizing NIT reduced referrals of<br />

≤F2 disease by over 80% and resulted in over 50% increase<br />

in the detection of cirrhosis. There was a significant reduction<br />

in liver transplantation rates by approximately 15%. The<br />

overall expenditure was reduced by a fifth, primarily through<br />

reductions in referrals for low risk cases, costs related to HCC<br />

management and emergency inpatient admissions. Discussion:<br />

The use of NIT led to a reduction in inappropriate referrals and<br />

improvement in the early detection of cirrhosis and its complications.<br />

These changes are likely to result in significant health<br />

gains and economic cost savings. With NAFLD prevalence<br />

likely to increase, policy needs to address these challenges.<br />

Modeling suggests that the Camden and Islington NAFLD pathway<br />

employing FIB-4 and ELF to guide GPs’ risk stratification<br />

of patients will achieve these goals. A full clinical effectiveness<br />

and cost consequence analysis is underway.<br />

Disclosures:<br />

Simcha Jong - Grant/Research Support: Siemens<br />

Julie Parkes - Stock Shareholder: iQur (spouse is shareholder)<br />

William M. Rosenberg - Advisory Committees or Review Panels: Janssen, Merk,<br />

Gilead, Merk, Gilead, GSK; Board Membership: iQur Limited, iQur Limited;<br />

Consulting: siemens; Speaking and Teaching: siemens, Roche<br />

The following authors have nothing to disclose: Ankur Srivastava, Anna Gola,<br />

Laura Fenlon, Petra Scantlebury, Sudeep Tanwar, Hannah Liu, Ruth E. Gailer,<br />

Sarah Morgan, Alex Warner, Karen Sennett, James O’Beirne, Emmanuel Tsochatzis<br />

1456<br />

Impact of Daclatasvir-Sofosbuvir Combination Treatment<br />

on Medical Events and Costs in Patients Infected with<br />

Genotype 3 Hepatitis C Virus<br />

Boris Gorsh 1 , Bruce E. Sill 1 , Shalini Hede 1 , Catherine St-Laurent<br />

Thibault 2 , Divya Moorjaney 3 , Michael Ganz 3 , Anupama<br />

Kalsekar 1 , Yong Yuan 1 ; 1 Bristol-Myers Squibb, Plainsboro, NJ;<br />

2 Evidera, Montreal, QC, Canada; 3 Evidera, Boston, MA<br />

Purpose: Patients infected with genotype (GT) 3 of the hepatitis<br />

C virus (HCV) fail antiviral therapy more often than patients<br />

infected with other GTs. Combination daclatasvir-sofosbuvir<br />

(DCV+SOF) is a significant advancement as it requires 12<br />

rather than 24 weeks for other combination therapies. Our<br />

objective is to estimate the number of medical events and<br />

direct medical costs avoided if, otherwise untreated, GT3 HCV<br />

patients were treated with DCV+SOF. Methods: We used a<br />

published Markov model [i] to estimate the number of medical<br />

events and costs experienced by a hypothetical cohort of<br />

1,000 HCV GT3 patients (mean age of 54 years, 65% male)<br />

who received DCV+SOF or no treatment. The patient distribution<br />

we used was based on McGarry et al (2012). [ii] We<br />

assumed that 89% of the DCV+SOF-treated patients would<br />

achieve a sustained viralogic response based on results of the<br />

ALLY-3 study. Medical events and costs were assessed for three<br />

scenarios: overall HCV population (F0-F4) receives no treatment,<br />

only the F3/F4 subpopulation receives DCV+SOF, and<br />

all patients (F0-F4) receive DCV+SOF. Total medical events<br />

were defined as the sum of compensated cirrhosis, decompensated<br />

cirrhosis, hepatocellular carcinoma, liver transplant, or<br />

liver-related death. Medical costs include monitoring costs and<br />

costs for treating complications and adverse events. Disease<br />

progression and costs were modelled over time horizons of 20<br />

and 50 years. Results: Medical events and direct medical costs<br />

were lowest when all patients were treated with DCV+SOF (76<br />

events; $14,343,786). When only F3/F4 patients received<br />

DCV+SOF, medical events and costs were substantially higher<br />

(335 events; $68,414,213). In the scenario where no one<br />

was treated, events and costs were the highest (729 events,<br />

$129,838,658). Similar results were seen over 50 years.<br />

Conclusion: Substantially fewer medical events and lower<br />

medical costs are accrued when all GT3 patients are treated<br />

with DCV+SOF compared with scenarios in which no one is<br />

treated or treatment is limited to F3/F4 patients. [i] McEwan P,<br />

Ward T, Yuan Y, et al. The impact of timing and prioritization<br />

on the cost-effectiveness of birth cohort testing and treatment<br />

for hepatitis C virus in the United States. Hepatology. 2013<br />

Jul;58(1):54-64. [ii] McGarry LJ, et al. Economic model of a<br />

birth cohort screening program for hepatitis C virus. Hepatology.<br />

2012 May;55(5):1344-55.<br />

Medical Events and Costs (Per Cohort)<br />

Disclosures:<br />

Boris Gorsh - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 921A<br />

Bruce E. Sill - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb<br />

Anupama Kalsekar - Employment: Bristol Myers Squibb<br />

Yong Yuan - Employment: Bristol Myers Squibb Company<br />

The following authors have nothing to disclose: Shalini Hede, Catherine St-Laurent<br />

Thibault, Divya Moorjaney, Michael Ganz<br />

1457<br />

Hospitalisation for primary biliary cirrhosis: a UK-PBC<br />

analysis of hospital episodes<br />

Gideon Hirschfield 2 , Luke Vale 3 , Tracy J. Mayne 1 , George F.<br />

Mells 4 , David Shapiro 1 , David Jones 5 ; 1 Intercept Pharmaceuticals,<br />

San Diego, CA; 2 Centre for Liver Research, University of Birmingham,<br />

Birmingham, United Kingdom; 3 Institute of Health & Society,<br />

Newcastle University, Newcastle, United Kingdom; 4 Division of<br />

Gastroenterology and Hepatology, Department of Medicine, University<br />

of Cambridge, Cambridge, United Kingdom; 5 Institute of<br />

Cellular Medicine, Newcastle University Medical School, Newcastle,<br />

United Kingdom<br />

Background: The advent of new therapies for patients with<br />

primary biliary cirrhosis (PBC) highlights the need to understand<br />

the current health burden of PBC. Objective: To assess<br />

the frequency and nature of hospitalisations associated with<br />

primary biliary cirrhosis (PBC). Methods: UK-PBC analysed all<br />

records from 2009 through 2014 in the Hospital Episodes Statistics<br />

database, containing information on all hospitalisation<br />

across the National Health Service in England, where the ICD-<br />

10 code for PBC (K74.3) appeared. We characterised primary<br />

diagnosis for each hospitalisation. Results: There were 21,275<br />

admissions with a PBC code over the 5-year observational<br />

period covering 1631 unique ICD-10s listed as the primary<br />

hospitalisation code. PBC was the primary code for 17% of<br />

admissions. The number of relevant hospitalisations increased<br />

from 3368 in 2009/2010 to 4995 in 2012/2014. The number<br />

per 100,000 hospitalisations increased from 23 to 32 over<br />

this period (Figure). The increase was almost completely driven<br />

by the inclusion of PBC as a secondary diagnosis. Liver transplants<br />

represented 5-10% of admissions with PBC as a primary<br />

code, and 1- 2% of any PBC-related admission. The next 20<br />

ICD-10 codes accounted for an additional 17% of admissions.<br />

The top 20 codes included known sequelae of PBC: ascites<br />

(2%), varices (2%), and liver cancer (1%). Remaining codes<br />

included: anaemia (iron deficiency and unspecified) 3%; respiratory<br />

(pneumonia, COPD and lower respiratory infections)<br />

3%; osteoporosis and fractures 2%; and various abdominal<br />

(unspecified abdominal pain, gastritis and gastroenteritis) 2%.<br />

Conclusion: Our UK-PBC data confirms a significant clinical<br />

need for patients with PBC, including a rising healthcare burden<br />

from disease. The increase in PBC-related diagnoses may<br />

be due to more accurate coding, greater co-morbidity as the<br />

result of increased longevity, or may represent the true impact<br />

of disease. New therapies preventing the progression of PBC to<br />

end-stage cirrhosis, and its complications, may help to reduce<br />

this rising burden.<br />

Disclosures:<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

David Jones - Consulting: Intercept, Pfizer, Novartis; Speaking and Teaching:<br />

Falk, Shire<br />

The following authors have nothing to disclose: Gideon Hirschfield, Luke Vale,<br />

George F. Mells<br />

1458<br />

Obeticholic Acid (OCA) has Incremental Efficacy in<br />

Patients with Higher Baseline Alkaline Phosphatase<br />

(ALP) Levels<br />

Tracy J. Mayne, Richard Pencek, Tonya Marmon, David Shapiro;<br />

Intercept Pharmaceuticals, Inc., San Diego, CA<br />

Objective: Examine ALP lowering stratified by baseline ALP<br />

among patients with primary biliary cirrhosis (PBC) treated with<br />

10 mg OCA. Methods: We pooled data from the 10 mg treatment<br />

arms of two phase 2 trials (n=20 and n=38) and one<br />

phase 3 trial (n=73) randomized, double-blind, placebo-controlled<br />

trials of OCA in patients with PBC. The percent change<br />

in ALP at 3 months was calculated based on baseline ALP stratified<br />

by 50 U/L. We ran a linear regression predicting percent<br />

change in ALP based on baseline ALP. Results: As shown in the<br />

first figure, both absolute and percent reduction in ALP were<br />

greater at higher baseline ALP levels. As shown in the second<br />

figure, for each incremental 100 U/L in baseline ALP, there<br />

was an incremental 5% reduction in ALP at 3 months (intercept<br />

= 17%) in patients treated with 10 mg OCA. Conclusion: A<br />

standard measure of drug efficacy is percent reduction in a<br />

specified marker. Statin efficacy is measured as percent reduction<br />

in low density lipid cholesterol (LDL-C); blood pressure<br />

medications are assessed by percent reduction in systolic and<br />

diastolic blood pressure. At a fixed percent reduction, absolute<br />

reduction in a biomarker will be greater with higher baseline<br />

levels: a 50% reduction in LDL-C at a baseline of 200 mg/dL<br />

is 100 mg/dL; a 50% reduction at 100 mg/dL is 50 mg/dL.<br />

In the case of OCA, the percent reduction in ALP increases as<br />

a function of baseline ALP. As shown by the regression model,<br />

OCA produces an average 59% reduction at a baseline ALP<br />

of 800 U/L, and an average 38% reduction at a baseline ALP<br />

of 400 U/L. Achieved level of ALP has been shown to be associated<br />

with improved progression-free survival (Lammers et al,<br />

Gastroenterology 2014;147:1338). These data indicate that<br />

patients with more progressed PBC, as measured by higher<br />

serum ALP, may receive incremental benefit when treated with<br />

OCA.


922A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Tracy J. Mayne - Employment: Intercept Pharmaceuticals<br />

Richard Pencek - Employment: Intercept Pharmaceuticals; Stock Shareholder:<br />

Intercept Pharmaceuticals<br />

Tonya Marmon - Employment: Intercept Pharmaceuticals, Inc; Stock Shareholder:<br />

Intercept Pharmaceuticals, Inc<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

1459<br />

Impact of Comorbidity on Inpatient Outcomes for<br />

Decompensated Cirrhosis<br />

Brandon Nokes 2,1 , Amit Kashyap 3 , Thomas D. Boyer 4,3 , Archita P.<br />

Desai 4,3 ; 1 Internal Medicine, Mayo Clinic Arizona, Phoenix, AZ;<br />

2 College of Medicine, University of Arizona, Tucson, AZ; 3 Gastroenterology<br />

and Hepatology, University of Arizona, Tucson, AZ;<br />

4 Liver Research Institute, University of Arizona, Tucson, AZ<br />

Background: Decompensated cirrhosis is an increasingly<br />

common reason for admission. Inpatients with cirrhosis and<br />

comorbidities have worse outcomes. While current comorbidity<br />

indices used for risk adjustment report number of comorbidities,<br />

the complex interaction between cirrhosis, comorbidity and outcomes<br />

may be better understood by capturing naturally occurring<br />

subgroups of cirrhosis patients with common comorbidity<br />

profiles. Aim: We sought to describe inpatient health care utilization<br />

due to decompensated cirrhosis and to create comorbidity<br />

profiles for patients admitted with decompensated cirrhosis.<br />

Methods: We examined the 2008-2011 NIS and included all<br />

admissions with a diagnosis of cirrhosis and a complication of<br />

cirrhosis. Means and proportions with standard errors and confidence<br />

intervals were used to describe the sample. Latent class<br />

mixture modeling was used to identify distinct and common<br />

comorbidity profiles among binary categorical variables indicating<br />

the presence or absence of comorbidities. Generalized<br />

linear model analysis was used to quantify associations among<br />

comorbidity profiles and outcomes. Results: Of the full NIS,<br />

414521 (1.3%) admissions had a cirrhosis diagnosis of which<br />

237553 (57%) met inclusion criteria. The average patient was<br />

58 years old, white (58%), male (62%) with hepatitis C (29%)<br />

or alcohol related liver disease (51%), insured by Medicare<br />

(40%) and cared for in a large (66%), non-teaching (51%),<br />

rural (90%) hospital. Complications of cirrhosis were most often<br />

ascites (57%), portal hypertension (40%), hepatic encephalopathy<br />

(28%), and acute renal failure (23%). Compared to<br />

those with Elixhauser comorbidity index of 0, those with an<br />

index of 3 or more had 72% higher inpatient mortality (5.1%<br />

vs. 8.7%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 923A<br />

Figure. Changes in fatigue and work productivity scores during<br />

and after treatment with different anti-HCV regimens.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Maria Stepanova, Linda Henry,<br />

Fatema Nader, Sharon L. Hunt<br />

1461<br />

Cost-effectiveness of Daclatasvir in Combination with<br />

Sofosbuvir for the Treatment of Subjects with Genotype<br />

3 Chronic Hepatitis C Infection in the United States<br />

Catherine St-Laurent Thibault 1 , Divya Moorjaney 1 , Michael Ganz 1 ,<br />

Bruce E. Sill 2 , Shalini Hede 2 , Yong Yuan 2 , Anupama Kalsekar 2 ,<br />

Boris Gorsh 2 ; 1 Modeling and Simulation, Evidera, St-Laurent, QC,<br />

Canada; 2 Bristol-Myers Squibb Company, Plainsboro, NJ<br />

BACKGROUND: An open label phase III trial evaluated the<br />

efficacy and safety of daclatasvir (DCV) in combination with<br />

sofosbuvir (SOF) for treatment of subjects with genotype (GT)<br />

3 hepatitis C virus (HCV). This study evaluated the cost-effectiveness<br />

of DCV+SOF versus SOF in combination with ribavirin<br />

(RBV) over a time horizon of 20 years from the payer perspective<br />

in the United States (US). METHODS: A published Markov<br />

model was adapted to reflect US baseline demographic<br />

characteristics, treatment patterns, costs of drug acquisition,<br />

monitoring, disease management and adverse event management,<br />

and mortality risks. Clinical inputs were based on the<br />

ALLY3 trial for DCV+SOF and the VALENCE trial for SOF+RBV.<br />

The primary cost-effectiveness outcome was the incremental<br />

cost-utility ratio (ICUR). Life-years, incidence of complications<br />

(compensated cirrhosis, decompensated cirrhosis, hepatocellular<br />

carcinoma, liver transplant and liver-related death),<br />

number of patients achieving sustained virological response<br />

(SVR) and the total cost per SVR were used as the secondary<br />

outcomes. Costs (2014 USD) and QALYs were discounted at<br />

3% per year. An estimated cost was used for DCV based on<br />

the current standard of care. Deterministic sensitivity analyses<br />

(DSA) (varying SVR probabilities [+/-10%], health state and<br />

therapy costs [+/-20%]), probabilistic sensitivity analyses (PSA)<br />

(with probabilistic sampling of key inputs at each of the 1,000<br />

iterations using their respective standard error) and scenario<br />

analyses (evaluating the impact of various time horizons [5, 10<br />

and 80 years] along with analyses for specific subpopulations<br />

[treatment-naïve, treatment-experienced, cirrhotic or non-cirrhotic])<br />

were conducted to assess the robustness of the results<br />

to changes in inputs and assumptions. RESULTS: Compared<br />

with SOF+RBV, DCV+SOF was a dominant treatment in the<br />

base case as well as in almost all scenarios (i.e., treatment-experienced.<br />

non-cirrhotic, lower DCV price, 5 and 10 year time<br />

horizon). In the cirrhotic and treatment naïve population scenarios,<br />

DCV+SOF was less costly but also slightly less effective<br />

compared to SOF+R. DSA for the overall population showed<br />

that ICURs were sensitive to variations in the probability of<br />

achieving SVR for DCV+SOF and SOF+RBV. PSA for the overall<br />

population indicated that DCV+SOF yielded an ICUR below<br />

$50,000 per QALY gained in 78.7% of all iterations compared<br />

to SOF+RBV. CONCLUSION: DCV+SOF is a dominant<br />

option compared with SOF+RBV in the US for the overall GT3<br />

HCV patient population. It may be important to understand the<br />

real-world SVR rates, as variations in this efficacy parameter<br />

can have a significant impact on the ICUR.<br />

Disclosures:<br />

Bruce E. Sill - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb<br />

Yong Yuan - Employment: Bristol Myers Squibb Company<br />

Anupama Kalsekar - Employment: Bristol Myers Squibb<br />

Boris Gorsh - Employment: Bristol-Myers Squibb; Stock Shareholder: Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Catherine St-Laurent Thibault,<br />

Divya Moorjaney, Michael Ganz, Shalini Hede<br />

1462<br />

The cost-effectiveness of ombitasvir/paritaprevir/ritonavir,<br />

dasabuvir and ribavirin (3D+R) in patients with<br />

liver transplantation after genotype 1 hepatitis C virus<br />

infection<br />

Sammy Saab 2 , Yuri Sanchez 3 , Caroline Huber 1 , Alice Wang 3 ,<br />

Timothy R. Juday 3 ; 1 Precision Health Economics, Los Angeles, CA;<br />

2 Pfleger Liver Institute, University of California, Los Angeles, Los<br />

Angeles, CA; 3 AbbVie, Inc., North Chicago, IL<br />

Background: Hepatitis C virus (HCV) is the most common indication<br />

for liver transplantation, and HCV infection recurs in<br />

over 90% of patients within a month of transplant. Liver fibrosis<br />

progresses more rapidly after recurrence, and treatment<br />

options historically have been limited in this population. New<br />

interferon-free regimens have demonstrated safety and efficacy<br />

in the post-transplant HCV population; we explore the cost-effectiveness<br />

of regimens with FDA-approval or Phase III trial<br />

results in the post-liver transplant setting from a US healthcare<br />

system perspective. Methods: A cohort of post-liver transplant<br />

patients with recurrent HCV genotype 1 and fibrosis stage<br />

F0-F2 was modeled using a two-phase Markov model. In the<br />

first phase, all patients underwent one of three treatment scenarios:<br />

no treatment (NT), pegylated interferon and ribavirin<br />

for 48 weeks (PR48), or ombitasvir/paritaprevir/ritonavir,


924A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

dasabuvir and ribavirin for 24 weeks (3D+R). This phase distributed<br />

patients into post-treatment disease states, including<br />

sustained virologic response (SVR), based on treatment efficacy.<br />

The second phase followed the cohort’s post-treatment<br />

progression through further liver fibrosis and death. Outcome<br />

measures included lifetime costs (medical and pharmaceutical),<br />

quality-adjusted life years (QALYs) and incremental cost-effectiveness<br />

ratios (ICERs) associated with each treatment. SVR<br />

and adverse event rates were based on phase III clinical trials.<br />

Transition probabilities, utilities and medical costs were<br />

extracted from published literature; treatment costs assumed<br />

100% adherence and reflected wholesale acquisition costs<br />

from December 2014 Red Book. Results: In a post liver-transplant<br />

setting, the use of 3D+R had the lowest discounted overall<br />

costs ($423,585) and highest gain in QALYs (11.3) compared<br />

to NT ($724,757 cost, 8.25 QALYs) or PR48 ($658,411 cost,<br />

8.73 QALYs). Use of 3D+R also generated net cost-savings<br />

for the healthcare system given the lower cost of this regimen<br />

($169,018) compared to the incremental discounted costs of<br />

NT ($301,172) and PR48 ($234,826). Conclusions: Compared<br />

to other regimens for treatment of post-liver transplant<br />

genotype 1 HCV recurrence, 3D+R is a cost-effective option<br />

that generates significantly higher quality-adjusted survival and<br />

lower costs. Use of 3D+R in the post-liver transplant population<br />

could also generate net cost savings for the healthcare system<br />

compared to previous standards of care.<br />

1463<br />

Gut Microbiota Alterations are an Independent Prognostic<br />

Factor for 90-day hospitalization in Cirrhosis<br />

Jasmohan S. Bajaj 1 , Naga Betrapally 2 , Douglas M. Heuman 1 , Melanie<br />

White 1 , Ariel Unser 1 , Edith A. Gavis 1 , Phillip B. Hylemon 1 ,<br />

Leroy Thacker 1 , Masoumeh Sikaroodi 2 , Patrick M. Gillevet 2 ; 1 VCU<br />

and McGuire VAMC, Richmond, VA; 2 Microbiome Analysis Center,<br />

Geoge Mason University, Manassas, VA<br />

Gut microbiota abnormalities (dysbiosis) worsens with the progression<br />

of cirrhosis in cross-sectional <strong>studies</strong>. However the<br />

impact of fecal dysbiosis on hospitalizations in cirrhotics is<br />

unclear. Aim: Define the association of dysbiosis with 90-day<br />

hospitalizations in cirrhosis. Methods: 284 cirrhotics (56 yrs,<br />

MELD 11, 35% alcohol, 38% prior HE, 30% Diabetes) underwent<br />

stool multi-tagged sequencing analysis to determine the<br />

relative microbial family abundances including beneficial commensals<br />

such as Lachnospiraceae, Ruminococceae & Clostridiales<br />

XIV. Subjects were followed for 90 days and non-elective<br />

hospitalizations were noted. Binary logistic regression using<br />

90-day hospitalizations as the dependent variable with MELD,<br />

PPI use, HE, alcohol etiology, diabetes, and bacteria significant<br />

on univariate analyses as predictors was performed.<br />

Results: Of the 284, 25 were lost to follow-up and 5 had elective<br />

hospitalizations. 94 (37%) were non-electively hospitalized<br />

within 90 days (median 35, IQR 21-78 days). The major<br />

(n=87) hospitalization reason were liver-related, thus no further<br />

sub-analysis was done(HE=46, Infection=14, Anasarca=13,<br />

GI bleed=10, others=4). Those who were hospitalized had<br />

a worse cirrhosis severity, were younger, and higher PPI use<br />

(Table). Alcoholic etiology was similar (25% vs. 18%, p=0.4).<br />

Stool microbiota showed a lower relative abundance of Bacteroidaceaeae<br />

(11% vs 26%, p=0.002), Lachnospiraceae (8%<br />

vs 14%,p=0.001), Ruminococceae (3% vs. 7%,p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 925A<br />

years; HBV/HCV/alcohol/NASH/others = 1/16/9/4/6).<br />

Data were statistically compared and examined both before<br />

and 1 month after B-RTO. Results: Preoperative Child-Pugh (CP)<br />

score was higher in the HE group than in the GV group (8.4<br />

vs. 6.2 pts., p < 0.01), and both dilatation of splenic vein<br />

(12.7 vs. 9.0 mm, p < 0.01) and constriction of portal vein<br />

(9.6 vs. 12.0 mm, p < 0.01) pre-B-RTO were significant in the<br />

HE group compared with in the GV group. Moreover, while<br />

splenic venous blood flow in the GV group was all hepatopetal,<br />

blood flow in 80% of the patients in the HE group was<br />

hepatofugal before the procedure. Whereas increased portal<br />

blood flow, elevated portal venous pressure, and expanded<br />

portal vein diameter following B-RTO were statistically significant<br />

in both groups, an improvement in CP score was more<br />

marked in the HE group (HE: 8.4 to 7.0 pts., p < 0.01; GV:<br />

6.2 to 5.9 pts., p < 0.05). Splenic vein diameter was significantly<br />

reduced in the HE group (12.7 to 9.9 mm, p < 0.01),<br />

meanwhile it was markedly increased in the GV group (8.9 to<br />

9.4 mm, p = 0.07), suggesting that splenic venous blood flow<br />

in the HE group changed from hepatofugal to hepatopetal after<br />

B-RTO. The decrease in blood ammonia levels in response to<br />

shunt occlusion was also significant in both groups, but it was<br />

statistically more marked in the HE group (HE: 161.5 to 95.2<br />

μg/dL, p < 0.01; GV: 90.8 to 69.4 μg/dL, p < 0.05). This<br />

effect lasted at least until 6 months after the procedure (before:<br />

161.5 μg/dL, after 1 month: 95.2 μg/dL, after 3 months:<br />

69.7 μg/dL, after 6 months: 64.6 μg/dL). Conclusions: Much<br />

like the dramatic effect on GV, B-RTO has an extremely strong<br />

and long-lasting effect on refractory HE. Moreover, compared<br />

with those treated for GV, patients treated for HE exhibit highly<br />

characteristic “hemodynamics in the portal–splenic venous system”<br />

both before and after the B-RTO procedure, which may<br />

lead to markedly improved hepatic function including ammonia<br />

metabolism.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tatsuro Nishimura, Tsuyoshi<br />

Ishikawa, Ryo Sasaki, Yuki Aibe, Shogo Shiratsuki, Kazuhito Matsunaga, Takuya<br />

Iwamoto, Taro Takami, Isao Sakaida<br />

1465<br />

Lactulose treatment improves cognition, quality of life<br />

and intestinal microbiota in minimal hepatic encephalopathy<br />

patients: results from a multi-center, randomized,<br />

controlled trial<br />

Jiyao Wang 1 , Jiang-bin Wang 2 , Jia Shang 3 , Xin-Min Zhou 4 , Xiao-<br />

Lin Guo 5 , Xuan Zhu 6 , Li-Na Meng 7 , Hai-Xing Jiang 8 , Yu-Qiang<br />

Mi 9 , Jian-Ming Xu 10 , Jin-Hui Yang 11 ; 1 GI & Hepatology, Zhong<br />

Shan Hospital,Fudan University, Shanghai, China; 2 Gastroenterology,<br />

China-Japan Union Hospital, Jilin University, Changchun,<br />

China; 3 Infectious Disease, Departmentof Infectious Diseases,<br />

Henan Provincial People’s Hospital, Zhengzhou, China; 4 Digestive<br />

Diseases,, Xijing Hospital, the Fourth Military Medical University,<br />

Xi’an, China; 5 Digestive Diseases, The First Hospital of Jilin University,<br />

Changchun, China; 6 Digestive Diease, The first Affiliated<br />

Hospital of Nanchang University, Nanchang, China; 7 Digestive<br />

Diseases, Zhejiang Provincial Hospital of Tranditional Chinese<br />

Medicine, Hangzhou, China; 8 Digestive Diseases, The First AffiliatedHospital<br />

of Guangxi Medical University,, Nanning, China;<br />

9 Hepatology, Tianjin Second People’s Hospital, Tianjin, China;<br />

10 Digestive Diseases, The First Affiliated Hospital of Anhui Medical<br />

University, Hefei, China; 11 Hepatology, The second Affiliated Hospital<br />

of Kunming Medical University, Kunming, China<br />

Aims As we investigated recently, minimal hepatic encephalopathy<br />

(MHE) is high prevalent in China (39.8 %) and those<br />

patients with low quality of life (QoL). Following previous findings,<br />

this study is to assess the efficacy and safety of lactulose<br />

in treatment of MHE in aspect of cognitive function, QoL, and<br />

intestinal microbiota. Methods It is a nationwide, multi-centered,<br />

randomized controlled trial. Patients with liver cirrhosis<br />

were screened by number connection test A (NCT-A) and digit<br />

symbol test (DST) both positive to determine MHE diagnose.<br />

These MHE patients were 2: 1 randomized to receive lactulose<br />

30-60ml/twice per day (Gp-L) or no therapy (Gp-NL). At month<br />

2, the psychometric tests (NCT-A and DST), the QoL questionnaire,<br />

the Child-Pugh score, the ammonia and the intestinal<br />

flora (metastats, QIIME) were reassessed. The primary endpoint<br />

was the recovery rate defined as the normalization of<br />

at least one of the psychometric tests. Results Of 575 patients<br />

screened, 98 (17.04%) met the inclusion and exclusion criteria<br />

(Gp-L n=67, Gp-NL n=31). In all, 11(11.22%) patients were<br />

lost to follow-up (Gp-L n=8, Gp-NL n=3). At month 2, the MHE<br />

recovery rate in Gp-L (41/59, 69.49%) is significantly higher<br />

than in Gp-NL (6/28, 21.43%) (p=0.0000). The QoL in Gp-L<br />

was significantly improved from baseline (p=0.0000) in all 5<br />

aspects of physical function, psychological well-being, relief of<br />

symptoms, social function and self-evaluation. The improved<br />

score of QoL questionnaire in Gp-L (12.63±15.10) was larger<br />

than Gp-NL (6.36±12.16) (p=0.0523). Two-month lactulose<br />

treatment significantly improved the severity of liver cirrhosis<br />

(assessed by Child-Pugh score) (p=0.0013) decreased ammonia<br />

level (delta= -63.16±3.60) (p=0.0000) from baseline.<br />

Furthermore, the intestinal microbiota changed in both autochthnous<br />

taxa and non-autochthnous taxa in Gp-L, an increase of<br />

Bifidobacteriaceae (2.9% to 5.8%), Lactobacillaceae (4.3% to<br />

6%), Ruminococcaceae Ruminococcus (1.2% to 2.2%), and a<br />

decrease of Rikenellaceae Alistipes (4.5% to 3.3%) were seen<br />

at the phylum/order level. All patients were tolerated lactulose,<br />

no severe side effects was reported. Conclusion Lactulose is<br />

effective for the recovery of MHE in aspect of cognition, QoL<br />

in patients with cirrhosis, which may be associated with the<br />

microbial changes. Regard this, the prophylactic therapy by<br />

lactulose is suggested for MHE patients. Clinical trial nr: ChiC-<br />

TR-TRC-12002342<br />

Disclosures:<br />

The following authors have nothing to disclose: Jiyao Wang, Jiang-bin Wang,<br />

Jia Shang, Xin-Min Zhou, Xiao-Lin Guo, Xuan Zhu, Li-Na Meng, Hai-Xing Jiang,<br />

Yu-Qiang Mi, Jian-Ming Xu, Jin-Hui Yang<br />

1466<br />

Transient elastography related cut off parameters in<br />

patients with advanced chronic liver disease independently<br />

predict decompensation and mortality but<br />

lack prediction of varices.<br />

Rafael Paternostro, Monika Ferlitsch, Alexandra Etschmaier,<br />

Remy Schwarzer, Thomas Reiberger, Mattias Mandorfer, Michael<br />

Trauner, Markus Peck-Radosavljevic, Arnulf Ferlitsch; Vienna<br />

Hepatic Hemodynamic Lab, Div. of Gastroenterology and Hepatology,<br />

Internal Medicine III, Medical University of Vienna, Vienna,<br />

Austria<br />

Background & Aims: Transient elastography (TE) is a validated<br />

tool to stage liver fibrosis and identify patients with advanced<br />

chronic liver disease (ACLD) by liver stiffness >10kPa. The<br />

Baveno VI consensus conference on portal hypertension<br />

proposed to select a TE value of 20kPa to stratify patients<br />

with ACLD. Aim of this study was to evaluate the predictive<br />

value of TE regarding (i) decompensation and (ii) mortality.<br />

Methods: 227 patients with chronic liver diseases were evaluated<br />

by TE. 55 patients had liver stiffness


926A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

included. 132 (85%) underwent additional HVPG measurements.<br />

Hepatic decompensation (ascites, HE, jaundice) and<br />

mortality was recorded during follow-up. Results: Baseline<br />

characteristics: age: 52±11years, 73% male, main etiologies:<br />

47.7% viral, 31% ALD, 21.3% others. Median TE values were<br />

29.1kPa (11.3-75). Mean HVPG was:15±6.5mmHg. Median<br />

follow up was 967 days(89.2-1609.2). In total 56/155<br />

patients decompensated or died during follow up. Hepatic<br />

decompensation or death occurred significantly more often in<br />

patients with TE>20kPa: 49/106 (46.2%) than in patients with<br />

TE20kPa is an independent risk factor<br />

for decompensation or mortality in ACLD patients. Considering<br />

that numbers of ACLD patients with TE150G/L are low, still around 50% showed varices – not<br />

supporting the Baveno VI recommendation to avoid screening<br />

endoscopy in this constellation. The 20kPa stratification strategy<br />

should be prospectively evaluated in larger cohorts.<br />

Disclosures:<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

Mattias Mandorfer - Consulting: Janssen; Speaking and Teaching: AbbVie, Gilead,<br />

Janssen, Boehringer Ingelheim, Bristol-Myers Squibb, Roche<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

The following authors have nothing to disclose: Rafael Paternostro, Monika Ferlitsch,<br />

Alexandra Etschmaier, Remy Schwarzer, Arnulf Ferlitsch<br />

1467<br />

Development of an Electronic Diary (E-Diary) for Caregivers<br />

(CG) of Patients with Overt Hepatic Encephalopathy<br />

(OHE)<br />

R Todd Frederick 1 , Marwan Ghabril 2 , Karin Coyne 3 , Mary Kay<br />

Margolis 3,4 , Michael Santoro 5 , Dion F. Coakley 5 , Masoud Mokhtarani<br />

5 , Bruce F. Scharschmidt 5 , Marzena Jurek 5 ; 1 California Pacific<br />

Medical Center, San Francisco, CA; 2 Indiana University, Indianapolis,<br />

IN; 3 Evidera, Bethesda, MD; 4 Patient-Centered Outcomes<br />

Research Institute (PCORI), Washington, DC; 5 Horizon Therapeutics<br />

(formerly Hyperion Therapeutics, Inc.), Brisbane, CA<br />

Care of patients with OHE requires education and involvement<br />

of the CG, as most episodes occur outside of the hospital or<br />

clinic. However, some CG may try to manage the episodes<br />

without informing or receiving input from the medical provider.<br />

Moreover, lack of systematic patient evaluation and/or communication<br />

with the provider by CG may hinder identification of<br />

OHE episodes in clinical practice and represents a challenge<br />

in the design and interpretation of OHE endpoints in clinical<br />

trials. Study Design: A 3 part approach was used to develop a<br />

daily e-diary for CG of OHE patients. Part 1 was conducted at<br />

2 clinical sites and included concept elicitation, i.e. interviews<br />

with 12 CG who had cared for a patient with OHE within the<br />

last three months, to determine the signs and symptoms CG use<br />

to identify OHE. In Part 2, a daily e-diary for CG was developed<br />

that included questions with branching logic and skip<br />

patterns for standardized feedback, automatic reminders for<br />

completion, and real time alerts sent to the site if CG answers<br />

were suggestive of OHE. This e-diary was evaluated for comprehensiveness<br />

and understandability by 10 CG in semi-structured<br />

interviews. Part 3 was a 3-day field test among 10 CG<br />

to assess real-world usability on their own devices. The final<br />

e-diary was translated and culturally adapted with CG input<br />

into 16 languages. Results: Part 1 CG collectively described<br />

33 different HE-related symptoms experienced by the patients<br />

they cared for. Forgetfulness/confusion and sleep difficulties<br />

(e.g., day-night inversion) were reported by all CG; most also<br />

described speech problems (difficulty finding words [92%] or<br />

repeating words or phrases [50%]). No one symptom emerged<br />

as most important and most CG (83%) reported that they did<br />

not contact a physician unless there was progression or non-resolution<br />

of OHE symptoms. Part 2 interviews identified 7 items<br />

(speech difficulties, unusual behavior, forgetfulness/confusion,<br />

orientation, and level of consciousness), which adequately and<br />

comprehensively captured what CG observe and monitor at<br />

home and that the e-format was user friendly. Part 3 confirmed<br />

the usability of the e-diary on the CG’ personal devices, including<br />

web-enabled smart phones (iPhone, Android), laptop, and<br />

desktop computers. Conclusion: A comprehensive, easy to<br />

use CG e-diary was developed for use in clinical trials of OHE<br />

and in clinical practice. Real time communication with health<br />

care providers and standardized evaluation of manifestations<br />

of OHE by CG will improve documentation of endpoints in<br />

trials of OHE and may lead to earlier medical intervention and<br />

improved patient outcomes in clinical practice.<br />

Disclosures:<br />

R Todd Frederick - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, Salix, Vital Therapies, Hyperion, AbbVie<br />

Marwan Ghabril - Grant/Research Support: Salix<br />

Karin Coyne - Consulting: Evidera<br />

Michael Santoro - Employment: Horizon Therapeutics, Inc.<br />

Dion F. Coakley - Employment: Horizon Therapeutics<br />

Masoud Mokhtarani - Employment: Hyperion; Stock Shareholder: Hyperion<br />

Marzena Jurek - Employment: Horizon Therapeutics, Inc.<br />

The following authors have nothing to disclose: Mary Kay Margolis, Bruce F.<br />

Scharschmidt<br />

1468<br />

Muscle mass in liver transplant candidates: From the Fitness,<br />

Life Enhancement, and Exercise in Transplantation<br />

(FLEXIT) Consortium<br />

Jennifer C. Lai 1 , Elizabeth J. Carey 2 , Connie W. Wang 1 , Srinivasan<br />

Dasarathy 3 , Michael A. Dunn 4 ; 1 UCSF, San Francisco, CA;<br />

2 Mayo Clinic Arizona, Scottsdale, AZ; 3 Cleveland Clinic, Cleveland,<br />

OH; 4 University of Pittsburgh, Pittsburgh, PA<br />

Background: Sarcopenia, frailty, and deconditioning are distinct<br />

terms for anatomic and functional disturbances related to<br />

muscle wasting, a frequent and often lethal manifestation of<br />

advanced cirrhosis. We hypothesized that sarcopenia measured<br />

on cross sectional imaging as an indicator of muscle<br />

wasting could be consistently evaluated as a common parameter,<br />

and that its clinical significance in liver transplant candidates<br />

from multiple centers would confirm earlier single-center


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 927A<br />

reports. Methods: Adult liver transplant candidates listed in<br />

2012 with an abdominal CT scan within 3 months of listing<br />

were included. Our 4 centers executed a data use agreement<br />

and agreed on common definitions of clinical variables, key<br />

parameters to measure muscle mass, and outcomes. Abdominal<br />

skeletal muscle area was measured as the sum of cross-sectional<br />

areas of psoas, paraspinal, and abdominal wall muscles<br />

at L3, normalized for height (SMI, cm 2 /m 2 ). Associations<br />

between SMI and baseline characteristics were evaluated<br />

by Pearsons correlations, Wilcoxon, or Kruskal-Wallis tests.<br />

Adjusted competing risks analysis evaluated association of SMI<br />

with death/delisting for being too sick for LT. Results: Included<br />

were 224 subjects: 25% female, 54% non-Hispanic White,<br />

54% HCV, 14% ETOH, 11% NASH, 46% HCC, median age<br />

59y, median BMI 28. Disease-related characteristics: median<br />

MELD 14, median albumin 3.1, 49% with ascites. Median<br />

(IQR) total abdominal skeletal muscle index was 46 cm 2 /m 2<br />

(36-55). SMI had low correlation with MELD (r=-0.27;p


928A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

analysis identified altered levels of metabolites associated with<br />

bile acids, heme metabolism, sphingosine and lipids in HPS<br />

relative to controls. Levels of 6 of 8 primary bile acids and<br />

only 2 of 19 secondary bile acids were increased in HPS relative<br />

to control (others unchanged) suggesting altered excretion<br />

and gut microbial metabolism that may affect the vasculature.<br />

Products of heme metabolism including bilirubin, biliverdin and<br />

urobilinogen were also increased in HPS suggesting enhanced<br />

hemoxygenase driven heme metabolism and/or impaired<br />

clearance which may increase carbon monoxide production.<br />

Levels of sphingosine metabolites, potent regulators of endothelin<br />

and nitric oxide synthase signaling, were also elevated in<br />

HPS. Finally, global increases in fatty acid levels with reduced<br />

monoglycerol production were found in HPS suggesting<br />

enhanced fatty acid release or decreased clearance. Random<br />

forest analysis showed that selected metabolites distinguished<br />

HPS from controls with 71% accuracy. Conclusion: Human HPS<br />

has plasma metabolic signatures consistent with alterations<br />

in sphingosine, bile acid, heme and fatty acid metabolism,<br />

suggesting novel mechanistic pathways and possible links to<br />

experimental HPS. Validation in larger cohorts is required.<br />

Disclosures:<br />

Michael B. Fallon - Grant/Research Support: Bayer/Onyx, Eaisi, Gilead, Grifolis<br />

The following authors have nothing to disclose: Michael J. Krowka, Kimberly A.<br />

Forde, Karen Krok, MAMTA PATEL, Grace Lin, Jae K. Oh, Carl Mottram, Paul D.<br />

Scanlon, Sachin Batra, David S. Goldberg, Steven M. Kawut<br />

and of BMI 27.12 ± 5.24. The frequency of malnutrition was<br />

14.7% for BIA, 25% for BIVA, 32% for MAMC, 13% for TSF,<br />

and 14 for % BMIa, In the Rxc graph only patients with clinical<br />

ascites displayed fluid retention by BIVA (55.1% vs 44.9%<br />

without ascites, p=0.005). Also, higher stages of Child-Pugh<br />

were significantly associated with a higher rate of malnutrition<br />

and fluid retention (data not shown). Kaplan-Meier curves disclosed<br />

a higher mortality in the malnourished group identified<br />

by BIVA (39.7% vs. 21.4% in non malnourished, p=


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 929A<br />

were significant. On multi-variate analysis, recent alcohol (OR<br />

4.0, p=0.003) and illegal turns (OR 1.4, p=0.005) remained<br />

significant predictors of real-life crashes. Conclusion: In this<br />

largest-studied cirrhosis cohort, MHE subjects were likely to<br />

have worse simulator and real-life crash rates that remains<br />

stable over 1 year. Only MHE patients who performed poorly<br />

on navigation and had quit alcohol within 3 years were likely<br />

to be unsafe drivers.<br />

Summarized of results<br />

Table shows summarized of results from driving simulation, real<br />

life traffic history and 6-month follow up in 163 patients.<br />

Disclosures:<br />

Richard K. Sterling - Advisory Committees or Review Panels: Merck, Vertex, Salix,<br />

Bayer, BMS, Abbott, Gilead, Baxter, Jansen; Grant/Research Support: Merck,<br />

Roche/Genentech, Pfizer, Gilead, Boehringer Ingelheim, Bayer, BMS, Abbott<br />

Puneet Puri - Advisory Committees or Review Panels: Health Diagnostic Laboratory<br />

Inc.; Consulting: NPS Pharmaceuticals Inc.<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Douglas M. Heuman - Consulting: Bayer, Grifols, Genzyme; Grant/Research<br />

Support: Exilixis, Novartis, Bayer, Bristol Myers Squibb, Scynexis, Ocera, Mannkind,<br />

Salix, Globeimmune, Roche, SciClone, Wyeth, Otsuka, Ikaria, UCB, Celgene,<br />

Centocor, Millenium, Osiris, AbbVie, Gilead; Speaking and Teaching:<br />

Otsuka, Astellas<br />

Jasmohan S. Bajaj - Advisory Committees or Review Panels: Salix, Merz, otsuka,<br />

ocera, grifols, american college of gastroenterology; Grant/Research Support:<br />

salix, otsuka, grifols<br />

The following authors have nothing to disclose: Mette M. Lauridsen, Vishwadeep<br />

Ahluwalia, Ariel Unser, Melanie White, R. Todd Stravitz, Scott C. Matherly,<br />

Velimir A. Luketic, Mohammad S. Siddiqui, Michael Fuchs<br />

1473<br />

Randomized Controlled Trial on Efficacy of Nutritional<br />

Therapy on Cognitive Functions and Health Related<br />

Quality of Life in Patients of Liver Cirrhosis with Minimal<br />

Hepatic Encephalopathy<br />

Sudhir Maharshi, Barjesh C. Sharma, Sanjeev Sachdeva, Siddharth<br />

Srivastava; Gastroenterology, G B Pant Institute of Post<br />

graduate Medical Education and Research, New delhi, India<br />

Background & Aims: Minimal hepatic encephalopathy impairs<br />

health related quality of life (HRQOL), predicts development of<br />

overt hepatic encephalopathy and associated with poor prognosis.<br />

There is no study on nutritional management in patients<br />

with minimal hepatic encephalopathy. We assessed the effects<br />

of nutritional therapy on cognitive functions and HRQOL in<br />

patients of cirrhosis with minimal hepatic encephalopathy.<br />

Methods: A randomized controlled trial conducted in a tertiary<br />

care setting on patients of cirrhosis with minimal hepatic<br />

encephalopathy who were randomized to nutritional therapy<br />

(group A: 30-35 kcal/kg/day and 1 . 0-1 . 5 gram of vegetable<br />

protein/kg/day) and no nutritional therapy (group B: diet as<br />

patients were taking before) for 6 months. Minimal hepatic<br />

encephalopathy was diagnosed based on psychometry hepatic<br />

encephalopathy score (PHES). HRQOL was assessed by sickness<br />

impact profile (SIP) questionnaire. Primary endpoints were<br />

improvement or worsening in minimal hepatic encephalopathy<br />

and improvement in HRQOL. Results: 120 patients were<br />

randomized to group-A (n = 60, age 42 . 1±10 . 3 yr, 48 men)<br />

and group-B (n = 60, age 42 . 4±9 . 6 yr, 47 men). There was<br />

no significant difference in baseline characteristics between<br />

the two groups. Baseline PHES (-8 . 12±1 . 32 vs -8 . 53±1 . 38;<br />

p=0 . 08) and SIP score (14 . 25±5.8 vs 15 . 44±5 . 03; p=0 . 85)<br />

were comparable in both the groups. Reversal of minimal<br />

hepatic encephalopathy was higher in group A (71 . 1% vs<br />

22 . 8%; p=0 . 001). Improvement in PHES (ΔPHES 3 . 86±3 . 58<br />

vs 0 . 52±4 . 09; p=0 . 001) and HRQOL (Δ SIP 3 . 24±3 . 63 vs<br />

0 . 54±3 . 58; p=0 . 001) were also higher in group A compared<br />

to group B. Overt hepatic encephalopathy developed in 10%<br />

of patients in group A vs 21 . 7% in group B (P=0 . 04). Conclusions:<br />

Nutritional therapy is effective in treatment of minimal<br />

hepatic encephalopathy and associated with improvement in<br />

HRQOL.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sudhir Maharshi, Barjesh C.<br />

Sharma, Sanjeev Sachdeva, Siddharth Srivastava<br />

1474<br />

Usefulness of Balloon-Occluded Retrograde Obliteration<br />

(B-RTO) as a Therapeutic Procedure to Improve Liver<br />

Function in Cirrhotic Patients with Porto-Systemic Shunts<br />

Manabu Nakazawa, Yukinori Imai, Satsuki Ando, Kayoko Sugawara,<br />

Mie Inao, Nobuaki Nakayama, Satoshi Mochida; Gastroenterology<br />

& Hepatology, Saitama medical university hospital,<br />

Saitama, Japan<br />

Aim: Porto-systemic shunts cause refractory hepatic encephalopathy<br />

in patients with liver cirrhosis. The shunts also contribute<br />

to derange liver function through reduction of blood flow in<br />

the portal vein. The usefulness of B-RTO as a therapeutic procedure<br />

for improvement of liver function as well as attenuation of<br />

refractory hepatic encephalopathy was evaluated in cirrhotic<br />

patients. Methods: The subjects were 34 cirrhotic patients with<br />

porto-systemic shunts showing refractory hepatic encephalopathy.<br />

The extent of liver dysfunction classified according to the<br />

Child-Pugh classification were grade A, B and C in 3, 18 and<br />

13 patients, respectively. A balloon catheter was inserted into<br />

the shunts followed by injection of 5% ethanolamine oleate<br />

through the catheter under balloon inflation. The balloon was<br />

kept inflation for between 6 and 48 hours depending on<br />

sizes of the shunts. Results: Drainage vessels of the targeted<br />

shunts were the splenorenal, mesocaval, paraumbilical, azygos<br />

and splenorenal plus azygos veins in 20, 4, 6, 2 and<br />

2 patients, respectively. B-RTO was successfully done in 29<br />

patients (85%), and hepatic encephalopathy did not recur in<br />

24 (83%) patients. In patients receiving successful B-RTO procedures,<br />

HH15, LHL15 and LU15 values assessed by Tc-99m<br />

GSA scintigraphy were not different between at baseline and<br />

1 months after the procedures, while doppler ultrasound examinations<br />

revealed that blood flows in the portal vein (cm 3 /sec;<br />

mean±SD) increased significantly (from 2.2±0.9 to 8.4±2.4,<br />

p


930A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

therapeutic procedure to improve liver function as well as to<br />

attenuate hepatic encephalopathy through increase of portal<br />

venous flows in cirrhotic patients with porto-systemic shunts.<br />

Disclosures:<br />

Satoshi Mochida - Grant/Research Support: Chugai, MSD, Tioray Medical,<br />

BMS; Speaking and Teaching: MSD, Toray Medical, BMS, Tanabe Mitsubishi<br />

The following authors have nothing to disclose: Manabu Nakazawa, Yukinori<br />

Imai, Satsuki Ando, Kayoko Sugawara, Mie Inao, Nobuaki Nakayama<br />

1475<br />

Six-week administration of lactulose in patients with cirrhosis<br />

does not alter the gut bacterial flora<br />

Amit Goel 1 , Aditya N. Sarangi 2 , Ankur Singh 1 , S. Avani 1 , Rakesh<br />

Aggarwal 1,2 ; 1 Gastroenterology, Sanjay Gandhi Postgraduate<br />

Institute of Medical Sciences, Lucknow, India; 2 Biomedical informatics<br />

Center, Sanjay Gandhi Postgraduate Institute of Medical<br />

Sciences, Lucknow, India<br />

Background: Benefit of lactulose in hepatic encephalopathy<br />

(HE) may be mediated, at least partially, by alterations in gut<br />

flora. However, data on change in gut flora post-lactulose are<br />

sparse. We therefore studied gut microbial composition in<br />

patients with liver cirrhosis (LC) and the effect of 6 weeks of<br />

lactulose use on it, using next-generation sequencing. Methods:<br />

30 patients with LC (male 22; median [range] age: 42 [29-65]<br />

y; body mass index 22.7 [17.3-32.3] Kg/m 2 ) of any cause<br />

or severity but no prior history of HE, and 18 healthy controls<br />

(HC; male 14; 45 [24-67] y; 23.3 [20.0-25.0] Kg/m 2 ) were<br />

enrolled. Persons with gastrointestinal symptoms, or recent (preceding<br />

6 weeks) use of lactulose or another drug affecting<br />

gut motility or flora were excluded. A morning stool specimen<br />

was collected from each subject; 19 patients also provided<br />

a follow-up specimen after 6 weeks of lactulose intake. A<br />

cDNA library for V3 region of bacterial 16S rRNA from each<br />

stool was subjected to paired-end Illumina sequencing, and<br />

abundances of various bacterial operational taxonomic units<br />

(OTUs) were determined. Relationship of specimens with each<br />

other was studied using principal co-ordinate analysis (PCoA).<br />

Abundances of various OTUs, and indices of α and β diversity<br />

were compared between groups using t-test with Bonferroni<br />

correction; data before and after lactulose were compared<br />

using tests for paired data. Results: The 67 specimens yielded a<br />

median (range) of 504,182 (171,799-1,267,206) high-quality<br />

reads. In PCoA, gut flora from LC clustered separately from<br />

HC and showed lower diversity (Chao 1 index: 1975 vs 2802;<br />

observed species: 1082 vs 1485; Shannon index: 4.8 vs 5.6;<br />

p=0.0001 each). Median abundances of major phyla (Bacteroidetes<br />

[75.2% vs 67.2%], Firmicutes [17.5% vs 18.4%]<br />

and Proteobacteria [6.6% vs 8.3%]) were similar in LC and<br />

HC; however, Lentisphaerae [0% vs 0.01%], Cyanobacteria<br />

[0.01% vs 0.52%] and Tenericutes [0.01% vs 0.05%] were<br />

less abundant in LC than HC (p0.005%<br />

of all the bacterial sequences had comparable abundances<br />

before and after lactulose. The pre- and post-lactulose specimens<br />

showed no separation on PCoA, and had similar α diversity<br />

indices (Chao 1: 2755 vs 2788; observed species 1732<br />

vs 1761; Shannon index: 4.9 vs 4.9; all p=ns). Conclusion:<br />

Gut flora in LC differs from healthy persons. However, administration<br />

of lactulose for 6 weeks did not produce a discernible<br />

change in gut flora in patients with LC, suggesting that alteration<br />

in gut flora may play a minor role in improving HE.<br />

Disclosures:<br />

The following authors have nothing to disclose: Amit Goel, Aditya N. Sarangi,<br />

Ankur Singh, S. Avani, Rakesh Aggarwal<br />

1476<br />

Erectile Dysfunction in patients with cirrhosis<br />

Sergio Maimone 1 , Giovanni Oliva 1,2 , Antonino Di Benedetto 2 ,<br />

Roberto Filomia 1,2 , Gaia Caccamo 1 , Carlo Saitta 1 , Irene Cacciola<br />

1,2 , Giovanni Squadrito 1,3 , Giovanni Raimondo 1,2 ; 1 Division<br />

of Clinical and Molecular Hepatology, Universitary Hospital of<br />

Messina, Messina, Italy; 2 Department of Clinical and Experimental<br />

Medicine, University Hospital of Messina, Messina, Italy; 3 Department<br />

of Human Pathology, University Hospital of Messina, Messina,<br />

Italy<br />

BACKGROUND AND AIMS - Erectile dysfunction (ED) is a clinical<br />

disorder frequently observed in men suffering from several<br />

chronic diseases as diabetes, obesity, metabolic syndrome,<br />

cardiovascular diseases. There is little information, however,<br />

about ED in patients with advanced chronic liver disease. Aims<br />

of our study were to evaluate the prevalence of ED in cirrhotic<br />

patients and to investigate clinical and biochemical factors<br />

possibly associated with ED occurrence in these patients.<br />

PATIENTS AND METHODS - We prospectively analyzed 147<br />

consecutive cirrhotic patients (mean age 61.5± 15.5 years) 99<br />

of whom were in class A and 48 in class B of Child-Pugh (CP).<br />

Physical examination, electrocardiography, biochemistry as<br />

well as serum levels of thyroid stimulating hormone (TSH), total<br />

testosterone (TT), free testosterone (FTT), sex hormon binding<br />

globulin (SHBG), and prolattin (PRL) were evaluated in each<br />

patient. Known ED risk factors such as smoking habit, alcohol<br />

abuse, diabetes, cardiovascular diseases (i.e., hypertensive<br />

cardiovascular disease, ischemic heart disease), depression,<br />

hypogonadism, treatment with diuretics and beta blockers<br />

were evaluated as well. All patients completed the following<br />

questionnaires: International Index of Erectile Function (IIEF)<br />

scale (that classifies ED into 4 categories: absent, mild, moderate<br />

and severe), Self-rating Anxiety Scale (SAS) (that evaluates<br />

grades of depression), and ANDROTEST (that assesses<br />

the presence of hypogonadism jointly to serum level of FTT).<br />

Data processing and statistical analysis were performed by<br />

SPSS software. RESULTS - ED was revealed in 86/147 (58.5%)<br />

patients, 31 (36%) of whom had mild, 7 (8.1%) moderate and<br />

48 (55.9%) severe ED. No statistically significant differences<br />

were found when patients with and those without ED were<br />

compared for age, smoking habit, alcohol consumption, ethiology<br />

of liver disease, concomitant arterial hypertension, therapy<br />

with diuretics and beta blockers, depression, thyroid and sexual<br />

hormone profiles. In analogy, SAS test and ANDROTEST<br />

provided similar results in the two groups. On the contrary,<br />

at univariate logistic regression analysis ED was significantly<br />

associated with CP class B (p=0.04), presence of diabetes<br />

(p=0.04), and cardiovascular diseases (p=0.01). However,<br />

only presence of cardiovascular disease maintained the significant<br />

association with ED occurrence at the multivariate logistic<br />

regression analysis (OR 3.7, CI 95%1.06-13.44, p=0.03).<br />

CONCLUSIONS - Erectile dysfunction affects the majority of<br />

patients with cirrhosis and it is significantly associated with<br />

cardiovascular diseases in these patients.<br />

Disclosures:<br />

Giovanni Raimondo - Speaking and Teaching: BMS, Gilead, Roche, Merck,<br />

Janssen, Bayer, MSD<br />

The following authors have nothing to disclose: Sergio Maimone, Giovanni<br />

Oliva, Antonino Di Benedetto, Roberto Filomia, Gaia Caccamo, Carlo Saitta,<br />

Irene Cacciola, Giovanni Squadrito


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 931A<br />

1477<br />

Perceptions of the ‘cirrhotic coagulopathy’ influence low<br />

yield testing for cerebral hemorrhage in patients with<br />

cirrhosis and hepatic encephalopathy: an international,<br />

multi-specialty survey<br />

Elliot B. Tapper 1 , Laura Mazer 2 ; 1 Gastroenterology, Beth Israel<br />

Deaconess Medical Center, Boston, MA; 2 Surgery, Beth Israel Deaconess<br />

Medical Center, Boston, MA<br />

Patients with cirrhosis and hepatic encephalopathy (HE) are<br />

often evaluated for intracranial hemorrhage (ICH) even in the<br />

absence of focal neurologic findings or trauma due to perceptions<br />

of a tendency toward bleeding - the so-called ‘cirrhotic<br />

coagulopathy.’. There is limited data on the prevalence and<br />

clinical impact of such perceptions. Methods: We conducted<br />

a prospective survey of physicians in internal medicine (IM),<br />

emergency medicine (EM), general surgery (GS) and gastroenterology<br />

(GI) at 35 centers in 5 countries. All respondents<br />

were presented with a standard HE patient presentation (somnolence,<br />

no focal neurologic findings, no evidence of a fall/<br />

trauma). Subjects rated their likelihood to order computed<br />

tomography (CT) of the head to exclude ICH on a 7 point<br />

scale from -3 to +3.They were asked to re-rate their likelihood<br />

after seeing the patient’s labs (including INR 2.4 and platelet<br />

count 59) and again after seeing the results of a study<br />

demonstrating the low yield of CT for this presentation (CGH<br />

2015;13(1):165). Results: 1033 physicians responded, 86%<br />

from the US. The proportions of IM, ED, GS and GI were 55%,<br />

29%, 7% and 9%. The proportion of attendings, fellows and<br />

residents were 42%, 7% and 49%. The baseline tendency to<br />

order a CT was lowest among GI attendings, median -2 interquartile<br />

range (IQR) (-2 – 0.75) and highest among EM physicians,<br />

2 IQR (1-3). At baseline 56.0% of respondents were<br />

likely to order a head CT. All groups (by specialty, level of<br />

training and familiarity with cirrhosis) were significantly more<br />

likely (p < 0.001 for all and to 0.03 for attending GI) to order<br />

a head CT after seeing the labs. Of the subjects unlikely to<br />

order a head CT at baseline, 114 (25.4%) would order after<br />

the labs. All groups were less likely to order head CTs after<br />

seeing the research data (mean difference -1.51 95% CI (-1.61<br />

- -1.41), p < 0.0001). Of the 651 likely to order a head CT<br />

after the labs, 313 (48.1%) were dissuaded by the published<br />

yield data. 61.2% of IM physicians interested in head CTs<br />

were most likely to report disinterest after exposure to the data<br />

compared to 31.3% of EM physicians, p = 0.0002. Conclusions:<br />

Misperceptions regarding the coagulopathy of cirrhosis<br />

are highly prevalent and affect clinical decision-making, but<br />

are amenable to education.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elliot B. Tapper, Laura Mazer<br />

1478<br />

Prospective Multicenter Observational Study of Overt<br />

Hepatic Encephalopathy (OHE): Nomenclature and Disease<br />

Burden<br />

Charles S. Landis 1 , Marwan Ghabril 2 , Vinod K. Rustgi 3 , Adrian<br />

M. Di Bisceglie 4 , Benedict Maliakkal 5 , Don C. Rockey 6 , John M.<br />

Vierling 7 , Richard Rowell 8 , Michael Santoro 8 , Aimee Enriquez 8 ,<br />

Marzena Jurek 8 , Masoud Mokhtarani 8 , Dion F. Coakley 8 , Bruce F.<br />

Scharschmidt 8 ; 1 University of Washington, Seattle, WA; 2 Indiana<br />

University, Indianapolis, IN; 3 University of Pittsburgh, Pittsburgh,<br />

PA; 4 St Louis University, St Louis, MO; 5 University of Rochester,<br />

Rochester, NY; 6 Medical University of South Carolina, Charleston,<br />

SC; 7 Baylor College of Medicine, Houston, TX; 8 Horizon Therapeutics<br />

(formerly Hyperion Therapeutics, Inc.), Brisbane, CA<br />

Overview: OHE, a complication of advanced liver disease, is<br />

devastating to patients and families, burdensome to society,<br />

and challenging with respect to both diagnosis and nomenclature,<br />

evidenced by recent ISHEN and AASLD/EASL guidelines.<br />

However, there is little prospectively collected data regarding<br />

either the presenting manifestations of OHE or its associated<br />

economic burden, as reflected, for example, by frequency of<br />

recurrence and hospitalizations. Study Design: Surveys pertaining<br />

to OHE were sent to over 200 centers in 23 countries;<br />

176 centers (67% university and 29% public), which followed<br />

a median of ~40 patients with OHE responded. A subset of<br />

centers, which followed a median of ~100 OHE patients was<br />

invited to participate in an observational study. To be eligible<br />

for enrollment, patients were required to have a clinical diagnosis<br />

of cirrhosis and at least 1 episode of OHE within 30<br />

days prior to consent (qualifying OHE). Once enrolled, patients<br />

were followed to capture OHE recurrence (on-study OHE). Clinical<br />

manifestations of both qualifying and on-study OHE episodes<br />

as recorded in the patients’ charts were noted. Results:<br />

269 patients with a qualifying OHE event were enrolled at 31<br />

sites in North America and Europe; data from 260 patients<br />

who were followed for up to 8 months (mean=73 days) were<br />

available for this interim analysis (median age=61; median<br />

time since OHE diagnosis=1.1 years, 61% male, 87% Caucasian).<br />

The most frequently documented clinical manifestations<br />

of both qualifying and on-study OHE episodes included confusion<br />

(73%), change in mental status (56%), disorientation<br />

(46%), lethargy (44%) and asterixis (44%). West Haven grade<br />

was recorded in only 27% of all OHE episodes. Fifty seven<br />

patients had 101 on-study OHE episodes. The median time<br />

for occurrence of 2 nd , 3 rd and 4 th OHE episodes was 28, 19<br />

and 11 days, respectively. Of 87 and 47 patients followed for<br />

at least 3 or 4 months, 32% and 34% experienced a mean<br />

(SD) of 0.7 (1.5) or 0.9 (1.8) on-study OHE episodes, respectively.<br />

57% of qualifying and 84% of on-study OHE episodes<br />

occurred while patients were on rifaximin; 65% of qualifying<br />

and 91% of on-study OHE episodes were associated with<br />

hospitalization. Conclusions: Among patients enrolled in this<br />

multi-center and multi-national observational study, clinical disorientation,<br />

level of consciousness and asterixis were the most<br />

commonly documented manifestations of OHE, whereas West<br />

Haven was infrequently recorded. Despite the availability of<br />

rifaximin, OHE, as reflected by frequent recurrence and hospitalizations,<br />

continues to represent a major unmet clinical need<br />

and societal burden.<br />

Disclosures:<br />

Charles S. Landis - Grant/Research Support: Gilead, Abbvie, BMS<br />

Marwan Ghabril - Grant/Research Support: Salix<br />

Vinod K. Rustgi - Advisory Committees or Review Panels: Abbvie, BMS, Merck;<br />

Grant/Research Support: Gilead<br />

Adrian M. Di Bisceglie - Advisory Committees or Review Panels: Gilead, AbbVie,<br />

Novartis, Bayer, BTG; Grant/Research Support: Gilead, AbbVie


932A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Benedict Maliakkal - Speaking and Teaching: Valeant Pharma (Salix), Gilead,<br />

AbbVie, Bristol Myers Squibb<br />

Don C. Rockey - Grant/Research Support: Gilead, Actelion, Hyperion<br />

John M. Vierling - Advisory Committees or Review Panels: Abbvie, Bristol-Meyers-Squibb,<br />

Gilead, Hyperion, Intercept, Janssen, Novartis, Merck, Sundise,<br />

HepQuant, Salix, Immuron, Exalenz, Chronic Liver Disease Foundation; Board<br />

Membership: Clinical Research Centers of America, LLC; Grant/Research Support:<br />

Abbvie, Bristol-Meyers-Squibb, Eisai, Gilead, Hyperion, Intercept, Janssen,<br />

Novartis, Merck, Sundise, Ocera, Mochida, Immuron, Exalenz, Conatus; Speaking<br />

and Teaching: GALA, Chronic Liver Disease Foundation, ViralEd, Chronic<br />

Liver Disease Foundation, Clinical Care Options<br />

Michael Santoro - Employment: Horizon Therapeutics, Inc.<br />

Aimee Enriquez - Employment: Horizon Pharma<br />

Marzena Jurek - Employment: Horizon Therapeutics, Inc.<br />

Masoud Mokhtarani - Employment: Hyperion; Stock Shareholder: Hyperion<br />

Dion F. Coakley - Employment: Horizon Therapeutics<br />

The following authors have nothing to disclose: Richard Rowell, Bruce F.<br />

Scharschmidt<br />

1479<br />

Sarcopenia predicts moratility after transjugular intrahepatic<br />

portosystemic shunt placement<br />

Joshua Anderson 1,3 , Alagappan A. Annamalai 1 , Joseph L. Robinson<br />

1 , Tram T. Tran 1 , Walid S. Ayoub 1 , Jonah Hirschbein 2 , Georgios<br />

Voidonikolas, 1 , Irene K. Kim 1 , Vladamir Donchev 1 , Nicholas<br />

N. Nissen 1 , Andrew S. Klein 1 , Vinay Sundaram 1 ; 1 Cedars-Sinai<br />

Medical Center, Los Angeles, CA; 2 Radiology, Univeristy of California<br />

Davis Medical Center, Sacramento, CA; 3 University of Florida<br />

Health Science Center, Jacksonville, FL<br />

Aim: Sarcopenia is a marker of frailty and predicts mortality<br />

in cirrhotic patients. Our aim was to determine if sarcopenia<br />

predicted mortality after transjugular intrahepatic portosystemic<br />

shunt (TIPS) placement, when adjusting for Model for Endstage<br />

Liver Disease (MELD) score Methods: We retrospectively<br />

reviewed patients who underwent TIPS procedure at a single<br />

center from 2003-2014. CT images were used to determine<br />

sarcopenia, using measurement of L3 skeletal muscle index<br />

compared to gender specific thresholds of ≤38.5 cm2/m2 for<br />

women and ≤52.4 cm2/m2 for men. Cox proportional hazards<br />

regression was used to determine risk factors for mortality<br />

after TIPS. Results: A total of 96 patients were included,<br />

of which 52 had sarcopenia and 44 did not. Median length<br />

of follow-up was 303 days (range 3-4035 days). Indication<br />

for TIPS were primarily ascites (n=34, 35.8%) and variceal<br />

bleeding (n=56, 58.9%). Sarcopenic patients had a greater<br />

proportion of males (77% vs 38%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 933A<br />

Myocardial blood flow pre and post-exercise in patients<br />

with compensated cirrhosis compared with healthy controls<br />

(HV=healthy, CLD=liver disease)<br />

6.2, P=0.001, and OR 7.0, 95% CI 2.4-21, P


934A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

for primary and secondary prophylaxis should target this predisposition<br />

and points to an important area of future research.<br />

Disclosures:<br />

Rajiv Jalan - Consulting: Ocera Therapeutics, Conatus; Grant/Research Support:<br />

Grifols, Gambro; Patent Held/Filed: Yaqrit, Cyberliver<br />

The following authors have nothing to disclose: Rosaria Spinella, Rohit Sawhney,<br />

Peter Holland-Fischer, Naina Shah, Rajeshwar Mookerjee<br />

1483<br />

Progressive Loss of Hematopoietic Stem Cells and Associated<br />

Niche Cells in Advanced Cirrhosis<br />

Chhagan Bihari, Sheetalnath B. Rooge, Dhananjay Kumar, Priyanka<br />

Saxena, Lovkesh Anand, Smriti Shubham, Archana Rastogi,<br />

Anupam Kumar, Shiv K. Sarin; Institute of Liver and Biliary Sciences,<br />

Delhi, India<br />

Background: Bone marrow (BM) is a reservoir for immunological<br />

and hematological cells and dysfunction of both has been<br />

reported in cirrhosis. The mechanisms underlying these dysfunctions<br />

are not well understood. It is important to know the state<br />

of BM in cirrhotics, as hepatic regeneration using growth factors<br />

shown some promise. Aim: To study the effects of disease<br />

severity and portal hypertension on bone marrow response in<br />

cirrhosis. Patients and Methods: BM were done as part of a<br />

work up of cirrhotics (n=168) undergoing regenerative therapy<br />

protocol or pancytopenia over a period of 4 years. BM without<br />

any pathology from non-liver, pyrexia of unknown origin cases<br />

served as controls (n=44). BM assessed for routine examination<br />

and immunohistochemistry (IHC) for HSCs (CD34), and<br />

associated niche cells using cell type specific markers; Nestin<br />

(mesenchymal stem cells), S-100(Schwann cells), GFAP(neural<br />

fibres). BM HSCs were enumerated by fluorescence activated<br />

cell sorting. Cytokines and growth factors were analyzed in<br />

BM plasma (Cirrhosis=26, control= 6). Results: Routine BM in<br />

cirrhotics showed low cellularity, greater reticulinization, erythroid<br />

colony disarray as compared to controls. IHC showed<br />

significant reduction in CD34+ HSCs with increase in MELD (r=<br />

-0.545; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 935A<br />

1485<br />

Development of a Novel Clinician Reported Outcome<br />

Tool for Assessment of Hepatic Encephalopathy<br />

Jasmohan S. Bajaj 1,2 , Nathan M. Bass 3 , R Todd Frederick 4 ,<br />

Karin Coyne 5 , Mary Kay Margolis 5,6 , Dion F. Coakley 7 , Masoud<br />

Mokhtarani 7 , Marzena Jurek 7 , Bruce F. Scharschmidt 7 ; 1 Virginia<br />

Commonwealth University, Richmond, VA; 2 McGuire VA Hospital,<br />

Richmond, VA; 3 University of California, San Francisco, CA;<br />

4 California Pacific Medical Center, San Francisco, CA; 5 Evidera,<br />

Bethesda, MD; 6 Patient-Centered Outcomes Research Institute<br />

(PCORI), Washington, DC; 7 Horizon Therapeutics (formerly Hyperion<br />

Therapeutics, Inc.), Brisbane, CA<br />

The West Haven scale (WH), used for decades in clinical practice,<br />

was neither developed as an outcome assessment tool<br />

using methodology acceptable to regulatory authorities, nor<br />

is it consistently applied. Although recent ISHEN and EASL/<br />

AASLD Guidelines have clarified HE nomenclature and treatment,<br />

there is no standardized tool for assessing overt HE<br />

(OHE). A standardized clinician reported outcome (ClinRO)<br />

tool to diagnose and grade OHE was developed, in compliance<br />

with methodologies accepted by regulatory authorities,<br />

for use in clinical trials as well as daily clinical practice. Study<br />

Design: A two part iterative Delphi method was employed<br />

using a 3-member steering committee and a panel of 40<br />

practicing hepatologists in the US, Canada, Europe, South<br />

America, and Australia to develop a new ClinRO instrument<br />

(HE Grading Instrument, HEGI) for OHE identification and<br />

grading. Part 1 involved open-ended concept elicitation from<br />

the panel regarding clinical findings they use in their daily<br />

practice to diagnose and rate OHE severity. Part 2 included<br />

testing the HEGI for clarity, usability, and inter- and intra-rater<br />

reproducibility. To assess inter-rater reproducibility, 34 panelists<br />

viewed 5 videos depicting HE episodes of differing severity<br />

and determined the presence/absence of OHE, as well as its<br />

severity. Intra-rater reproducibility was assessed among a subset<br />

of the panelists who viewed the video vignettes a second<br />

time, in random order, and rated each episode again. Results:<br />

Of the 40 panelists, 97.5% had participated in clinical trials,<br />

52.5% saw > 10 HE patients/week, and 32% were from outside<br />

the US. Major OHE criteria recommended in Part 1 for<br />

inclusion in the HEGI included disorientation to time (90%),<br />

place (72.5%), and person (70.0%), asterixis (97.5%), lethargy<br />

(92.5%), and coma (100%). Assuming exclusion of other<br />

causes, consensus was reached that any of the following constituted<br />

a minimum requirement for diagnosis of OHE: (1) any<br />

disorientation, (2) presence of both lethargy and asterixis, or<br />

(3) coma. The overall adjudicated concordance in Part 2 for<br />

discriminating between presence/absence of OHE using HEGI<br />

was 97.1%; the overall concordance for rating episode severity<br />

was 87.1%; and the overall concordance for reproducibility<br />

was 98%. Conclusions: The HEGI, a ClinRO tool for diagnosis<br />

and grading of OHE episodes, was developed using the principles<br />

contained in the FDA Guidance for Development of Patient<br />

Reported Outcomes for instrument validity and reproducibility.<br />

It is anticipated to aid in the identification of OHE in clinical<br />

practice and enhance the reliability of endpoint assessment in<br />

clinical trials for OHE.<br />

Disclosures:<br />

Jasmohan S. Bajaj - Advisory Committees or Review Panels: Salix, Merz, otsuka,<br />

ocera, grifols, american college of gastroenterology; Grant/Research Support:<br />

salix, otsuka, grifols<br />

Nathan M. Bass - Advisory Committees or Review Panels: Quest Diagnostics<br />

R Todd Frederick - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, Salix, Vital Therapies, Hyperion, AbbVie<br />

Karin Coyne - Consulting: Evidera<br />

Dion F. Coakley - Employment: Horizon Therapeutics<br />

Masoud Mokhtarani - Employment: Hyperion; Stock Shareholder: Hyperion<br />

Marzena Jurek - Employment: Horizon Therapeutics, Inc.<br />

The following authors have nothing to disclose: Mary Kay Margolis, Bruce F.<br />

Scharschmidt<br />

1486<br />

Usefulness of Balloon-Occluded Retrograde Obliteration<br />

(B-RTO) as a Consolidation Procedure after Anticoagulation<br />

Therapy in Cirrhotic Patients with Portal Vein<br />

Thrombosis<br />

Mie Inao, Kazuki Hirahara, Kayoko Sugawara, Nobuaki<br />

Nakayama, Yukinori Imai, Satoshi Mochida; Department of Gastroenterology<br />

& Hepatology, Saitama Medical University, Moroyama-Machi,<br />

Japan<br />

Aim: Although anticoagulation therapies with Xa inhibitors and<br />

antithrombin concentrates were shown to be effective for attenuation<br />

of portal vein thrombosis in cirrhotic patients, aggravation<br />

or recurrence of the lesions may occur following the<br />

therapies leading to derangement of liver function. Decrease<br />

of blood flow in the portal vein as a consequence of porto-systemic<br />

shunts may responsible for thrombosis development.<br />

Thus, the usefulness of B-RTO as a consolidation procedure<br />

after anticoagulation therapies was evaluated. Methods: The<br />

subjects were 43 patients (23 men and 20 women, aged from<br />

40 to 76 years old) with liver cirrhosis complicating portal vein<br />

thrombosis. Both danaparoid Na (2,500 units/day) and antithrombin<br />

concentrates (1,500 units/day) were intravenously<br />

administrated for 3 days followed by danaparoid Na injections<br />

for further 11 days. Patients seen in April 2013 and later<br />

received B-RTO procedures after anticoagulation therapies,<br />

when porto-systemic shunts were observed on CT and/or MRI<br />

imaging. A balloon catheter was inserted into the shunts followed<br />

by injection of 5% ethanolamine oleate through the catheter<br />

under balloon inflation. The balloon was kept inflation for<br />

6 to 48 hours depending on sizes of the shunts. Results: Immediately<br />

after anticoagulation therapies, portal vein thrombosis<br />

was completely disappeared in 11 patients (25%) and the<br />

sizes of thrombosis were attenuated in 15 patients (35%), while<br />

the lesions did not change in 17 patients (40%). B-RTO was<br />

additionally done in 4 patients; 2 patients showing complete<br />

thrombosis disappearance and 2 patients failing to achieve<br />

thrombosis attenuation. Following B-RTO procedures, thrombosis<br />

did not recur in both of the former patients and the lesions<br />

disappeared in both of the latter patients despite that anticoagulation<br />

therapies were ineffective. In contrast, in 39 patients<br />

without additional B-RTO procedures, thrombosis recurred in<br />

4 among 9 patients after thrombosis disappearance and was<br />

aggravated in 6 among 15 patients achieving thrombosis<br />

attenuation. Conclusion: B-RTO was effective as a consolidation<br />

procedure after anticoagulation therapies for patients with<br />

portal vein thrombosis even in those failing to achieve attenuation<br />

of the lesions when porto-systemic shunts responsible for<br />

decrease of blood flows in the portal vein were observed.<br />

Disclosures:<br />

Satoshi Mochida - Grant/Research Support: Chugai, MSD, Tioray Medical,<br />

BMS; Speaking and Teaching: MSD, Toray Medical, BMS, Tanabe Mitsubishi<br />

The following authors have nothing to disclose: Mie Inao, Kazuki Hirahara,<br />

Kayoko Sugawara, Nobuaki Nakayama, Yukinori Imai


936A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1487<br />

Osteopontin is a new prognostic marker of liver cirrhosis<br />

Radan Bruha 1 , Marie Jachymova 2 , Jaromir Petrtyl 1 , Karel Dvorak 1 ,<br />

Martin Lenicek 2 , Petr Urbanek 3 , Libor Vitek 2 ; 1 4th Intenal Clinic,<br />

Charles University in Prague, 1st Faculty of Medicine, Prague,<br />

Czech Republic; 2 Institute of Clinical Biochemistry and Laboratory<br />

Diagnostics, Charles University in Prague, 1st Faculty of medicine,<br />

Prague, Czech Republic; 3 Internal Clinic, Central Military Hospital,<br />

Prague, Czech Republic<br />

Background and Aims: Portal hypertension leads to major<br />

complications of cirrhosis. Invasively measured hepatic venous<br />

pressure gradient (HVPG) is a strong prognostic indicator in<br />

patients with cirrhosis and portal hypertension. Recently, osteopontin<br />

has emerged as a new marker with a possible relation<br />

to fibrosis and cirrhosis and close correlation to portal hypertension.<br />

Our aim was to evaluate the relationship of osteopontin<br />

plasma concentrations to the prognosis of patients with<br />

cirrhosis. Methods: A cohort of 154 patients with liver cirrhosis<br />

(112 ethylic, 108 men, age 34-72 years) were enrolled. The<br />

HVPG was measured by standard catheterization using the<br />

balloon wedge technique. The mean follow-up of the patients<br />

was 3.7±2.6 years (maximal length 7 years). Osteopontin was<br />

measured by ELISA method. Results: The mean value of HVPG<br />

was 16.18±5.6 mmHg. Patients with osteopontin values above<br />

80 ng/ml had a significantly lower cumulative survival rate<br />

compared to those with osteopontin ≤80 ng/ml (37% vs. 56%,<br />

p=0.00035; OR 2.23, 95% CI 1.06-4.68; Fig 1). Similarly,<br />

the HVPG cut-off value of 10 mm Hg divided patients into 2<br />

groups with significantly different probabilities of cumulative<br />

survival (39% for those with HVPG>10 mm Hg compared to<br />

65% for those with HVPG≤10 mm Hg; p=0.0086, OR 2.92,<br />

95% CI 1.09-7.76). Mortality in patients with at least one risk<br />

factor (HVPG above 10 mm Hg or plasma OPN above 80<br />

ng/l) was more than twice as high compared to patients without<br />

any risk factors (OR=2.34); in those with both risk factors,<br />

mortality was more than five times as high (OR=5.1) compared<br />

to patients without any risk factors. Conclusions: Osteopontin is<br />

a strong prognostic factor for overall survival in patients with<br />

cirrhosis. Supported by SVV 260 156/2015.<br />

Figure 1: Cumulative proportion of surviving patients with plasma<br />

OPN levels below and above 80 ng/ml using the Kaplan-Meier<br />

method.<br />

1488<br />

Liver Injury Independently Predicts Incident Cardiovascular<br />

Disease (CVD) Risk<br />

Kaku So-Armah 1 , Janet Tate 3,2 , Joseph K. Lim 2 , Vincent Lo Re 14 ,<br />

Vincent Marconi 4 , Jeffrey Samet 1 , Cynthia Gibert 8 , David Leaf 9,10 ,<br />

Adeel Butt 6,7 , Matthew B. Goetz 9,10 , David Rimland 4,11 , Maria C.<br />

Rodriguez-Barradas 12,13 , Matthew Budoff 5 , Chung-Chou Chang 15 ,<br />

Amy Justice 2,3 , Matthew Freiberg 16 ; 1 Boston University School of<br />

Medicine, Boston, MA; 2 Yale University School of Medicine, New<br />

Haven, CT; 3 VA Connecticut Healthcare System, West Haven,<br />

CT; 4 Emory University School of Medicine, Atlanta, GA; 5 LA<br />

Biomed Research Institute, Torrance, CA; 6 VA Pittsburgh Healthcare<br />

System, Pittsburgh, PA; 7 Hamad Healthcare Quality Institute<br />

and Hamad Medical Corporation, Doha, Qatar; 8 Washington DC<br />

VA Medical Center, Washington, DC; 9 David Geffen School of<br />

Medicine at UCLA, Los Angeles, CA; 10 VA Greater Los Angeles<br />

Healthcare System, Los Angeles, CA; 11 Atlanta VA Medical Center,<br />

Atlanta, GA; 12 Michael E. DeBakey VA Medical Center, Houston,<br />

TX; 13 Baylor College of Medicine, Houston, TX; 14 University of<br />

Pennsylvania School of Medicine, Philadelphia, PA; 15 University of<br />

Pittsburgh Medical Center, Pittsburgh, PA; 16 Vanderbilt University<br />

School of Medicine, Nashville, TN<br />

Background: Multiple etiologies of liver injury are associated<br />

with cardiovascular disease (CVD) e.g., hepatitis C or B (HCV,<br />

HBV) and HIV co-infection, alcohol, and obesity. We hypothesize<br />

that liver injury is a mechanism of increased CVD risk.<br />

We assessed whether liver fibrosis index 4 (FIB4) predicted<br />

incident CVD. Liver-related comorbidities may have extra-hepatic<br />

contribution to CVD risk. Thus, we performed sensitivity<br />

analyses excluding people with HCV, HBV, HIV, alcohol use<br />

disorder, and obesity. Methods: Participants from the Veterans<br />

Aging Cohort Study Virtual Cohort (VACS VC), without CVD<br />

at baseline (first clinical visit after 4/1/2003) were included.<br />

Liver injury was categorized using baseline FIB4 (calculated<br />

using age, liver transaminases and platelets) and ICD-9 codes<br />

for cirrhosis and hepatic decompensation. Total non-fatal and<br />

fatal CVD was assessed using VA, VA Fee For Service and<br />

Medicare ICD-9 and procedure codes for myocardial infarction,<br />

heart failure, coronary heart disease and stroke, and<br />

National Death Index cause of death data. Follow-up ended<br />

after a CVD event, or death, or on 9/30/2012. Cox regression<br />

analyses were used to estimate CVD risk adjusting for<br />

confounders including renal disease. Results: After excluding<br />

21488 people with prevalent CVD, 104731 people remained.<br />

Mean (SD) age was 49.6 (9.5) years in this largely male (90%)<br />

cohort followed for a median (IQR) of 7 (3, 9) years. LDL cholesterol<br />

and body mass index decreased with increasing FIB4.<br />

African American race, HIV, HCV, HBV, current smoking, type<br />

2 diabetes, alcohol use disorder, cocaine use, and anemia<br />

were more common among those with FIB4>1.45 than those<br />

with FIB43.25 or a clinical diagnosis of cirrhosis or<br />

hepatic decompensation had significantly elevated CVD risk<br />

despite also having higher mortality rates (see Table below).<br />

This association persisted among those without heart failure or<br />

without HCV, HBV, HIV, alcohol abuse/dependence diagnosis<br />

and BMI>30 kg/m 2 . Conclusions: FIB4 and severe liver injury<br />

independently predict risk of CVD in VACS VC. Future <strong>studies</strong><br />

should assess if treatment of liver injury reduces CVD risk.<br />

Disclosures:<br />

The following authors have nothing to disclose: Radan Bruha, Marie Jachymova,<br />

Jaromir Petrtyl, Karel Dvorak, Martin Lenicek, Petr Urbanek, Libor Vitek


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 937A<br />

GPC in culture-positive SBP and/or bacteremia were observed.<br />

Patients with GNB had more advanced liver disease and more<br />

severe inflammatory response compared to GPC ones. A variety<br />

of drug-resistant microorganisms and a high rate of E. Faecium<br />

have emerged. Due to high rates of resistance in currently<br />

recommended therapy and prophylaxis, the choice of optimal<br />

antibiotic therapy should be individualized.<br />

Disclosures:<br />

The following authors have nothing to disclose: Alexandra Alexopoulou, Larisa E.<br />

Vasilieva, Sotiria Papadaki, Danai Agiasotelli, Sophia Pouriki, Athanasia Tsiriga,<br />

Marina G. Toutouza, Spyros P. Dourakis<br />

Disclosures:<br />

Joseph K. Lim - Consulting: Merck, Boehringer-Ingelheim, Gilead, Bristol Myers<br />

Squibb, Janssen; Grant/Research Support: AbbVie, Boehringer-Ingelheim, Bristol<br />

Myers Squibb, Gilead, Hologic, Janssen<br />

Matthew Budoff - Grant/Research Support: GE, NIH<br />

The following authors have nothing to disclose: Kaku So-Armah, Janet Tate, Vincent<br />

Lo Re, Vincent Marconi, Jeffrey Samet, Cynthia Gibert, David Leaf, Adeel<br />

Butt, Matthew B. Goetz, David Rimland, Maria C. Rodriguez-Barradas, Chung-<br />

Chou Chang, Amy Justice, Matthew Freiberg<br />

1489<br />

Increasing frequency of gram-negative multi-drug resistant<br />

bacteria and Enterococcus faecium in spontaneous<br />

bacterial peritonitis and bacteremia in patients with<br />

decompensated cirrhosis<br />

Alexandra Alexopoulou 1 , Larisa E. Vasilieva 1 , Sotiria Papadaki 1 ,<br />

Danai Agiasotelli 1 , Sophia Pouriki 1 , Athanasia Tsiriga 2 , Marina<br />

G. Toutouza 2 , Spyros P. Dourakis 1 ; 1 2nd Dept of Medicine, Medical<br />

School University of Athens, Athens, Greece; 2 Microbiology<br />

Department Hippokration GNA, Athens, Greece<br />

Purpose/Background: Spontaneous bacterial peritonitis (SBP)<br />

and bacteremia in cirrhotic patients are historically caused<br />

by Gram-negative bacteria (GNB) almost exclusively Enterobacteriaceae.<br />

Recently, an increasing rate of infections with<br />

Gram-positive cocci (GPC) and multi-drug resistant (MDR)<br />

microorganisms was demonstrated. The purpose of this study<br />

was to assess possible recent changes of the bacteria causing<br />

SBP or bacteremia in patients with decompensated cirrhosis.<br />

Methods: We prospectively recorded 130 cases (68.5%<br />

males) with culture-positive SBP or spontaneous bacteremia<br />

without SBP during a 3-year-period (2012-2014). Results: The<br />

76.2% of patients had health care-associated (HCA) or nosocomial<br />

infections. GPC were found in half of the cases. The most<br />

prevalent organisms in a descending order were E. coli (33),<br />

Enterococcus spp (30, including 17 E. Faecium), Streptococcus<br />

spp (25), K. pneumonia (16), S. aureus (8), P. aeruginosa<br />

(5), other GNB (11) and anaerobes (2). Twenty five (19.2%)<br />

of the isolated bacteria were MDR, including extended-spectrum-beta-lactamase-producing<br />

(ESBL)-GNB (9), carbapenemase-producing<br />

(KPC)-K. pneumonia (5), P. aeruginosa (5), A.<br />

baumannii (2), E. Faecium Vancomycin-resistant (2), KPC colistin-Resistant-K.<br />

pneumonia (1) and metallo-Β-Lactamase producing<br />

(MBL)-E. Coli (1). MDR bacteria were more frequently<br />

isolated in HCA/nosocomial than in community-acquired infections<br />

(100% vs 70.5%) (P=0.002). E. Faecium was always<br />

associated with HCA/nosocomial infections. Patients with MDR<br />

bacteria or E. Faecium had longer hospitalization compared<br />

to the rest of patients (P=0.013). Patients with GNB had more<br />

advanced liver disease and higher neutrophil-to-leucocyte ratio<br />

compared to those with GPC [MELD 21 (17-26) vs 18 (13-<br />

25) respectively, (P=0.059), total bilirubin 4.98 (2,2-11,1) vs<br />

2.75 (1.79-5.90) respectively, (P=0.046] and [84 (74-91) vs<br />

78 (69-85) respectively, (P=0.018)]. Third-generation cephalosporin-resistance<br />

was observed in 43.1% and quinolone-resistance<br />

in 49.6%. Conclusion: Similar frequencies of GNB and<br />

1490<br />

Non-invasive assessment of portal hypertensive gastropathy:<br />

feasibility of contrast-enhanced ultrasonography<br />

for stomach wall<br />

Hitoshi Maruyama, Soichiro Kiyono, Takayuki Kondo, Tadashi<br />

Sekimoto, Osamu Yokosuka; Department of Gastroenterology<br />

and Nephrology, Chiba University Graduate School of Medicine,<br />

Chiba, Japan<br />

Background/Aim: Portal hypertensive gastropathy (PHG) is one<br />

of the major complications in cirrhosis, and endoscopy may be<br />

the only diagnostic tool. The analysis of microbubble-induced<br />

enhancement of the stomach wall on the sonogram may detect<br />

the hemodynamic difference between patients with and without<br />

PHG. The aim was to elucidate the efficacy of transabdominal<br />

contrast-enhanced ultrasound (CEUS) as a non-invasive<br />

diagnostic tool of PHG. Methods: This is a prospective study<br />

performed in 54 subjects, 40 cirrhosis patients (60.4 ± 10.5<br />

years; 11 females; Child A 25, B 15) with esophageal varices<br />

who underwent the measurement of hepatic venous pressure<br />

gradient (HVPG) and 14 controls (41.4 ± 13.1 years; female<br />

4). Contrast effects using perflubutane microbubble agent<br />

were assessed at the upper stomach wall by CEUS. The peak<br />

enhancement time (PT, time from contrast onset to maximum<br />

enhancement) and the intensity ratio between pre-enhancement<br />

and peak-enhancement (IR) were calculated using time-intensity<br />

curve. Results: PHG was detected in 15 cirrhosis patients<br />

(37.5 %; mild 12, severe 3). The PT and IR were 8.5s and<br />

8.3s (p=0.348), and 1.22 and 1.56 (p


938A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Hitoshi Maruyama, Soichiro<br />

Kiyono, Takayuki Kondo, Tadashi Sekimoto, Osamu Yokosuka<br />

1491<br />

Videogame Training Improves Brain Plasticity In Covert<br />

Hepatic Encephalopathy Through Changes In Brain Connectivity<br />

Vishwadeep Ahluwalia, James Wade, Frederick G. Moeller, Joel<br />

L. Steinberg, Michael Lennon, Ariel Unser, Melanie White, Douglas<br />

M. Heuman, Edith A. Gavis, Jasmohan S. Bajaj; VCU and<br />

McGuire VAMC, Richmond, VA<br />

Cognitive improvement using videogames could be a non-pharmacologic<br />

covert HE(CHE) therapy but its effect on brain networks<br />

is unclear. Aim: Assess the impact of videogame-related<br />

cognitive therapy on brain networks in CHE. Methods: CHE<br />

cirrhotics underwent an 8-wk trial with 4 visits with CHE-related<br />

[Paper/pencil & inhibitory control test (ICT)] & unrelated<br />

tests (Hopkins Verbal learning test HVLT). Visit 1 & 2 were 2<br />

wks apart to gauge learning.At visit 2 an iPod loaded with 2<br />

games; IQ Boost & Arcade Challenge was given for 4 wks. Visit<br />

3 was 4 wks post-iPod when cognitive testing was repeated &<br />

trial ended. A subgroup (n=13) underwent MRI at visit 2 & 3<br />

to define functional/structural connectivity change. Changes in<br />

brain MRI & tests related/unrelated to videogames were analyzed.Resting<br />

state analysis produces functional brain network<br />

maps linked with attention, memory &other domains. Tractography<br />

estimates the spatial probability distribution of white<br />

matter tracts between two regions;it was performed between<br />

memory-related (HVLT-related) regions such as hippocampus,<br />

angular (AG) & anterior middle temporal gyri (AMTG). Pre-/<br />

post-ipod differences between networks were tested after<br />

region of interest creation. Results: 20 CHE pts (12 men, 56<br />

yrs, MELD 10) were recruited. There was a significant improvement<br />

in paper-pencil tests between visits 1 & 2. ICT & HVLT<br />

improved only during the iPod phase. 27% of pts’ performance<br />

normalized at visit 4 (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 939A<br />

Disclosures:<br />

The following authors have nothing to disclose: Tsuyoshi Ishikawa, Ryo Sasaki,<br />

Tatsuro Nishimura, Yuki Aibe, Shogo Shiratsuki, Kazuhito Matsunaga, Takuya<br />

Iwamoto, Taro Takami, Isao Sakaida<br />

1493<br />

Inadequate Food And Water Intake In Patients With Cirrhosis:<br />

Need Of A Personalized Nutritional Approach<br />

Ferdinando A. Giannone 1,3 , Marco Domenicali 1,3 , Maurizio Baldassarre<br />

1,3 , Paola De Pasquale 1 , Silvia Boffelli 1,2 , Maristella Laggetta<br />

1,3 , Maurizio Biselli 1,2 , Mauro Bernardi 1,2 , Paolo Caraceni 1,3 ;<br />

1 Department of Surgical and Medical Sciences, Alma Mater Studiorum<br />

University of Bologna, Bologna, Italy; 2 U.O. Semeiotica<br />

Medica, Azienda Ospedaliero-Universitaria di Bologna - Policlinico<br />

S. Orsola-Malpighi, Bologna, Italy; 3 Center for Applied<br />

Biomedical Research (CRBA), Alma Mater Studiorum University of<br />

Bologna, Bologna, Italy<br />

Background/Aim. Malnutrition is a frequent complication in<br />

patients with cirrhosis, contributing to excess morbidity and<br />

mortality. Although practical guidelines have been published,<br />

its management at the individual level is quite complex since<br />

cirrhotic patients constitute a very heterogeneous population in<br />

terms of needs, intake and utilization of nutrients as well as motivation<br />

and adherence to dietary programs. This study aimed to<br />

assess whether the food and water intake of cirrhotic patients<br />

is adequate to the needs of their specific clinical condition<br />

according to the recommended nutritional indications. Methods.<br />

Outpatients with cirrhosis were consecutively screened<br />

for inadequate nutrition by using a scored algorithm, including<br />

as risk factors altered BMI, recent body weight changes,<br />

presence of diabetes, ascites or hepatic encephalopathy (HE).<br />

All patients at risk were offered to be evaluated by a certified<br />

nutritionist. Patients were then asked to fill up a weekly food<br />

diary, which provided information on the amount and origin<br />

of the food, water intake, and, physical activity, and to perform<br />

a hand-grip strength (HGS) test to assess muscle wasting.<br />

Results. Of the 144 pts screened, only 16% showed no risk of<br />

inadequate nutrition, while 37% presented 1 risk factor and<br />

47% ≥2 risk factors. Child-Pugh and MELD scores were 5.5<br />

and 7.4 in the group with no risk, 6.8 and 11.7 in pts with<br />

one risk factor, and 7.6 and 12.8 in those with ≥2 risk factors.<br />

Of the 121 pts at risk, 86 pts accepted the nutritional evaluation:<br />

reduced muscle strength was found in 56% of pts, while<br />

65% properly returned the weekly food diary. Interestingly,<br />

the adherence was higher in those with ≥2 risk factors and/or<br />

more advanced disease, particularly if transplant candidates.<br />

Despite all patients had received general nutritional indications<br />

by their hepatologists, the analysis of the diary showed that<br />

70% of pts had an inadequate intake of calories, 41% of water<br />

and 36% of salt. Daily and/or weekly distribution of macronutrients<br />

was also inadequate in the vast majority: carbohydrates<br />

in 88%, sugar in 73%, proteins in 86%, lipids in 36 %, and<br />

fibers in 95%. No physical activity was present in 79% of pts.<br />

Conclusions. This study shows that the vast majority of cirrhotic<br />

patients have an inadequate intake of food, water and macronutrients,<br />

which was often associated to muscle wasting and<br />

often related to a scarce knowledge of nutritional facts. This<br />

highlights the need of a personalized nutritional approach,<br />

including educational and motivational aspects, which takes<br />

into account the complex interacting, often conflicting, nutritional<br />

needs of each individual patient.<br />

Disclosures:<br />

Paolo Caraceni - Advisory Committees or Review Panels: GSK; Consulting: BMS;<br />

Grant/Research Support: Grifols; Speaking and Teaching: Baxter, Kedrion<br />

The following authors have nothing to disclose: Ferdinando A. Giannone, Marco<br />

Domenicali, Maurizio Baldassarre, Paola De Pasquale, Silvia Boffelli, Maristella<br />

Laggetta, Maurizio Biselli, Mauro Bernardi<br />

1494<br />

Myocardial strain as a noninvasive measure of circulatory<br />

dysfunction and systolic function in liver cirrhosis.<br />

Carolina F. Pimentel 1,4 , Maria Lucia Ferraz 1 , Miguel Osman D.<br />

Aguiar 2 , Adriano M. Gonzalez 1,2 , Gabriel D. Branco 2 , Marcel<br />

Superbia 2 , Ana Cristina A. Feldner 1 , Michelle Lai 5 , Wilson<br />

Mathias 3 , Mario Kondo 1 ; 1 Gastroenterology, Federal University of<br />

Sao Paulo, Sao Paulo, Brazil; 2 Echocardiology, Sao Luiz Hospital,<br />

Sao Paulo, Brazil; 3 Heart Institute, University of Sao Paulo, Sao<br />

Paulo, Brazil; 4 Liver Transplant, Euryclides de J Zerbini Transplant<br />

Hospital, Sao Paulo, Brazil; 5 Gastroenterology and Hepatology,<br />

Beth Israel Deaconess Medical Center, Harvard Medical School,<br />

Boston, MA<br />

Cirrhosis is related to hyperdynamic circulation with insufficient<br />

organ perfusion, activation of RAAS and compensatory<br />

increase in cardiac output. Left ventricular longitudinal myocardial<br />

strain is a non-invasive echocardiographic marker of<br />

myocardial contractility, able to identify subclinical systolic dysfunction<br />

(SD). We evaluated cirrhotic patients and measured<br />

strain rate, parameters for cardiac dysfunction and serum renin<br />

levels and analyzed them by Child Classes. Method: A cohort<br />

of 103 cirrhotic outpatients was studied. SD was defined as<br />

strain>-18% or ejection fraction (EF)< 55%. and diastolic dysfunction<br />

(DD) findings as E/A8. Results: Non-alcoholic<br />

liver cirrhosis was the most common etiology (70%).<br />

Child stages were A (66%), B (21%) and C(13%). Alcohol<br />

consumption was unrelated to any parameters. SD was more<br />

frequently diagnosed by strain rate than EF (5% vs. 0.9%).<br />

All patients had normal ventricular measurements and 24%<br />

evidence of DD. These patients had strain rates consistent with<br />

decreased myocardial contractility, even without anatomical<br />

abnormalities. Both strain rate and serum renin levels were<br />

significantly correlated with Child scores (p=0.002, r=-0.3;<br />

p=0.03, r=0.3, respectively), showing increased myocardial<br />

contractility and progressive RAAS activation with increasing<br />

Child scores. Mean strain rates and mean serum renin levels<br />

were statistically different among Child classes (see figure).<br />

Conclusions: Myocardial strain rate and serum renin level correlated<br />

with Child stages consistent with a progressively more<br />

hyperdynamic circulatory state with worsening liver decompensation.<br />

Strain rate was more sensitive than ejection fraction in<br />

diagnosing SD and is a promising tool for early diagnosis of<br />

subclinical systolic dysfunction and circulatory dysfunction in<br />

cirrhosis.


940A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

this cohort, with 64% of the cohort having at least one highrisk<br />

parameter. Table 1 shows prevalence of high risk HRV<br />

parameters. SDNN and SDANN correlated with Child scores<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 941A<br />

and dry BMI, mid-arm muscle circumference, SMI and VMI.<br />

Logistic regression analysis revealed that INR [O.R. 6.581<br />

(95% C.I. 1.330 – 32.569), p = 0.021] and log-transformed<br />

leptin [O.R. 0.425 (0.221 – 0.816), p < 0.010] were independently<br />

associated with malnutrition. Conclusions: Low serum<br />

leptin and elevated INR are associated with malnutrition in hospitalized<br />

patients with cirrhosis and may represent biomarkers<br />

for early identification of malnutrition in this population.<br />

Disclosures:<br />

The following authors have nothing to disclose: Vikrant Rachakonda, Jaideep<br />

Behari<br />

Disclosures:<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Thoetchai Peeraphatdit, Niyada<br />

Naksuk, Roongruedee Chaiteerakij<br />

1497<br />

The World Congress of Gastroenterology Guidelines for<br />

the Diagnosis of Left Ventricular Diastolic Dysfunction<br />

Differ from the American Society of Echocardiography<br />

Standard Recommendation<br />

Thoetchai Peeraphatdit 1,2 , Niyada Naksuk 3 , Lewis R. Roberts 2 ,<br />

Roongruedee Chaiteerakij 2,4 ; 1 Internal Medicine, University of<br />

Minnesota, Minneapolis, MN; 2 Gastroenterology and Hepatology,<br />

Mayo Clinic College of Medicine, Rochester, MN; 3 Cardiovascular<br />

Diseases, Mayo Clinic College of Medicine, Rochester,<br />

MN; 4 Medicine, Chulalongkorn University, Bangkok, Thailand<br />

Purpose: Diagnostic criteria for left ventricular diastolic dysfunction<br />

(LVDD) proposed by the World Congress of Gastroenterology<br />

(WCG) has never been tested and are different from<br />

the American Society of Echocardiography (ASE) criteria. This<br />

study determined the agreement of LVDD diagnosed by these<br />

two guidelines, the performance of the WCG criteria, and<br />

their prognostic values. Methods: We retrospectively enrolled<br />

914 consecutive patients with normal left ventricular ejection<br />

fraction (LVEF) and available data on diastolic function undergoing<br />

liver transplant (LT) at a tertiary medical center between<br />

1994 and 2014. According to the ASE, LVDD was diagnosed<br />

based on a stepwise approach using multiple parameters.<br />

However, by the WCG, LVDD was defined as an early to<br />

late ventricular filling velocities ratio (E/A) 200 ms. The agreement of LVDD diagnosed by<br />

the two guidelines was analyzed using the Kappa statistic.<br />

The performance of the WCG criteria was determined using<br />

the ASE criteria as a gold standard. Results: Mean age was<br />

53.7±10.6 years; 65.2% were male. During 5.6±4.2 years<br />

follow up after LT, 21.6% patients had died. Overall LVEF was<br />

65.8±5.1%. By the ASE criteria, 47.2% of patients had LVDD;<br />

whereas the prevalence of LVDD was 19.6%, 61.0%, 64.5%<br />

and 13.9% by the WCG criteria of E/A 200 ms,<br />

either one of the criteria, or both criteria, respectively. There<br />

was minimal agreement between LVDD diagnosed by the 2<br />

WCG criteria (Kappa=0.10). Moreover, there was generally<br />

poor or slight agreement between the ASE and the WCG criteria<br />

(Kappa=0.08-0.18, Table). The WCG criteria had sensitivities<br />

of 50-70% and specificities of 58-62% (Table). However,<br />

in the multivariate Cox regression, only LVDD diagnosed by<br />

the WCG criterion of E/A


942A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

in sickle hepatopathy. Despite diagnostic challenges with HB<br />

in SCD, DB, ALP and PLT appear to be respectively associated<br />

with increasing portal pressures and mortality.<br />

Disclosures:<br />

The following authors have nothing to disclose: David M. Chascsa, Ohad Etzion,<br />

Hawwa Alao, Courtney Fitzhugh, John F. Tisdale, Caterina Minniti, Matthew<br />

Hsieh, David E. Kleiner, Ahmed M. Gharib, Varun K. Takyar, Jason L. Eccleston,<br />

Theo Heller, Christopher Koh<br />

1499<br />

A half of hepatic hydrothorax in cirrhotic patients were<br />

caused by porous diaphragm syndrome due to radiofrequency<br />

ablation for hepatocellular carcinoma<br />

Kenichi Miyoshi, Masahiko Koda, Toshiaki Okamoto, Tomomitsu<br />

Matono, Takaaki Sugihara, Keiko Hosho, Jun-ichi Okano; Tottori<br />

University School of Medicine, Yonago, Japan<br />

Aims Hepatic hydrothorax is defined as a significant pleural<br />

effusion in cirrhotic patients without underlying pulmonary or<br />

cardiac disease. One of causes of refractory hepatic hydrothorax<br />

is porous diaphragm syndrome. However, the frequency<br />

of porous diaphragm syndrome in hepatic hydrothorax and<br />

the cause of porous diaphragm syndrome are unknown. We<br />

aimed to evaluate its frequency, clinical background, and therapeutic<br />

value of them. Methods From August 2005 to April<br />

2015, 124 cirrhotic patients hospitalized due to hepatic<br />

ascites are enrolled. Patient’s data hospitalized several times<br />

during its period were referred to the most serious cases. We<br />

compared clinical background, and therapeutic value between<br />

patients with and without hepatic hydrothorax. Four patients<br />

with the rupture of hepatocellular carcinoma (HCC), 3 hemodialysis<br />

patients, and 4 patients taking warfarin were excluded.<br />

In 17 of 115 patients from September 2007 with hydrothorax,<br />

we examined the presence of diaphragmatic defects with<br />

intraperitoneal injection of perflubutane (approved by the local<br />

ethics committee). We analyzed clinical background between<br />

patients with and without porous diaphragm syndrome. Results<br />

Forty-one patients (34.2%) in 124 cirrhotic patients with ascites<br />

had hepatic hydrothorax. There was no difference between<br />

patients with and without hydrothorax on clinical background<br />

and therapeutic value. Ten (58.8%) of 17 patients with hydrothorax<br />

had porous diaphragm syndrome. Accordingly, 10<br />

(8.7%) of 115 cirrhotic patients with ascites complicated diaphragmatic<br />

defects. Patients with porous diaphragm syndrome<br />

had higher serum albumin levels and fewer hepatic encephalopathy<br />

events, and more past history of radiofrequency ablation<br />

(RFA) for HCC adjacent to diaphragm than those without<br />

porous diaphragm syndrome. Multivariate analysis showed<br />

that only RFA was an independent factor for porous diaphragm<br />

syndrome. Tour of 10 patients with porous diaphragm syndrome<br />

underwent thoracoscopic diaphragm plication. In all<br />

these patients, pleural effusion disappeared or decreased.<br />

Conclusions 58.8% of patients with hepatic hydrothorax were<br />

caused by porous diaphragm syndrome. Porous diaphragm<br />

syndrome was related to RFA for HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kenichi Miyoshi, Masahiko<br />

Koda, Toshiaki Okamoto, Tomomitsu Matono, Takaaki Sugihara, Keiko Hosho,<br />

Jun-ichi Okano<br />

1500<br />

Sarcopenic obesity impairs prognosis of patients with<br />

cirrhosis<br />

Motoh Iwasa, Ryosuke Sugimoto, Yoshinao Kobayashi, Yoshiyuki<br />

Takei; Department of Gastroenterology and Hepatology, Mie University<br />

Graduate School of Medicine, Tsu, Japan<br />

Aims: It has been reported that there is a high incidence of serious<br />

events and poor outcomes when sarcopenia accompanies<br />

cirrhosis. In terms of body fat, it was recently reported that the<br />

prognosis is poor after resection of hepatocellular carcinoma<br />

in cirrhosis patients with decreased visceral fat. In the present<br />

study, skeletal muscle mass and visceral fat mass of patients<br />

with cirrhosis were measured, and the relationships between<br />

these body composition measurements and prognosis were<br />

investigated. Furthermore, time-dependent changes in body<br />

composition measurements were also examined. Methods: A<br />

total of 161 patients with cirrhosis (94 men, 67 women; 67 ±<br />

9 years) who received medical care and nutritional assessment<br />

were investigated. For body composition, in addition to height<br />

and weight, visceral fat area (VFA, cm 2 ) and upper limb skeletal<br />

muscle mass (kg) were measured using a multi-frequency<br />

bioelectrical impedance analysis device. Subsequently, setting<br />

the cut-off value of VFA at 100 cm 2 for visceral obesity and the<br />

cut-off value of sarcopenia at 1.7 kg/m 2 for men and 1.2 kg/<br />

m 2 for women, patients were classified, and their outcomes<br />

were observed. Furthermore, 60 patients with cirrhosis who<br />

underwent multiple body composition measurements during<br />

the observation period were extracted, and their initial and<br />

final measurements were compared. Results: Although there<br />

was a weak positive correlation between VFA and albumin,<br />

significant correlations were not observed between muscle<br />

mass and blood test results. Patients were classified into the<br />

following 4 groups: normal body composition; visceral obesity;<br />

sarcopenia; and sarcopenic obesity. There were 60, 61, 25,<br />

and 15 patients, respectively, in each group. Sarcopenia and<br />

sarcopenic obesity were more common in men, and sarcopenic<br />

obesity was more common in elderly individuals. In the comparison<br />

of 3 groups (visceral obesity, sarcopenia, and sarcopenic<br />

obesity groups), 22 (36%), 12 (48%), and 10 (67%) deaths<br />

were noted, respectively, during a mean observation period<br />

of 1005 days, and the prognosis was significantly worse in<br />

sarcopenic obesity, followed by sarcopenia and visceral obesity,<br />

in that order (P < 0.05). A significant positive correlation<br />

was observed only between percent change in skeletal muscle<br />

mass and percent change in serum albumin. Sarcopenia is a<br />

common feature of advanced cirrhosis, and transitions were<br />

observed from normal body composition to sarcopenia and<br />

from obese to sarcopenic obesity. Conclusions: Body composition<br />

is a prognostic factor for cirrhosis, and a better body<br />

composition may be advantageous for long-term survival in<br />

patients with cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Motoh Iwasa, Ryosuke Sugimoto,<br />

Yoshinao Kobayashi, Yoshiyuki Takei


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 943A<br />

1501<br />

Portal vein thrombosis, mortality and hepatic decompensation<br />

in patients with cirrhosis: A meta-analysis<br />

Jonathan G. Stine 1 , Puja M. Shah 2 , Scott L. Cornella 3 , Sean R.<br />

Rudnick 1 , George R. Stukenborg 4 , Patrick Northup 1 ; 1 Gastroenterology<br />

& Hepatology, University of Virginia, Charlottesville, VA;<br />

2 Surgery, University of Virginia, Charlottesville, VA; 3 Medicine,<br />

University of Virginia, Charlottesville, VA; 4 Public Health Sciences,<br />

University of Virginia, Charlottesville, VA<br />

Background: Non-neoplastic portal vein thrombosis is a common<br />

complication of cirrhosis. Treatment options can be risky<br />

and are not without complications. To make the most appropriate<br />

choice, clinicians need to know what the evidence<br />

says about the clinical importance of portal vein thrombosis<br />

in patients with cirrhosis. Objectives: To determine the clinical<br />

impact of portal vein thrombosis in terms of both mortality<br />

and hepatic decompensation (variceal hemorrhage, ascites,<br />

portosystemic encephalopathy) in adult patients with cirrhosis.<br />

Methods: We identified observational research <strong>studies</strong><br />

through February 2015 from MEDLINE, EMBASE, Scopus, Science<br />

Citation Index, AMED, Google Scholar, the Cochrane<br />

Library and the relevant grey literature. Two independent<br />

reviewers screened citations and extracted data. We specifically<br />

excluded any <strong>studies</strong> involving liver transplantation. We<br />

graded the strength of evidence and determined the magnitude<br />

of effect by calculating DerSimonian and Laird random effects<br />

odds ratios to obtain aggregate estimates of effect size and<br />

95% confidence intervals. Between-study variability and heterogeneity<br />

were assessed. Main results: After reviewing 226<br />

citations, we included 3 <strong>studies</strong> with 2,436 participants. The<br />

majority of included patients were Caucasian. Viral hepatitis<br />

was the most common etiology of cirrhosis (42%) and alcoholic<br />

liver disease was the second most prevalent (21%). Median<br />

MELD scores ranged from 10-15 for each individual study. The<br />

prevalence of portal vein thrombosis in aggregate was 7.2%.<br />

Patients with portal vein thrombosis had an increased risk of<br />

mortality (OR 1.69, 95% CI 1.18-2.42, p


944A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1503<br />

Real Time Pressure Volume Loops in Cirrhosis: Characterization<br />

of Systolic and Diastolic Function and Validation<br />

of Doppler Indices with the Gold Standard<br />

Cristina Ripoll 2,1 , Raquel Yotti 3 , Yolanda Benito 3 , Diego Rincón 2 ,<br />

Maria Vega Catalina 2 , Jaime Elizaga 3 , Javier Bermejo 3 , Rafael<br />

Bañares 2 ; 1 Innere Medizin I, Martin-Luther-Universität Halle-Wittenberg,<br />

Halle (Saale), Germany; 2 Digestive Diseases, Hospital Universitario<br />

Gregorio Marañón, IiSGM. CiberEHD, Madrid, Spain;<br />

3 Cardiology, Hospital General Universitario Gregorio Marañón,<br />

IiSGM, Madrid, Spain<br />

Studies reporting abnormal left ventricular (LV) properties in<br />

cirrhosis have frequently used load-dependent indices. Aims:<br />

1) to evaluate systolic and diastolic LV properties with the gold<br />

standard, LV conductance catheter, which allows obtention<br />

of real-time pressure volume loops, 2) to test the validity of<br />

Doppler-echocardiography indices compared to the gold standard,<br />

3) to analyze the impact of severity of liver disease,<br />

neurohormonal activation and beta-blocker (BB) treatment on<br />

LV properties. Methods. Pressure-volume loops were obtained<br />

in 9 patients with cirrhosis (Child A:3;B:2;C:4; refractory ascites:2)<br />

in the context of liver transplant evaluation and 9 controls<br />

undergoing LV catheterization. Invasive gold-standard systolic<br />

(maximal elastance, Emax) and diastolic indexes (relaxation<br />

and stiffness) were correlated to Doppler-echocardiography<br />

indices including peak ejection intraventricular pressure difference<br />

(EIVPD) and strain rate (SR). Gold-standard validated<br />

Doppler indexes in cirrhosis (n=59; Child A:15;B:25;C:19;<br />

refractory ascites:9) were compared to matched controls.<br />

The influence of the severity of liver disease, neurohormonal<br />

activation and BB was evaluated. Nonparametric tests were<br />

used. IRB approval was obtained. Results: Emax correlated<br />

only with EIPVD (r=0.75, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 945A<br />

1505<br />

Comparison Of Thrombin Generation After Addition Of<br />

Thrombomodulin Or Activated Protein C: Looking For A<br />

Marker Of Hypercoagulability With Low Variability In<br />

Patients With Cirrhosis<br />

Cédric Duron 1 , Armando Abergel 1 , Géraldine Lamblin 1 , Aurelien<br />

Lebreton 2 , Thomas Sinegre 3 ; 1 Hepatogastroenterolgy, CHU, Clermont<br />

Ferrand, France; 2 CHU Clermont Ferrand, Clermont ferrand,<br />

France; 3 CHU Clermont Ferrand, Clermont Ferrand, France<br />

Background and Aims: Cirrhosis is characterized by a defective<br />

hemostasis, and may result in bleeding complications as<br />

well as in thrombotic events. Conventional coagulation tests<br />

such as the Prothrombin Time and aPTT are prolonged in<br />

patients with cirrhosis, assessing only the defects of procoagulants.<br />

Thrombin Generation (TG) testing consist in a dynamic<br />

study of the coagulation, measuring generation and inhibition<br />

of the thrombin by pro and anticoagulant factors. This new<br />

approach reveals that TG could be equal or higher in patients<br />

with cirrhosis than in healthy volunteers. The aim of this study<br />

was: 1) to assess the variability of TG over time, and 2) the<br />

effect of Child-Pugh on TG. Methods: Fifty two patients with a<br />

diagnosis of cirrhosis, without hepatocellular carcinoma, were<br />

included in the study, in 2 groups: uncomplicated cirrhosis and<br />

complicated disease defined by the presence of hemorrhage,<br />

infection or alcoholic hepatitis in the last 3 months. 18 healthy<br />

volunteers were recruited as controls. Patients or controls using<br />

drugs known to modify the coagulation system and/or having<br />

a history of spontaneous thrombosis were excluded. The<br />

stability of TG in platelet-poor plasma testing, with or without<br />

addition of thrombomodulin (TM) or activated Protein C<br />

(PCa), has been assessed over 12 weeks by 3 repeated Calibrated<br />

Automatic Thrombinography measurements at week<br />

0, 6 and 12. Results: The intra-individual variability of endogenous<br />

thrombin potential (ETP) after addition of TM was low,<br />

about 5%, in uncomplicated cirrhosis and healthy volunteers.<br />

The inter-individual variability was about 100%, and was not<br />

influenced by age, sex or cirrhosis etiology (alcoholic or not).<br />

TG, after addition of PCa, varies over time in cirrhosis and<br />

healthy groups. In contrast, TG significantly varies over time<br />

in patients with complicated cirrhosis. TG was significantly<br />

higher in patients with cirrhosis Child B (n=20) than in Child<br />

A (n=21) and healthy volunteers (n=18). No difference was<br />

found between Child B and Child C. The ratio TG without/with<br />

TM increases with hepatic function. No coagulation imbalance<br />

persists after in vitro protein C (PC) normalisation. Conclusions:<br />

TG should be evaluated after addition of thrombomodulin, and<br />

not after addition of activated Protein C in patients with cirrhosis.<br />

Thrombin Generation increased according to the hepatic<br />

function. Thrombin generation over time, in the same patient,<br />

after addition of thrombomodulin should be evaluated as a risk<br />

factor of hypercoagulability. Hypercoagulability appears to be<br />

related to PC deficiency. A largest study should identify factors<br />

affecting inter and intra individual variabilities.<br />

Disclosures:<br />

Armando Abergel - Consulting: gilead, msd, bms; Speaking and Teaching: abbvie<br />

The following authors have nothing to disclose: Cédric Duron, Géraldine Lamblin,<br />

Aurelien Lebreton, Thomas Sinegre<br />

1506<br />

In Vivo Assessment Of Portal Hypertensive Colopathy<br />

And Clinical Outcome Of Patients With Liver Cirrhosis<br />

With Confocal Laser Endomicroscopy (CLE)<br />

Steffen Zopf 1 , Gian Eugenio Tontini 2 , Michael Vieth 3 , Markus<br />

F. Neurath 1 , Helmut Neumann 1 ; 1 Medical Department 1, Friedrich-Alexander<br />

University of Erlangen, Erlangen, Germany; 2 IRCCS<br />

Policlinico San Donato, San Donato Milanese, Italy; 3 Institute of<br />

Pathology, Klinikum Bayreuth, Bayreuth, Germany<br />

Background and Aims: Recent data has highlighted the role<br />

of mucosal integrity for bacterial translocation in the gut which<br />

is also discussed as a major cause for development of spontaneous<br />

bacterial peritonitis (SBP) and/ or hepatic encephalopathy<br />

(HE) in patients with liver cirrhosis. CLE has emerged as a<br />

valuable tool for real time diagnosis of mucosal integrity and<br />

allows in vivo imaging of commensal bacteria in the gut. Our<br />

aim was to prospectively assess the value of CLE for in vivo<br />

diagnosis of portal hypertensive colopathy and its association<br />

to Child-Pugh class, Model for End Stage Liver Disease (MELD),<br />

HE and development of SBP. Methods: Patients with established<br />

diagnosis of liver cirrhosis and portal hypertension were<br />

prospectively included. Clinical, biochemical, and ultrasound<br />

criteria, including portal vein thrombosis, ascites and collateral<br />

portosystemic vessels were assessed in addition to endoscopic<br />

criteria (e.g. esophageal varices, portal gastropathy) and betablocker<br />

intake. Fluoresceine aided CLE was performed in every<br />

patient in the sigmoid colon, rectosigmoid junction, and rectum.<br />

Afterwards biopsies were taken, unless contraindicated,<br />

for corresponding histopathological analysis. Results: Overall,<br />

more than 14,700 CLE images were collected. Confocal<br />

imaging revealed dilation and/or ectasia of microvessels, congestion<br />

of blood flow, edema, and a non-specific increase of<br />

the cellular infiltrate within the lamina propria. These findings<br />

were directly correlated to Child-Pugh class and MELD score<br />

with patients at higher scores showing more distinct changes<br />

of the microarchitecture. Of note, disturbed mucosal integrity,<br />

as observed by CLE, was strongly correlated with occurrence<br />

of HE and SBP, even in the follow-up of the patients. The procedure<br />

was well tolerated by the patients, and no adverse<br />

events were observed. Conclusions: Fluoresceine guided CLE<br />

in patients with liver cirrhosis and portal hypertensive colopathy<br />

is safe and well tolerated. In vivo imaging revealed similar<br />

microscopic changes of portal colopathy as conventional histology<br />

without the need of physical biopsies. Of note, confocal<br />

imaging corresponds to clinical outcome parameters, including<br />

development of HE and/or SPB.<br />

Disclosures:<br />

The following authors have nothing to disclose: Steffen Zopf, Gian Eugenio Tontini,<br />

Michael Vieth, Markus F. Neurath, Helmut Neumann<br />

1507<br />

Conjugated bile acids suppress the HPA axis via activation<br />

of hypothalamic glucocorticoid receptors in a<br />

rodent model of chronic biliary obstruction<br />

Matthew McMillin 1,2 , Gabriel A. Frampton 1 , Stephanie Grant 1 ,<br />

Matthew A. Quinn 3 , Sharon DeMorrow 2,1 ; 1 Texas A&M Health<br />

Science Center, College of Medicine, Temple, TX; 2 Central Texas<br />

Veterans Healthcare System, Temple, TX; 3 Department of Health<br />

and Human Services, National Institute of Environmental Health<br />

Sciences, NIH, Research Triangle Park, NC<br />

Suppression of the hypothalamic pituitary adrenal axis (HPA)<br />

occurs during cholestasis. We previously demonstrated that<br />

serum bile acids gain entry to the hypothalamus in a model<br />

of cholestasis. The aim of the current study was to determine


946A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

the effects of bile acid signaling on the HPA axis. Methods:<br />

Sprague Dawley rats were fed cholestyramine for 3 days<br />

prior to sham or bile duct ligation (BDL) surgery; corticosterone<br />

levels, hypothalamic corticotropin releasing hormone<br />

(CRH) expression, and serum and hypothalamic bile acid levels<br />

were assessed. Apical sodium-dependent bile acid transporter<br />

(ASBT)- and glucocorticoid receptor (GR)-specific Vivo<br />

morpholinos were infused in the lateral ventricle (1mg/kg/<br />

day) for 3 days, followed by injection of the bile acids cholic<br />

acid (CA), deoxycholic acid (DCA), chenodeoxycholic acid<br />

(CDCA), taurocholic acid (TCA) or glycochenodeoxycholic<br />

acid (GCDA) into the 3 rd ventricle (20 pmol) and HPA activity<br />

was assessed after 6 hr. Hypothalamic neurons were treated<br />

with various bile acids (10 mM) with or without the GR antagonist<br />

RU486 or after ASBT knockdown, and CRH expression<br />

and secretion was assessed. Bile acid uptake was assessed<br />

with the fluorescent bile acid derivative cholyl-lysyl fluorescein<br />

(CLF). The ability of bile acids to activate GR was assessed ex<br />

vivo using a commercially available competition assay and in<br />

vitro by subcellular localization of GR. Results: Cholestyramine<br />

attenuated the HPA axis suppression and increase in hypothalamic<br />

bile acid levels observed during cholestasis. Treatment of<br />

hypothalamic neurons with various bile acids suppressed CRH<br />

expression and secretion in vitro. In vivo, HPA axis suppression<br />

was only evident after injection of conjugated bile acids<br />

(TCA or GCDA). Uptake of CLF was evident in hypothalamic<br />

neurons but was inhibited in ASBT-silenced neurons. TCA- or<br />

GCDA-driven suppression of hypothalamic CRH expression<br />

and circulating corticosterone could be prevented in vitro and<br />

in vivo by inhibiting ASBT-mediated bile acid uptake. Both TCA<br />

and GCDA compete for GR binding sites with endogenous<br />

ligands in an ex vivo competition assay and caused nuclear<br />

translocation of GR in vitro. Inhibiting GR activation in vitro<br />

and in vivo prevented bile acid-driven suppression of the HPA<br />

axis. Conclusions: Taken together our data support the hypothesis<br />

that during cholestatic liver injury, bile acids gain entry to<br />

the brain, are transported into neurons through the ASBT transporter<br />

and can activate GR to suppress the HPA axis. These<br />

data also support the broader hypothesis that bile acids may<br />

act as central modulators of hypothalamic peptides that may be<br />

altered during liver disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Matthew McMillin, Gabriel A.<br />

Frampton, Stephanie Grant, Matthew A. Quinn, Sharon DeMorrow<br />

1508<br />

Characteristics, indications and outcomes in cirrhotic<br />

patients with portal hypertension undergoing endoscopic<br />

retrograde cholangiography: a case-control<br />

study<br />

Ricardo Macías-Rodríguez, Jose Luis Rodriguez-Garcia, Astrid<br />

Ruiz-Margáin, Aldo Torre; Gastroenterology, INCMNSZ, Mexico,<br />

Mexico<br />

Background: Endoscopic retrograde cholangiography (ERC)<br />

is a useful tool for diagnostic and therapeutic biliary tract diseases.<br />

In patients with cirrhosis, portal hypertension leads to<br />

clinical and biochemical alterations which can exert higher<br />

risk for complications related to ERC. However, scarce data<br />

regarding ERC in cirrhosis are available in medical literature.<br />

Aim:to evaluate the indications, general characteristics and<br />

outcomes of ERC in patients with cirrhosis and portal hypertension.<br />

Methods: This was a case-control study including 37<br />

patients (71 procedures) with cirrhosis and portal hypertension<br />

(C) and 37 age and sex-matched controls without cirrhosis<br />

(NC), undergoing ERC and evaluated 30 days post-ERC. Characteristics<br />

related to ERC (cannulation, sphincterotomy, stenting<br />

and complication rate) and cirrhosis (Child-Pugh and MELD<br />

scores; presence of ascites, hepatic encephalopathy (HE) and<br />

esophageal varices) were recorded. Mann-Whitney’s U and X 2<br />

were used as appropriate.Logistic regression and ROC analysis<br />

with Youden index were used to analyze complications<br />

in cirrhotics. Results: Main etiology of cirrhosis was hepatitis<br />

C infection (22%). Patients were classified as Child A/B/C<br />

(4/18/15), mean MELD was 17.8±6. Complications of cirrhosis<br />

were ascites (46%), esophageal varices (63%; large<br />

esophageal varices 43.7%) and HE (16%). Main differences<br />

in baseline characteristics between groups showed changes<br />

inherent to cirrhosis in biochemical parameters(higher bilirubin<br />

and lower PT, platelets, sodium and albumin). Main indication<br />

in C and NC group for ERC was choledocholithiasis in 46%<br />

and 48%, respectively p=0.816. Cannulation rate was the<br />

same in both groups, (97%, p=1.000). Biliary sphincterotomy<br />

was performed more frequently in NC patients compared to<br />

C (60 vs 35%, p=0.036); there was no difference in complications<br />

related to ERC between C and NC (10% vs 8%,<br />

p=0.677). Factors related to complications in cirrhosis were<br />

lower alkaline phosphatase levels(230± 71 vs 619± 550mg/<br />

dL) and sphincterotomy rate(71 vs 20% for complication and<br />

no complication subgroup, p=0.010). In the multivariable analysis<br />

only sphincterotomy was independently associated to the<br />

presence of complications (OR 9.8[1.7-56.3], p=0.011). ROC<br />

analysis yielded a MELD score>16 to best predict complications<br />

after ERC in cirrhosis (Se 71.4%, Sp 57.8%, AUC 0.616).<br />

Conclusion:Outcomes after ERC in patients with cirrhosis are<br />

similar to non-cirrhotics despite of the alteration in coagulation<br />

parameters and the presence of disease-specific complications,<br />

however a more cautiously approach in patients with cirrhosis<br />

undergoing sphincterotomy and MELD>16 is needed.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ricardo Macías-Rodríguez, Jose<br />

Luis Rodriguez-Garcia, Astrid Ruiz-Margáin, Aldo Torre<br />

1509<br />

Efficacy of Ornithine Phenylacetate (OP) in lowering<br />

plasma ammonia after upper gastrointestinal bleeding<br />

(UGIB) in cirrhotic patients: a multicenter, randomized,<br />

double blind trial<br />

Meritxell Ventura-Cots 1 , German Soriano 3,4 , Macarena Simon-Talero<br />

1 , Maria Torrens 1 , Albert Blanco 2 , Jose Antonio Arranz 2 , Mar<br />

Concepción 3 , Edilmar A. Alvarado 3 , Cristina Gely 3 , Eva Roman 3 ,<br />

Joan Genescà 1,4 , Juan Cordoba 1,4 ; 1 Liver Unit, Hospital Vall d’Hebron,<br />

Barcelona, Spain; 2 Metabolopathies Unit, Hospital Vall<br />

d’Hebron, Barcelona, Spain; 3 Department of Gastroenterology,<br />

Hospital de la Santa Creu i Sant Pau, Barcelona, Spain; 4 Instituto<br />

de Salud Carlos III, CIBEREHD, Madrid, Spain<br />

OP has been proposed to decrease ammonia in uncontrolled<br />

<strong>studies</strong>. Methods: Thirty-eight consecutive cirrhotic patients,<br />

aged between 18-85 years and with a creatinine level lower<br />

than 1.5 mg/dL were enrolled within 24 h of an UGIB. Patients<br />

were randomized (1:1) to receive OP (10g/24h continuous<br />

infusion for 5 days) or placebo, in addition to the usual treatment,<br />

and followed up for 28 days. Ammonia and related<br />

metabolites were assessed by HPLC and mass spectroscopy.<br />

The hypothesis was that OP would produce a 25μmol/L venous<br />

ammonia difference in the means between the groups during<br />

the first 24h. Secondary outcomes included: plasma ammonia<br />

changes at any time point, pharmacokinetic profile, hepatic<br />

encephalopathy and clinical outcome, and confirmation of<br />

safety and tolerability of OP. Treatment groups were compared<br />

using a one-sided exact Wilcoxon rank-sum test. Results: a


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 947A<br />

progressive decrease in ammonia concentration was observed<br />

in both groups, being higher in OP group at each time point,<br />

without significant differences. Decrease in ammonia in the first<br />

24h in the OP group [-20.40 μmol/L (-39.20, -2.80); median<br />

(IQR)] was twice that of the placebo group [-11.88 μmol/L<br />

(-36.75, 26.35)], without statistical significance. The subanalysis<br />

by Child-Pugh score showed a statistically significant<br />

ammonia decrease in Child-Pugh C treated patients at 36 h, as<br />

well as in the time normalized area under the curve (TN-AUC)<br />

0-120h in the OP group [40.16 μmol/L (37.7, 42.6 )] vs placebo<br />

group [ 65.5 μmol/L (54, 126); p=0.036]. A progressive<br />

decrease in plasma glycine and glutamine were observed in<br />

the treated group as compared to the placebo group. These<br />

differences reached statistical significance after 24 h of treatment<br />

and were maintained thereafter. Glutamine decrease<br />

associated with the appearance of phenylacetylglutamine in<br />

urine: u-PAG accumulated TN-AUC 12-120h placebo [5768<br />

μmol (1769,8808)] vs TN-AUC 12-120h OP [106851μmol<br />

(78802, 146666); p50 mmol/L). Ammonemia was higher in patients<br />

with HE than in those without (108 (95%CI: 86-123) vs 57<br />

(95%CI: 50-81), p=0.0004). In patients with HE, ammonemia<br />

was significantly correlated to severity of HE assessed by GCS<br />

(p=0.001). Overall, 34 HE patients experienced worsening of<br />

their neurological status during hospitalisation. Ammonemia<br />

at admission was significantly higher in those patients (128<br />

(95%CI: 108-147) vs 87 (95%CI: 64-110), p=0.007). In multivariate<br />

analysis, parameters independently associated with<br />

worsening of neurological status were MELD score and ammonia<br />

levels at admission (p=0.01 and 0.02, respectively). Conclusion:<br />

Ammonia levels at admission are correlated with the<br />

severity of HE and with outcome in cirrhotic patients with HE.<br />

Disclosures:<br />

Marika Rudler - Speaking and Teaching: Gilead, Mayoly<br />

Thierry Poynard - Grant/Research Support: Gilead; Stock Shareholder: BioPredictive<br />

Nicolas Weiss - Speaking and Teaching: Norgine, Alpha Wasserman<br />

The following authors have nothing to disclose: Simona Tripon, Maxime Mallet,<br />

Denis Monneret, Francoise Imbert-Bismut, Dominique Thabut<br />

1511<br />

Use of betablockers, previous hepatic encephalopathy<br />

and low albumin levels as risk factors of portal vein<br />

thrombosis in a cohort of cirrhotic patients<br />

Marta Lopez Gomez 1 , Elba Llop 1 , Angela Puente 2 , Juan de la<br />

Revilla 1 , Carlos Fernández-Carrillo 1 , Fernando Pons 1 , Jose Luis<br />

Martinez 1 , Natalia Fernández 1 , Maria Trapero 1 , Javier Crespo 2 ,<br />

José Luís Calleja 1 ; 1 Gastroenterology and Hepatology, Hospital<br />

Universitario Puerta de Hierro (MADRID. SPAIN), Pozuelo de<br />

Alarcon, Spain; 2 Hospital Universitario Marqués de Valdecilla,<br />

Santander, Spain<br />

Portal vein thrombosis(PVT) is a complication of liver cirrosis(LC).<br />

The aim of our study was to evaluate anual incidence<br />

of PVT and related risk factors.Methods: We retrospectively<br />

reviewed clinical and radiological data collected prospectively<br />

of consecutive cirrhotic patients included in the database<br />

of two Universitary Hospitals. Patients out of Milan criteria<br />

HCC, known PVT, TIPS and pregnancy were excluded. All<br />

patients with ultrasound diagnosis of PVT underwent MR or<br />

CTangiography.Results: From September 2013 to September<br />

2014, 747 cirrhotic patients were reviewed, 179 had exclusion<br />

criteria. Baseline characteristics are described in Table<br />

1. 23(4%) patients presented PVT during the inclusion period.<br />

Significant differences between patients with/without PVT<br />

were observed in: albumin (3.4SD0.8vs4.0SD0.5;p


948A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

HE(OR3.2 IC 1.1-8.; p0.03) were risk factors and high albumin<br />

levels(OR0.3IC0.2-0.8p=0.01) was as a protective factor.<br />

Besides, significant differences were observed in PVD(12.2SD-<br />

5vs10.7SD2;p=0.02) and SD(15SD3vs13SD2.6;p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 949A<br />

Room air PaO2, 100% FiO2 PaO2 and MAA are poor predictors<br />

of overall survival in those considered for MELD exception<br />

for HPS. Although post-LT mortality is higher in those with a<br />

100% FiO2 PaO2


950A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1516<br />

Linkage and interaction between portal hemodynamics<br />

and serum sodium concentration in cirrhosis<br />

Hitoshi Maruyama, Takayuki Kondo, Tadashi Sekimoto, Soichiro<br />

Kiyono, Osamu Yokosuka; Department of Gastroenterology and<br />

Nephrology, Chiba University Graduate School of Medicine,<br />

Chiba, Japan<br />

Background/Aim: Although a hyponatremia is frequently<br />

observed in patients with cirrhosis, the prognostic influence<br />

with respect to the presence of concomitant portal hemodynamic<br />

abnormality has yet to be determined. The study aimed<br />

to examine the linkage between portal hemodynamics and<br />

hyponatremia, and their interactions on the prognosis of cirrhosis.<br />

Methods: This study was based on the subgroup analysis<br />

using medical records collected in the prospective study<br />

regarding the influence of portal hemodynamics on clinical<br />

outcomes from November 2007 to November 2012. Clinical<br />

findings including portal hemodynamic parameters measured<br />

by Doppler ultrasound and serum sodium concentrations were<br />

examined with respect to the prognosis. Following patients<br />

were excluded: i) with hepatocellular carcinoma (HCC) diagnosed<br />

by contrast-enhanced CT or magnetic resonance imaging,<br />

or a treatment history of HCC, ii) receiving medications<br />

with vasoactive drugs such as β-blockers, antiviral therapy, or<br />

interventional procedures such as a transjugular intrahepatic<br />

portosystemic or peritoneo-venous shunt, prior to or during,<br />

the study period. Results: There were 153 cirrhosis patients<br />

(62.2 ± 12.0 years; median observation period, 34.1 m). Sixteen<br />

patients were accompanied with hyponatremia (Na <<br />

135 mEq/L), who showed a significantly greater frequency of<br />

possessing a splenorenal shunt (SRS; p = 0.0068), and 137<br />

patients without hyponatremia. Serum sodium concentrations<br />

were significantly lower in patients with a SRS than in those<br />

without (p = 0.0193). An increased prothrombin time-international<br />

normalized ratio was a significant predictive factor for<br />

developing hyponatremia a year later (8/96; Hazard ratio<br />

14.415; p = 0.028). The cumulative survival rate was significantly<br />

lower in patients with hyponatremia (46.7% at 1 and<br />

3 years) than in those without (91.8% at 1 year, 76.8% at 3<br />

years; P < 0.001), and that was significantly lower in patients<br />

who had developed hyponatremia after one year (100% at 1<br />

year, 62.5% at 3 years) than those who had not (100% at 1<br />

year, 89.0% at 3 years; P < 0.001). In addition, the cumulative<br />

survival rate was significantly worse in patients with both hyponatremia<br />

and a SRS (20% at one year) than in others (91.8%<br />

at 1 year, 76.3% at 3 years, and 63.1% at 5 years in patients<br />

with neither hyponatremia nor a SRS, p < 0.0001; 77.4% at 1<br />

year, 72.6% at 3 years and 58.1% at 5 years in patients with<br />

hyponatremia or a SRS, p = 0.014). Conclusions: There was<br />

a close linkage between the serum sodium concentration and<br />

portal hemodynamic abnormality, a presence of SRS, and their<br />

interaction may negatively influence the prognoses in cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hitoshi Maruyama, Takayuki<br />

Kondo, Tadashi Sekimoto, Soichiro Kiyono, Osamu Yokosuka<br />

1517<br />

Ammonia Produces Pathological Changes And Activation<br />

In Human Hepatic Stellate Cells And Is A Target For<br />

Therapy In Portal Hypertension<br />

Francesco De Chiara, Rajiv Jalan, Vairappan Balasubramaniyan,<br />

Fausto Andreola, Massimo Malago, Massimo Pinzani, Rajeshwar<br />

Mookerjee, Krista Rombouts; UCL, London, United Kingdom<br />

Background: Hepatic stellate cells (HSC) are vital to hepatocellular<br />

function and the liver’s response to injury. They share<br />

a phenotypic homology with astrocytes that are central in the<br />

pathogenesis of hepatic encephalopathy, a condition in which<br />

hyperammonemia plays a pathogenic role. This study tested<br />

the hypothesis that ammonia modulates human HSC activation<br />

in vitro and in vivo, and evaluated whether ammonia lowering,<br />

by using L-ornithine phenylacetate (OP), modifies HSC activation<br />

in vivo and reduces portal pressure in a bile duct ligation<br />

(BDL) model. Methods: Primary human HSCs were isolated<br />

and cultured. Proliferation (BrdU), metabolic activity (MTS),<br />

morphology (TEM, light and immunofluorescence microscopy),<br />

HSC activation markers by protein analysis, and changes in<br />

oxidative status (ROS) and endoplasmic reticulum (ER) were<br />

evaluated to identify effects of ammonia challenge (50 mM,<br />

100 mM, 300 mM) over 24-72hrs. Changes in plasma ammonia<br />

levels, markers of HSC activation, portal pressure, hepatic<br />

eNOS activity and expression of its regulatory proteins were<br />

quantified in hyperammonemic BDL animals, and after OP<br />

treatment. Results: Pathophysiological ammonia concentrations<br />

caused significant and reversible changes in cell proliferation,<br />

metabolic activity and activation markers of hHSC<br />

in vitro. Highly significant alterations in cellular morphology,<br />

characterised by cytoplasmic vacuolisation, ER enlargement<br />

and ROS production, pro-inflammatory gene expression were<br />

observed together with HSC-related activation markers such as<br />

α-SMA, myosin IIa, IIb, and PDGF-Rβ. Treatment with OP significantly<br />

reduced plasma ammonia (BDL 199.1mmol/L±43.65<br />

vs. BDL+OP 149.27mmol/L±51.1, P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 951A<br />

by, bacterial infections. We recently reported the effects of<br />

sepsis in the liver microcirculation of healthy rats; however the<br />

influence of bacterial infections on the liver sinusoidal circulation<br />

in cirrhotic livers remains unknown. Statins have shown<br />

beneficial effects ameliorating liver endothelial dysfunction in<br />

cirrhosis. The present study aimed at characterizing the effects<br />

of simvastatin on liver microvascular dysfunction in a cirrhotic<br />

rat model of endotoxemia. Methods: Lipopolysaccharide (LPS;<br />

1mg/kg) or saline was given to: (1) CCl 4<br />

-cirrhotic rats treated<br />

with placebo; (2) cirrhotic rats treated with simvastatin (25<br />

mg/kg/24 h, orally) from 3 days before LPS/saline injection.<br />

Hemodynamic parameters (Mean Arterial Pressure-MAP, Portal<br />

Pressure-PP, Portal Blood Flow-PBF, Hepatic Vascular Resistance-HVR),<br />

followed by Portal Perfusion Pressure (PPP) and<br />

microvascular functionality in response to acetylcholine were<br />

assessed. Underlying molecular/cellular mechanisms (liver<br />

fibrosis, HSC activation, inflammation, oxidative stress, NO<br />

release) were investigated. Results: LPS administration markedly<br />

aggravated hepatic microvascular dysfunction in cirrhotic rats<br />

(PPP +17%; vasodilatory response to Ach -27%; p


952A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

liver sections. Ascitic fluid was evaluated for bacterial growth<br />

in thioglycolate medium. Ex-vivo liver perfusion experiments<br />

were performed for assessing endothelium-dependent and –<br />

independent reactivity. Results: Compared with controls, rats<br />

with oral CCl4 gavage tended to show decreased survival and<br />

body weight gain, both of which were further worsened by<br />

enoxaparin 180 U/kg bw (p< 0.01). Rats with CCL4-induced<br />

cirrhosis showed altered laboratory parameters (increased<br />

INR, AST, ALT, bilirubin / decreased albumin, total proteins,<br />

glucose and platelets) regardless of enoxaparin treatment. In<br />

all experimental models, cirrhotic rats receiving saline and<br />

those receiving enoxaparin showed similar increases in the<br />

area of liver fibrosis compared with controls (p< 0.001). Rats<br />

with cirrhosis induced by oral CCl4 gavage and by BDL surgery<br />

developed increases of portal pressure and spleen-to-bw<br />

ratios compared with control rats (p< 0.001), regardless of<br />

enoxaparin treatment. Among rats with ascites, a similar proportion<br />

presented positive bacterial cultures (CCl4+saline 2<br />

of 7 vs. CCl4+enoxaparin 4 of 8, NS; BDL+saline 1 of 6 vs.<br />

BDL+enoxaparin 2 of 7, NS). Potential effects of enoxaparin<br />

on hepatic vascular reactivity were only observed in rats<br />

receiving enoxaparin 180 U/kg bw from the beginning of<br />

CCl4 administration, and consisted of increased portal venous<br />

resistance after addition of acetylcholine or S-nitroso acetylpenicillamine<br />

(SNAP, both p< 0.05) and increased sinusoidal<br />

resistance after addition of SNAP (p< 0.05). Conclusion: Our<br />

experimental data do not support a role of long-term treatment<br />

with enoxaparin for improving liver fibrosis, portal hypertension<br />

or endothelial dysfunction in cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jose Ignacio Fortea, Alexander<br />

Zipprich, Carolina Fernandez Mena, Christopher F. Rose, Juan Bañares, Marta<br />

Puerto, Cristina R. Bosoi, Jorge Almagro, Marcus Hollenbach, Marc-André Clément,<br />

Javier Vaquero, Rafael Bañares, Cristina Ripoll<br />

1521<br />

Neuro-inflammation in Cirrhotic Mice is Dependent on<br />

Gut Microbial Colonization: Implications for Hepatic<br />

Encephalopathy<br />

Dae Joong Kang 1 , Siddhartha Ghosh 1 , Daniel Carl 1 , Arun J.<br />

Sanyal 1 , Ryan B. Sartor 2 , Phillip B. Hylemon 1 , Huiping Zhou 1 , Runping<br />

Liu 1 , Xiang Wang 1 , Jing Yang 1 , Chunhua Jiao 1 , Jasmohan S.<br />

Bajaj 1 ; 1 VCU and McGuire VAMC, Richmond, VA; 2 Gnotobiotic<br />

Rodent Research Center, University of North Carolina, Chapel<br />

Hill, NC<br />

Neuro-inflammation has been proposed as a mechanism for<br />

hepatic encephalopathy (HE) but the source for this is not clear.<br />

Specifically, brain inflammation in cirrhotic models in the germfree<br />

mice (GF) condition has not been studied. Aim: Assess the<br />

impact of intestinal microbiota on brain inflammation in cirrhotic<br />

mice in a conventional (conv) and GF setting. Methods:<br />

We used C57BL/6 mice in a conv and GF setting. Sixteen mice<br />

were in the GF condition; of which 6 remained GF while 10<br />

were administered CCL4 1ml/kg in sesame oil using gavage<br />

twice/week till week 16. An age-balanced group of 10 conv<br />

mice were included; 5 remained healthy while cirrhosis using<br />

CCL4 was induced in 5 mice using 1ml/kg CCL4 administration<br />

twice a week for 12 weeks. On sacrifice, brain tissue was<br />

separated into cortex (Cx) and cerebellum (Cbl). Analysis was<br />

performed for mRNA of inflammation (IL-6, IL-1β) and activation<br />

of microglia and glia (IBA-1, GFAP) using qPCR (adjusted for<br />

GADPH) in both brain regions of all mice using Kruskall-Wallis<br />

tests. Results: All mice survived till end of the experiment. Both<br />

CCL4-treated groups (conv & GF) showed cirrhosis on sacrifice<br />

on visual and histological examination. Comparison within GF<br />

group: There was no significant difference in inflammation or<br />

microglial/glial activation in both brain regions in GF controls<br />

vs. CCL4 GF cirrhotics (table) Comparison within Conv<br />

group: There was a significant increase in mRNA of IL-1β,<br />

GFAP and IBA-1 in both Cbl and Cx of CCL4 conv cirrhotic<br />

mice compared to their conv non-cirrhotic counterparts. IL-6<br />

was unchanged. Comparison between GF Cirr and Conv cirr:<br />

Conv CCL4 cirrhotic mice showed a significant increase in Cbl<br />

and Cx IL-1β, GFAP and IBA-1compared to GF cirrhotic mice.<br />

IL-6 was again not different between groups. Conclusions: Glial<br />

and microglial activation and induction of IL-1β occurs in cirrhotic<br />

mice only in the presence of gut microbiota. This suggests<br />

an essential role of the gut microbiota in the development<br />

of neuro-inflammation in cirrhosis and could have implications<br />

for the gut-brain axis management in hepatic encephalopathy.<br />

Neuroinflammation and Glial/Microglial Expression between<br />

Groups<br />

Disclosures:<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Jasmohan S. Bajaj - Advisory Committees or Review Panels: Salix, Merz, otsuka,<br />

ocera, grifols, american college of gastroenterology; Grant/Research Support:<br />

salix, otsuka, grifols<br />

The following authors have nothing to disclose: Dae Joong Kang, Siddhartha<br />

Ghosh, Daniel Carl, Ryan B. Sartor, Phillip B. Hylemon, Huiping Zhou, Runping<br />

Liu, Xiang Wang, Jing Yang, Chunhua Jiao<br />

1522<br />

Emricasan, a pan caspase inhibitor, improves survival<br />

and portal hypertension in a murine model of long-term<br />

common bile-duct ligation<br />

Akiko Eguchi 1 , Alexander Wree 1 , Yukinori Koyama 1 , Casey Johnson<br />

1 , Ryota Nakamura 1 , Patricia C. Contreras 2 , Ariel E. Feldstein 1 ;<br />

1 UCSD, La Jolla, CA; 2 Conatus Pharmaceuticals, San Diego, CA<br />

Development of portal hypertension (PTH) is a central prognostic<br />

factor in patients with cirrhosis. Current pharmacotherapy<br />

remains limited and novel therapies are greatly needed. Circulating<br />

microparticles (MPs) are released by hepatocytes in<br />

a caspase dependent manner, are increased in circulation of<br />

patients with cirrhosis and contribute to PTH via induction of<br />

impair vasoconstrictor responses Here we tested the hypothesis<br />

that Emircasan (IDN-6556) a pan-caspase inhibitor ameliorates<br />

PTH via its anti-fibrotic effects and reduction in release<br />

of MPs. Methods: Secondary biliary cirrhosis was induced by<br />

long-term common bile-duct ligation (CBDL) in C57BL/6 mice.<br />

Controls were sham-operated, the abdominal cavity opened,<br />

but no ligature placed. Mice were treated with 10 mg/kg/<br />

day of IDN-6556 or placebo for 3 wks via i.p. injection (4<br />

groups, n= 7 -12 in each group).. After 3 wks, portal pressure<br />

was measured in each mouse and serum and livers were collected.<br />

Circulating MPs were isolated and characterized using<br />

various approaches including FACS analysis. Portal pressure<br />

was measured using catheter from ileocolic vein. Hepatocellular<br />

damage was assessed by liver histopathology, serum ALT<br />

levels, and TUNEL assay. Liver fibrosis was assessed by digital<br />

image analysis of Sirius red stained sections. Results: In


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 953A<br />

contrast to CBDL-Placebo group, nearly all CBDL-IDN-6556<br />

survived (p


954A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

HVPG ≥10mmHg (R=0.818, p=0.007). Conclusion We have<br />

demonstrated that parameters related to both hepatic architecture<br />

and splanchnic haemodynamics derived from non-contrast<br />

quantitative MR imaging techniques correlate significantly with<br />

HVPG.<br />

Disclosures:<br />

The following authors have nothing to disclose: Naaventhan Palaniyappan, Eleanor<br />

Cox, Andrew Austin, Indra Neil Guha, Susan Francis, Guruprasad P. Aithal<br />

1525<br />

The soluble guanylyl cyclase stimulator riociguat<br />

reduces liver fibrosis and portal pressure in cirrhotic rats<br />

Philipp Schwabl 1 , Ksenia Brusilovskaya 1 , Florian Riedl 1 , David<br />

J. Bauer 1 , Bastian Strobel 1 , Paul Supper 1 , Hubert Hayden 1 ,<br />

Michael Trauner 1 , Bruno K. Podesser 2 , Katharina Wochner 2 ,<br />

Markus Peck-Radosavljevic 1 , Thomas Reiberger 1 ; 1 Division of Gastroenterology<br />

and Hepatology, Department of Internal Medicine<br />

III, Medical University of Vienna, Vienna, Austria; 2 Department<br />

for Biomedical Research, Medical University of Vienna, Vienna,<br />

Austria<br />

Background: In liver cirrhosis nitric oxide (NO) signaling as<br />

well as the function of its key downstream target soluble guanylyl<br />

cyclase (sGC) is distorted. The sGC stimulator riociguat<br />

(RIO) is used for treatment of pulmonary hypertension. Since<br />

RIO has been shown to have additional antifibrotic activity,<br />

we aimed to investigate effects of RIO in rats with cirrhotic<br />

portal hypertension. Methods: Cirrhosis was induced by carbontetrachloride<br />

injections (CCl4, 8 weeks) or by bile duct ligation<br />

(BDL, 3 weeks) in Sprague Dawley rats. Controls received<br />

olive-oil (OO) or underwent sham operation (SO), respectively.<br />

RIO (1mg/kg/d, gavage) or vehicle (VEH) was administered<br />

from weeks 6-8 in CCl4/OO and from weeks 2-3 in BDL/<br />

SO animals. Mean arterial pressure (MAP), heart rate (HR),<br />

portal pressure (PP) and superior mesenteric artery blood flow<br />

(SMABF) were measured. Porto-systemic and spleno-renal shunting<br />

was assessed by colored microspheres technique. Hepatic<br />

fibrosis was quantified by hepatic hydroxyproline content (HP)<br />

and histology. Results: CCl4 rats presented with marked cirrhosis,<br />

hyperdynamic circulation, portal hypertension and elevated<br />

SMABF compared to OO animals. In CCl4 rats, RIO had no<br />

effect on HR, MAP and PP but significantly decreased SMABF<br />

(74±11 vs. 43±6 mL/min; p=0.036). Further, in CCl4 cirrhosis<br />

RIO significantly decreased spleen (1.3±0.1 vs. 1.0±0.1g;<br />

p=0.008) and liver (16.7±1.2 vs. 11.5±2.4g; p=0.015)<br />

weight. BDL rats presented with hepatosplenomegaly and portal<br />

hypertension. In the BDL model RIO significantly increased<br />

MAP (85±17 vs. 108±14mmHg; p=0.016) and decreased PP<br />

(13.2±2.5 vs. 10.1±2.4; p=0.048), but had no effect on HR,<br />

SMABF or shunting. HP was significantly decreased (286±147<br />

vs. 144±74 μg; p=0.039) in BDL-RIO rats, compared to BDL-<br />

VEH animals. Conclusions: RIO treatment significantly reduced<br />

liver fibrosis and ameliorated hyperdynamic circulation as well<br />

as portal pressure in biliary cirrhosis. In toxic cirrhosis RIO<br />

prevented development of hepatosplenomegaly and reduced<br />

splanchnic blood inflow compared to controls. RIO could be<br />

a good candidate for combination <strong>studies</strong> with other agents<br />

known to improve hepatic hemodynamics in cirrhosis of various<br />

etiologies.<br />

Disclosures:<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Markus Peck-Radosavljevic - Advisory Committees or Review Panels: Bayer, Gilead,<br />

Janssen, BMS, AbbVie; Consulting: Bayer, Boehringer-Ingelheim, Jennerex,<br />

Eli Lilly, AbbVie; Grant/Research Support: Bayer, Roche, Gilead, MSD, AbbVie;<br />

Speaking and Teaching: Bayer, Roche, Gilead, MSD, Eli Lilly, AbbVie, Bayer<br />

Thomas Reiberger - Consulting: Xtuit; Grant/Research Support: Roche, Gilead,<br />

MSD, Phenex; Speaking and Teaching: Roche, Gilead, MSD<br />

The following authors have nothing to disclose: Philipp Schwabl, Ksenia Brusilovskaya,<br />

Florian Riedl, David J. Bauer, Bastian Strobel, Paul Supper, Hubert<br />

Hayden, Bruno K. Podesser, Katharina Wochner<br />

1526<br />

Portal tract fibrosis is induced by hepatic arterial buffer<br />

response in portal hypertensive rats: A role for macrophage<br />

miR154 released via exosome<br />

Li Gong 1,2 , Chao Dong 1,3 , Teruo Utsumi 1 , Yasuko Iwakiri 1 ; 1 Internal<br />

Medicine, Section of Digestive Diseases, Yale University School<br />

of Medicine, New Haven, CT; 2 Anesthesiology, The third Xiangya<br />

Hospital, Central South University, Changsha, China; 3 General<br />

Surgery, Xiangya Hospital, Central South University, Changsha,<br />

China<br />

Background: The hepatic arterial buffer response increases<br />

hepatic arterial (HA) flow in response to decreased portal<br />

venous flow in portal hypertension. Hemodynamic stresses<br />

associated with increased HA flow may influence intrahepatic<br />

microenvironments around the HA, including resident fibroblasts<br />

of the portal tract region, portal myofibroblast (PMF),<br />

which is a major contributor to portal tract fibrosis. The relation<br />

between increased HA flow and portal tract fibrosis has<br />

not been determined. We hypothesize that hemodynamic<br />

stresses associated with increased HA flow promote infiltration<br />

of macrophages that modulate PMF function and contribute<br />

to the development of portal tract fibrosis in portal hypertension.<br />

Method: Male Sprague Dawley rats underwent sham<br />

operation or partial portal vein ligation (PPVL) to induce portal<br />

hypertension. Livers were isolated 10 days after the surgery<br />

for histological, immunohistochemical and biochemical analyses.<br />

Microarray analysis was performed to determine changes<br />

in gene profiling between sham and PPVL rats. To examine<br />

exosome transfer from macrophages to PMF, co-culture experiments<br />

were performed using macrophage-rich cell population<br />

isolated from spleens of rats that express green fluorescent<br />

protein (GFP) and PMF of normal rats. Exosomes were verified<br />

by scanning electron microscope, CD63 (an exosome marker),<br />

and nanoparticle tracking analysis. Results: Sirius red staining<br />

of livers determined that portal tract fibrosis significantly<br />

developed in rats 10 days after PPVL (2-fold, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 955A<br />

Hepatic arterial buffer response induces portal tract fibrosis in<br />

portal hypertension. miR154 in exosomes released from macrophages<br />

to PMF may play a role in PMF proliferation and<br />

contribute to portal tract fibrosis<br />

Disclosures:<br />

The following authors have nothing to disclose: Li Gong, Chao Dong, Teruo<br />

Utsumi, Yasuko Iwakiri<br />

1527<br />

Diabetes is Associated with Stool and Sigmoid Mucosal<br />

Microbiota Dysbiosis Independent of MELD score in Cirrhosis:<br />

Implications for Prognostication<br />

Jasmohan S. Bajaj 1 , Naga Betrapally 2 , Douglas M. Heuman 1 , Melanie<br />

White 1 , Kalyani Daita 1 , Ariel Unser 1 , Edith A. Gavis 1 , Phillip<br />

B. Hylemon 1 , Swati S. Almeida-Dalmet 2 , Masoumeh Sikaroodi 2 ,<br />

Patrick M. Gillevet 2 ; 1 VCU and McGuire VAMC, Richmond, VA;<br />

2 Microbiome Analysis Center, George Mason University, Manassas,<br />

VA<br />

Cirrhotics have an unfavorable gut microbial profile (dysbiosis)<br />

that is linked to a pro-inflammatory milieu and several<br />

complications. Type 2 diabetes(DM) can increase the risk of<br />

complications but the mechanisms are unclear. Aim: Define<br />

the effect of DM on gut microbiota in cirrhotics and its linkage<br />

with cirrhosis and DM severity.Methods: Cirrhotic outpts (with/<br />

without DM) underwent stool & sigmoid mucosal microbiota<br />

analysis (multi-tag pyrosequencing). DM severity (insulin use,<br />

HgbA1c within 6 mths) & cirrhosis severity (MELD) data were<br />

collected. The cirrhosis dysbiosis ratio (CDR: Lachnospiraceae<br />

+ Ruminococceae + Clostridiales XIV + Veillonellaceae/Enterobacteriaceae<br />

+ Bacteroidaceae, low score is worse) ratio of<br />

commensal vs. potentially pathogenic taxa in stool was calculated.<br />

LFSe analysis with multiple comparisons adjustment for<br />

stool/mucosal microbiota were performed between cirrhotics<br />

with/without DM & those on/not on insulin. Results: 287 cirrhotics<br />

(57yrs, MELD 11, 39% hepatic encephalopathy (HE),<br />

40% HCV, 17% NASH) gave stool while 72 cirrhotics (29%<br />

DM, 36% HE) underwent sigmoidoscopy. 30% (n=87) of cirrhotics<br />

had DM (45% on insulin, HgbA1c 6.9). MELD & HE%<br />

was similar but DM pts had higher %NASH & BMI (Table).<br />

Microbiome analysis: Stool: showed significant dysbiosis in<br />

DM with a low CDR. This was due to a higher Bacteroidaceae<br />

(29% vs 14%, p=0.04) &Streptococcaceae (0% vs 2%,<br />

p=0.02) with lower commensal abundance. The lower bacteria<br />

in DM were Ruminococcaeae (3% vs 8%, p


956A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1529<br />

Elevated galectin-3 in cirrhotic heart increases TNFα<br />

and inhibits cardiac contractility in rats<br />

Hongqun Liu, Samuel S. Lee; university of calgary, Calgary, AB,<br />

Canada<br />

Background: Galectin-3 plays a key pathogenic role in heart<br />

failure. However, the role of galectin-3 in cirrhotic cardiomyopathy<br />

is unclear. The present study aimed to investigate the<br />

effects of galectin-3 on cardiomyopathy. Methods: Two sets of<br />

rats were used in this study, one to study structural changes and<br />

another for a functional study. There were 4 groups (n=6 in<br />

each group) in each set: bile duct ligation (BDL); BDL + galectin-3<br />

inhibitor (N-acetyllactosamine, LacNAc, 5 mg/kg); sham<br />

operation and sham + LacNAc. The rats were sacrificed 4<br />

weeks after surgery. Cardiac mRNA and protein contents were<br />

evaluated by RT-PCR and Western blots (Galectin-3, PKCα,<br />

fibronectin, collagen I and III). Heart and serum TNFα was<br />

measured by ELISA. Isolated cardiomyocyte contractility was<br />

measured by an IonOptix cell-shortening video system. Results:<br />

the galectin-3, fibronectin and collagen I/ III ratio were significantly<br />

increased in cirrhotic hearts. TNFα was significantly<br />

increased in BDL rats in both serum and heart compared with<br />

sham controls (744 ± 84 pg/ml vs 566 ± 61 pg/ml in serum,<br />

p< 0.01 and 310 ± 32 pg/mg protein vs 189 ± 21 pg/<br />

mg protein in heart, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 957A<br />

Adipose triglyceride lipase (ATGL) is the rate-limiting enzyme<br />

for triglyceride hydrolysis in adipose tissue, muscle, and liver.<br />

In a murine model of cancer cachexia, global deletion of ATGL<br />

prevented tumor-induced weight loss by inhibiting both adipose<br />

tissue and skeletal muscle wasting. Whether tissue-specific<br />

ATGL inactivation can ameliorate chronic-liver disease<br />

associated cachexia is unknown. Our aim was to determine<br />

the role of adipose tissue ATGL on cachexia development in<br />

a murine model of biliary injury (bile duct ligation). Methods:<br />

Adipose-tissue specific ATGL-knockout (AAKO, N = 10) and<br />

wild type C57/BL6 (N = 10) male mice 10-15 weeks old were<br />

treated with bile duct ligation (BDL) or sham laparotomy. Calorie<br />

intake and body weight were measured weekly. Body composition<br />

was determined weekly using Echo-MRI, and 36-hour<br />

respiratory exchange ratio (RER) and activity were measured<br />

in an Oxymax Comprehensive Laboratory Animal Monitoring<br />

System (CLAMS). Animals were sacrificed 4 weeks after surgery.<br />

Results: At baseline, AAKO mice exhibited increased<br />

adipose tissue mass compared to WT mice. BDL-treated WT<br />

and AAKO mice developed moderate fibrosis by Sirius Red<br />

staining. There were no differences in caloric intake between<br />

genotypes and treatment groups. BDL significantly reduced fat<br />

mass and lean mass (% total body weight) in AAKO mice<br />

only (p


958A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

values of the spleen was 0.84. VS values of the spleen were<br />

significantly lower in the group of patients with F2 varices with<br />

the presence of the major shunt. Conclusion; In patients with<br />

liver cirrhosis, Vs values of the spleen by VTQ were useful in<br />

the detection of EGV. They were extremely useful in particular<br />

in detecting esophageal and gastric varices which require<br />

treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chikage Nakano, Takashi<br />

Nishimura, Tomoko Aoki, Kyohei Kishino, Yoshihiro Shimono, Kunihiro Hasegawa,<br />

Ryo Takata, Kazunori You, Akio Ishii, Tomoyuki Takashima, Yoshiyuki<br />

Sakai, Nobuhiro Aizawa, Naoto Ikeda, Hiroki Nishikawa, Yoshinori Iwata,<br />

Hirayuki Enomoto, Shuhei Nishiguchi, Hiroko Iijima<br />

1534<br />

All-oral direct acting antivirals can offer significant cost<br />

savings and cost effectiveness for treating hepatitis C<br />

in medically underserved areas in the United States:<br />

Impact of task-shifting on access<br />

Channa R. Jayasekera 1 , Nathaniel Smith 2 , Ryan B. Perumpail 1 ,<br />

Robert J. Wong 3 , Rachel Beckerman 2 , Zobair M. Younossi 4,5 ,<br />

Aijaz Ahmed 1 ; 1 Division of Gastroenterology and Hepatology,<br />

Stanford University Medical Center, Stanford, CA; 2 CB Partners,<br />

New York, NY; 3 Division of Gastroenterology and Hepatology,<br />

Alameda County Health System - Highland Hospital, Oakland, CA;<br />

4 Center for Liver Diseases, Inova Fairfax Hospital, Falls Church,<br />

VA; 5 Betty and Guy Beatty Center for Integrated Research, Inova<br />

Health System, Falls Church, VA<br />

Purpose: The superior safety and tolerability of all-oral<br />

direct-acting antivirals (DAAs) for chronic hepatitis C (HCV)<br />

allows specialist physicians to devolve treatment monitoring to<br />

mid-level healthcare providers, particularly in areas of health<br />

worker shortage (task-shifting). Based on our task-shifting experiences<br />

in rural clinics, we used a decision-analytic Markov<br />

model to assess 1) budgetary impact and 2) cost-effectiveness<br />

of task-shifted HCV treatment with second generation all-oral<br />

DAAs (2nd Gen DAA, e.g. ledipasvir/sofosbuvir) vs. first generation<br />

DAA (e.g. boceprevir)+peginterferon+ribavirin (PR+1st<br />

Gen DAA) Methods: We modelled a hypothetical cohort of<br />

1000 patients with HCV genotype 1 with demographics mirroring<br />

those in our rural clinics, over a lifetime horizon. Cost of<br />

treatment monitoring, cost per sustained virological response<br />

(SVR), cost per patient, and overall budget impact were compared.<br />

Cost-effectiveness of 2nd Gen DAA over PR+1st Gen<br />

DAA were evaluated from a US third-party payer perspective,<br />

using published clinical trial and real world SVR rates for the<br />

regimens. Treatment with 2nd Gen DAA was assumed to double<br />

treatment capacity; that is 1,000 patients receiving 2nd<br />

Gen DAA (intervention arm), 500 patients receiving PR+1st<br />

Gen DAA, and 500 untreated patients (comparator arm).<br />

Transition probability, utility, and cost estimates (2014 dollars)<br />

were based on literature and hepatologists consensus.<br />

Results: Driven by limited monitoring, shorter treatment duration,<br />

and lower reimbursement rates, 2nd Gen DAA treatment<br />

monitoring by mid-level providers led to reductions of 74% in<br />

monitoring costs, 38% in cost per SVR, and 25% in total cost<br />

of treatment compared with PR+1st Gen DAA. Assuming a willingness-to-pay<br />

threshold of $50,000 per quality-adjusted life<br />

year (QALY), 2nd Gen DAA treatment monitoring by mid-level<br />

providers demonstrated higher efficacy and lower total costs<br />

(i.e. a dominant strategy). Conclusions: In this model based<br />

on rural clinic populations, task-shifting of 2nd Gen DAA treatment<br />

monitoring from specialists to mid-level providers led to<br />

significant cost savings and was cost-effective. These findings<br />

support task-shifting as a potential strategy to overcome significant<br />

gaps in access to all-oral DAA regimens in medically<br />

underserved areas in the United States.<br />

Disclosures:<br />

Nathaniel Smith - Consulting: Gilead Sciences<br />

Rachel Beckerman - Consulting: Gilead Sciences<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

The following authors have nothing to disclose: Channa R. Jayasekera, Ryan B.<br />

Perumpail, Robert J. Wong<br />

1535<br />

Inappropriate Laboratory Utilization: Acute Viral Hepatitis<br />

Serologies<br />

Kamran Hussaini 1 , Muhammad B. Hammami 1 , Edward Charbek<br />

4,1 , Robin Chamberland 2 , Hamza Khalid 3,1 , Brent A. Neuschwander-Tetri<br />

3,1 ; 1 Internal Medicine, Saint Louis University, St.<br />

Louis, MO; 2 Pathology, Saint Louis University, St. Louis, MO; 3 Gastroenterology<br />

& Hepatology, Saint Louis University, St. Louis, MO;<br />

4 Pulmonary, Critical Care & Sleep Medicine, Saint Louis University,<br />

St. Louis, MO<br />

Purpose: Inappropriate laboratory tests contribute to excessive<br />

health care costs. We have observed acute viral hepatitis serologic<br />

testing in patients with chronic liver disease, sometimes<br />

repetitively, in the absence of substantially elevated aminotransferases.<br />

Our aim was to determine the rate of unnecessary<br />

testing for acute hepatitis A (HAV) and B (HBV) infections in<br />

our institution and estimate the impact on health care costs.<br />

Methods: The results of serum anti-HAV IgM and anti-HBc IgM,<br />

and alanine aminotransferase (ALT) and aspartate aminotransferase<br />

(AST) were collected from the Saint Louis University Hospital<br />

Electronic Medical record systems, Epic and Meditech,<br />

between January 2010 and Nov 2014. Due to an error in data<br />

collection no data was collected for June to December 2013.<br />

The highest value of ALT or AST within the 30 days before testing<br />

anti-HAV IgM and anti-HBc IgM were extracted. Serological<br />

tests without ALT and AST values were excluded. Results were<br />

categorized as shown in the table below; ALT or AST level ><br />

100 Units/L was chosen to be the lowest level appropriate for<br />

ordering the serologies. The cost of inappropriate testing was<br />

estimated based on Medicare reimbursement of $15.32 for<br />

anti-HAV IgM and $58.92 for anti-HBc IgM. Results: Anti-HAV<br />

IgM and anti-HBc IgM were tested a total of 2439 and 4462<br />

times, respectively over the approximately 5 year period. As<br />

shown in the table below, 72.7% of anti-HAV IgM and 82.7%<br />

of anti-HBc IgM tests were inappropriate. Medicare payment<br />

would have been on average $5,435 per year (3.37% of<br />

charges) for anti-HAV IgM and $43,494 per year (13.57%<br />

of charges) for anti-HBc IgM. If the threshold for inappropriate<br />

testing is set at an ALT and AST < 250 U/L, then 90.7% of anti-<br />

HAV IgM and 93.7% anti-HBc IgM tests were inappropriate.<br />

The actual health care costs may be higher based on higher<br />

reimbursement rates from commercial insurers. Conclusions:<br />

Most testing for acute viral hepatitis is ordered inappropriately.<br />

We predict this is a nationwide practice that needs to be<br />

addressed through education and other interventions that might


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 959A<br />

include warnings provided in the electronic ordering system<br />

and instituting a policy for laboratory personnel to confirm the<br />

test indication with the ordering provider before collecting and<br />

processing the specimen.<br />

Disclosures:<br />

Robin Chamberland - Advisory Committees or Review Panels: Bacterioscan, Inc.<br />

Brent A. Neuschwander-Tetri - Advisory Committees or Review Panels: Nimbus<br />

Therapeutics, Bristol Myers Squibb, Janssen, Mitsubishi Tanabe, Conatus,<br />

Scholar Rock<br />

The following authors have nothing to disclose: Kamran Hussaini, Muhammad B.<br />

Hammami, Edward Charbek, Hamza Khalid<br />

1536<br />

Cost-effectiveness analysis of sofosbuvir plus ledipasivir<br />

in the treatment of chronic hepatitis C virus genotype 1<br />

infection in real life practice in China<br />

Bing Li 1 , Jing Chen 2 , Cheng Wang 2,3 , Vanessa Wu 2 , April Wong 2 ,<br />

Qing Shao 1 , Dong Ji 1 , Fan Li 1 , Yudong Wang 2 , Xiaoxia Niu 1 , Jialiang<br />

Liu 1 , Wucai Yang 1 , Yiming Fu 1 , Guofeng Chen 1 , George K.<br />

Lau 2 ; 1 Liver Fibrosis Diagnosis and Treatment Center, 302 Hospital,<br />

Beijing, China; 2 Gastroenterology & Hepatology, Humanity &<br />

Health Medical Group, Hong Kong, China; 3 Department of Infectious<br />

Diseases and Liver Unit, Nanfang Hospital, Southern Medical<br />

University, Guangzhou, China<br />

Background and Aims: China has about 50 million individuals<br />

infected with chronic hepatitis C virus (HCV), among which<br />

genotype 1 accounts for almost 60%. New interferon (IFN)-<br />

free regimens to treat HCV genotype 1 are more effective than<br />

the standard of care (SOC) but substantially more expensive.<br />

Sofosibuvir (SOF), for example, has demonstrated high efficacy<br />

in patients with HCV genotype 1, 2, 3, or 4 infection. We<br />

aimed to evaluate the cost-effectiveness of 12-week sofosibuvir<br />

(Sovaldi®, Gilead) plus daclatasvir (Daklinza®, BMS) (SOF/<br />

DCV) compared to SOC in treatment-naïve HCV genotype 1b<br />

patients in real life practice in Chinese. Methods: A Markov<br />

transition model was constructed based on the natural history<br />

of HCV infection from the perspective of third-party payer. A<br />

cohort of patients were stratified by sex and cirrhosis stages<br />

and followed from age 50 to simulate the lifetime cost of HCV<br />

infection and IFN-free treatment as well as the health outcome<br />

in quality adjusted life years (QALYs). SVR rates were from<br />

real life data. Utility scores for QALYs and transition rates<br />

among different stages of HCV were from literature. Direct<br />

medical costs in China were used and costs were in US$. Both<br />

costs and health outcome were discounted at 3%. One-way<br />

sensitivity analysis and second order Monte Carlo simulation<br />

were undertaken to assess parameter uncertainty. Results: The<br />

SOF/LDV regimen added 0.84 QALY compared to SOC, at<br />

an incremental cost effectiveness ratio (ICER) of US$68,397<br />

per additional QALY. The ICERs ranged from US$108,672 to<br />

US$3,280 with respect to sex and presence of cirrhosis. The<br />

ICERs were lower in patients with cirrhosis than those without<br />

cirrhosis. One-way sensitivity analysis showed that the results<br />

were most sensitive to utility scores after successful treatment,<br />

cost of SOF/LDV treatment, discount rate, and transition probability<br />

of decompensate cirrhosis to death status. At a willingness<br />

to pay threshold of US$40,842 (6 times GDP per capita<br />

in China) per QALY, SOF/LDV regimen had a 58% probability<br />

of being cost-effectiveness. Threshold analysis showed that the<br />

cost of SOF/LDV treatment needs to be reduced by at least<br />

25% to make the regimen cost-effective in Chinese patients.<br />

Conclusion: Though the efficacy is significantly improved using<br />

IFN-free regimen compared to SOC, the IFN-free regimen is<br />

not cost-effective in most Chinese HCV genotype 1b patients<br />

mainly due to the high cost of IFN-free regimen and relatively<br />

low GDP per capital in China. Reduce the cost and shorten the<br />

duration of treatment may improve the cost-effectiveness in the<br />

future.<br />

Disclosures:<br />

George K. Lau - Consulting: Roche, Novartis, Roche, Novartis, Roche, Novartis,<br />

Roche, Novartis<br />

The following authors have nothing to disclose: Bing Li, Jing Chen, Cheng Wang,<br />

Vanessa Wu, April Wong, Qing Shao, Dong Ji, Fan Li, Yudong Wang, Xiaoxia<br />

Niu, Jialiang Liu, Wucai Yang, Yiming Fu, Guofeng Chen<br />

1537<br />

Cost-effectiveness analysis of sofosbuvir-based regimens<br />

for chronic hepatitis C virus genotype 1 infection<br />

in China<br />

Fan Li 1 , Guofeng Chen 1 , Qing Shao 1 , Dong Ji 1 , Bing Li 1 , Zhongbin<br />

Li 1 , Jialiang Liu 1 , Chunxiao Liu 1 , Tingting Wang 1 , Qiaomin<br />

Wang 1 , Jing Chen 2 , Vanessa Wu 2 , April Wong 2 , Cheng Wang 2 ,<br />

Yudong Wang 2 , George K. Lau 1,2 ; 1 Liver Fibrosis Diagnosis and<br />

Treatment Center, 302 Hospital, Beijing, China; 2 Gastroenterology<br />

& Hepatology, Humanity & Health Medical Group, Hong<br />

Kong, China<br />

Background and Aims: The nucleotide polymerase inhibitor<br />

sofosbuvir (SOF) has demonstrated high sustained virologic<br />

response (SVR) rates in treatment-experienced HCV infection<br />

but very costly. We aimed to evaluate the cost-effectiveness of<br />

sofosbuvir (SOF)-based regimens i.e. 12-week SOF (Sovaldi®,<br />

Gilead) plus peginterferon/ribavirin (SOF+PEG/RBV) and<br />

12-week SOF (Sovaldi®, Gilead) plus Daclatasvir (Daklinza®,<br />

BMS) (SOF+DCV) compared to SOC in treatment-experienced<br />

HCV genotype 1b patients in Chinese. Methods: A Markov<br />

model with a lifetime, third-party payer’s perspective was constructed<br />

to simulate the disease progression of a cohort of 50<br />

year old HCV patients. The patients were stratified by sex and<br />

cirrhosis status. Outcome measures included quality adjusted<br />

life years (QALYs), lifetime costs of HCV infection and incremental<br />

cost-effectiveness ratios (ICERs). Both cost and health<br />

outcome were discounted at 3%. The robustness of the results<br />

was tested by one-way sensitivity analysis and probabilistic<br />

sensitivity analysis. Results: Compared to SOC, SOF+DCV<br />

were cost-effective in patients with cirrhosis with an ICER of<br />

US$13,304 (15,842 and 10,233 for males and females<br />

respectively)(Table). SOF-based treatments were not cost-effective<br />

in patients without cirrhosis. The cost-effectiveness of SOFbased<br />

treatments was better in females than males across all<br />

types of patients. The results were sensitive to post SVR utility,<br />

costs of SOF-based treatments, and discount rate. At the WTP<br />

threshold of US$20,421 (3 times of GDP per capita in China),<br />

SOF+PEG/RBV were cost-effective in 88% and 97% non-cirrhotic<br />

and cirrhotic patients; SOF+DCV were cost-effective in<br />

40% and 98% non-cirrhotic and cirrhotic patients. Conclusion:<br />

In treatment-experienced HCV patients, SOF-based treatments<br />

may be a cost-effective alternative to SOC depending on the<br />

costs of SOF-based drugs.


960A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

George K. Lau - Consulting: Roche, Novartis, Roche, Novartis, Roche, Novartis,<br />

Roche, Novartis<br />

The following authors have nothing to disclose: Fan Li, Guofeng Chen, Qing<br />

Shao, Dong Ji, Bing Li, Zhongbin Li, Jialiang Liu, Chunxiao Liu, Tingting Wang,<br />

Qiaomin Wang, Jing Chen, Vanessa Wu, April Wong, Cheng Wang, Yudong<br />

Wang<br />

1538<br />

The Value of Cure Associated with Treating Treatment-naïve<br />

(TN) Chronic Hepatitis C (CH-C) Genotype 1<br />

(GT1): Are the New All Oral Regimens a Good Value to<br />

Society?<br />

Zobair M. Younossi 1,2 , Haesuk Park 3 , Douglas Dieterich 4 , Sammy<br />

Saab 5 , Aijaz Ahmed 6 , Stuart C. Gordon 7 ; 1 Center For Liver<br />

Disease, Department of Medicine, Inova Fairfax Hospital, Falls<br />

Church, VA; 2 Betty and Guy Beatty Center for Integrated Research,<br />

Inova Health System, Falls Church, VA; 3 University of Florida,<br />

Gainesville, FL; 4 Icahn School of Medicine at Mount Sinai, New<br />

York City, NY; 5 University of California Los Angeles, Los Angeles,<br />

CA; 6 Stanford University, Standford, CA; 7 Henry Ford Hospital,<br />

Detroit, MI<br />

Background and Aim: The approval of all-oral regimens has<br />

led to substantially high cure rates. Our aim was to assess the<br />

value of cure to society by treating HCV GT1. Methods: A Markov<br />

model for HCV GT1 projected long-term health outcomes,<br />

life-years (LYs), and quality-adjusted life-years (QALYs) gained<br />

for a lifetime horizon. The model compared current therapies in<br />

the US [2nd generation triple (sofosbuvir+PR & simeprevir+PR)<br />

and all-oral therapy treatments (ledipasvir/sofosbuvir and<br />

ombitasvir+paritaprevir/ritonavir+dasabuvir±ribavirin] with<br />

no treatment option. Sustained virologic response (SVR) rates<br />

were based on Phase III clinical trials. We assumed that 80%<br />

and 95% of HCV GT1 patients are eligible for 2nd generation<br />

triple and all-oral regimens. Transition probabilities, utility and<br />

mortality were based on literature review and consensus by<br />

hepatologists. Results: In 2015, 1.11 million individuals were<br />

TN with HCV-GT1. The model estimated that treating all eligible<br />

HCV GT1 patients with 2nd generation triple and all-oral<br />

therapies can result in 1.8 million and 2.9 million additional<br />

QALYs gained. Using a threshold of $50,000 to $150,000<br />

for the value of a QALY, these regimens led to savings of $91-<br />

$274 billion and $144-$432 billion. Also, the costs of these<br />

regimens were $80 billion and $93 billion. The value of cure<br />

was assessed by subtracting the costs of the regimens from the<br />

economic gains associated with the value of QALYs. The value<br />

of cure with 2nd generation triple and all oral regimens was<br />

estimated at $12-$194 billion and $51-$340 billion. Each<br />

HCV GT1 patient cured with 2nd generation triple and all oral<br />

regimens could save the U.S. between $10,381-$174,473<br />

and $46,141-$305,021. Conclusions: The recent evolution<br />

of regimens for HCV GT1 has increased efficacy and value of<br />

cure. Curing all eligible HCV-GT1 patients with all-oral regimen<br />

leads to substantial savings to the U.S. society.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Haesuk Park - Consulting: Gilead Science<br />

Douglas Dieterich - Advisory Committees or Review Panels: Gilead, BMS, Abbvie,<br />

Janssen, Merck, Achillion<br />

Sammy Saab - Advisory Committees or Review Panels: BMS, Gilead, Merck,<br />

Janssen; Grant/Research Support: Gilead; Speaking and Teaching: BMS, Gilead,<br />

Merck, Janssen, Salix, Onyx, Bayer, Janssen; Stock Shareholder: Achillion,<br />

Johnson and Johnson, BMS, Gilead<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

1539<br />

Rethinking pricing policy of direct antiviral agents to<br />

cure chronic hepatitis C infection at a population-level<br />

Francois Girardin, Nathalie Vernaz, Nicolas Goossens, Francesco<br />

Negro; Geneva University Hospital, Geneva, Switzerland<br />

Background Although cost-effectiveness of new direct antiviral<br />

agents for chronic hepatitis C viral infection (HCV) is a<br />

prerequisite for reimbursement, a population-level cure would<br />

induce staggering upfront costs. We investigated the correlation<br />

between sustained virological response (SVR) and drug<br />

costs. Methods We derived medication costs from Swiss public<br />

reimbursement prices and extracted published SVR data of<br />

currently available HCV therapies (ribavirin, pegylated interferon,<br />

simeprevir, boceprevir, telaprevir, sofosbuvir, ledipasvir,<br />

ombitasvir, paritaprevir, dasabuvir). We stratified cost-analysis<br />

by HCV genotype and presence/absence of cirrhosis (in treatment<br />

naïve patients). We measured the correlation between<br />

the costs and SVR. All cost results are reported in USD or CHF,<br />

as with the rounded exchange rate of 1 USD=1 Swiss franc<br />

might be considered (2015). Results For both patient groups,<br />

we found a correlation between SVR and cost, where R 2 was<br />

higher for naive cirrhotic patients (67.8%) compared to naive<br />

the non-cirrhotic patient group (56.4%) indicating that the<br />

more one spends on a treatment strategy, the higher the SVR.<br />

The costs extended between USD12,690 and USD133,750<br />

per SVR (Fig.1). Discussion There is a significant correlation<br />

between the rate of SVR and the costs of treatment strategy.<br />

Payment per patient recovery, regardless of dose and treatment<br />

duration, including risk-sharing with drug manufacturers<br />

would allow better budget control and long-term projections.<br />

Alternatively, pricing volume per year would also minimize<br />

resources consumption and maximize health benefits and allow<br />

the manufacturer to provide large amount of drug before the<br />

era of DAA generics.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 961A<br />

(IQR 24-52 months) with negative HBV-DNA. HCC incidence<br />

was higher in cirrhotic patients (global – 15.6%; 3.8% per<br />

year) than in patients without cirrhosis (global 1%; 0.2% per<br />

year), p


962A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

intrahepatic virological and inflammatory parameters (Median<br />

of HBV tDNA: 7.90 vs. 7.45Log 10<br />

copies/10 6 cells, P = 0.593;<br />

Median of cccDNA: 6.1 vs. 5.0Log 10<br />

copies/10 6 cells, P =<br />

0.052; Percentage of inflammation ≥ G2: 77.8% (21/27)<br />

vs. 84.6% (11/13), P = 0.933). Furthermore, comparable<br />

levels of intrahepatic HBV tDNA (Median: 7.89 vs. 7.90Log-<br />

10 copies/106 cells, P = 0.395) and significantly lower cccDNA<br />

(Median: 4.75 vs. 6.82Log 10<br />

copies/10 6 cells, P = 0.003) were<br />

presented in the patients with undetectable serum HBV DNA<br />

and normal ALT compared to the patients wits undetectable<br />

serum HBV DNA and abnormal ALT, and 61.5% (8/13) of the<br />

patients with undetectable serum HBV DNA and normal ALT<br />

were found intrahepatic inflammation ≥ G2. Conclusion: More<br />

than 60% HBV-related compensated cirrhosis patients with<br />

serum HBV DNA undetectable or at low levels and normal ALT<br />

had active intrahepatic inflammation and HBV replication, thus<br />

the administration of antiviral therapy might be strengthened.<br />

Disclosures:<br />

The following authors have nothing to disclose: Taotao Yan, Xiaoying Sha, Jing<br />

Wang, Li Jin, Yu Zhang, Ruitian Yi, Furong Cao, Yuan Yang, Jinfeng Liu, Yingli<br />

He, Tianyan Chen, Yingren Zhao<br />

1542<br />

Comparing antiviral efficacy of Tenofovir monotherapy<br />

and Tenofovir based combination therapy in antiviral<br />

resistant chronic hepatitis B: A interim result of multicenter<br />

cohort study<br />

Kyu sik Chung 1 , Beom Kyung Kim 1 , Ki Tae Yoon 2 , Hana Park 3 ,<br />

Chang Wook Kim 4 , Jin-Woo Lee 5 , Young Seok Kim 6 , Jung Il Lee 7 ,<br />

Jun Yong Park 1 , Seung Up Kim 1 , Do Young Kim 1 , Sang Gyune<br />

Kim 6 , Young-Joo Jin 5 , Hee Yeon Kim 4 , Seong Gyu Hwang 3 , Mong<br />

Cho 2 , Kwan Sik Lee 7 , Kwang-Hyub Han 1 , Sang Hoon Ahn 1 ;<br />

1 Department of Internal Medicine, Yonsei University College of<br />

Medicine, Seoul, Korea (the Republic of); 2 Department of Internal<br />

Medicine, Pusan National University School of Medicine, Yangsan,<br />

Korea (the Republic of); 3 Department of Internal Medicine,<br />

CHA Bundang Medical Center, Bundang, Korea (the Republic of);<br />

4 Department of Internal Medicine, The Catholic University of Korea<br />

College of Medicine, Seoul, Korea (the Republic of); 5 Department<br />

of Internal Medicine, Inha University College of Medicine, Incheon,<br />

Korea (the Republic of); 6 Department of Internal Medicine, Soonchunhyang<br />

University College of Medicine, Cheonan, Korea (the<br />

Republic of); 7 Department of Internal Medicine, Gangnam Severance<br />

Hospital, Seoul, Korea (the Republic of)<br />

Background: We aimed to compare antiviral efficacy between<br />

tenofovir (TDF) monotherapy and TDF-based combination therapy<br />

as a rescue therapy for chronic hepatitis B with resistance<br />

to antiviral agents. Methods: A total of 847 consecutive CHB<br />

patients treated with TDF monotherapy or TDF-based combination<br />

therapy (TDF plus entecavir, lamivudine or telbivudine)<br />

as a rescue therapy for resistance to antiviral agents were<br />

analyzed. Complete virological response (CVR) and biochemical<br />

response were defined as undetectable serum HBV-DNA<br />

(detection limit 20 IU/mL) and normalization of serum alanine<br />

aminotransferase (ALT) level, respectively. Results: At the time<br />

of rescue therapy, the median serum HBV DNA and ALT level<br />

were 3.36 log IU/mL and 28 IU/mL, respectively. During the<br />

follow-up (median 18.3 month), cumulative rates of CVR at<br />

24 months among patients with resistance to lamivudine only<br />

(n=474) were similar between those with TDF monotherapy<br />

and those with TDF-based combination therapy (95.6% vs.<br />

92.4%, respectively). Similar results (95.5% vs. 94.0%, respectively)<br />

were maintained in patients with multidrug-resistance<br />

(n=373) (resistance to lamivudine and entecavir [n=189], resistance<br />

to lamivudine and adefovir [n=137], and resistance to<br />

lamivudine, adefovir, and entecavir [n=47]) (all p>0.05). In<br />

HBeAg- positive patients, cumulative HBeAg seroconversion<br />

rate at 24 months was comparable between those with TDF<br />

monotherapy and those with TDF-based combination therapy<br />

(7.6% vs. 11.8%, P=0.258). Conclusions: TDF monotherapy<br />

showed similar antiviral efficacy compared with TDF-based<br />

combination therapy as a rescue therapy, even in patients with<br />

multi-drug resistance. Further <strong>studies</strong> with a longer follow-up<br />

are required to validate these results<br />

Disclosures:<br />

Ki Tae Yoon - Grant/Research Support: Handok; Speaking and Teaching: Gilead,<br />

BMS, MSD, Roche, GSK<br />

Chang Wook Kim - Consulting: Gilead Korea, MSD; Grant/Research Support:<br />

BMS Korea, Daewoong, Handok, Pharmicell, Pharmaking, Gilead Korea; Speaking<br />

and Teaching: BMS Korea, Daewoong<br />

The following authors have nothing to disclose: Kyu sik Chung, Beom Kyung Kim,<br />

Hana Park, Jin-Woo Lee, Young Seok Kim, Jung Il Lee, Jun Yong Park, Seung Up<br />

Kim, Do Young Kim, Sang Gyune Kim, Young-Joo Jin, Hee Yeon Kim, Seong Gyu<br />

Hwang, Mong Cho, Kwan Sik Lee, Kwang-Hyub Han, Sang Hoon Ahn<br />

1543<br />

Hepatitis B virus infection in irregular and refugee<br />

migrants in Naples, Italy<br />

Nicola Coppola 1 , Loredana Alessio 1 , Luciano Gualdieri 2 , Mariantonietta<br />

Pisaturo 3 , Caterina Sagnelli 4 , Nunzio Caprio 5 , Carmine<br />

Minichini 1 , Mario Starace 1 , Giuseppe Signoriello 1 , Giuseppe<br />

Pasquale 1 , Evangelista Sagnelli 1 ; 1 Mental Health and Public Medicine,<br />

Second University of Naples, Naples, Italy; 2 Medical Center,<br />

Centro per la Tutela della Salute degli Immigrati, Naples, Italy;<br />

3 Medical Center, Centro di Accoglienza “La tenda di Abramo”,<br />

Caserta, Italy; 4 Medical Center, Centro Sociale ex Canapificio,<br />

Caserta, Italy; 5 Medical center, Centro Suore Missionarie della<br />

Carità, Naples, Italy<br />

Aim: to define the characteristic of HBV infection in a cohort of<br />

irregular or refugee immigrants living in Italy, a country with an<br />

HBsAg prevalence nearly 1% and a HBV universal vaccination<br />

covering subjects under 35. Methods: A screening for HBV,<br />

HCV and HIV infections was offered free of charge and of<br />

bureaucratic procedures to 1,254 illegal or refugee immigrants<br />

living from 5 years or more in Naples or in its surroundings<br />

(Italy). Of these 1,254, 1,212 (96.6%), 831 irregular and<br />

381 refugees, accepted to be screened at one of 4 first-level<br />

clinical centers operating for years in this setting, with proven<br />

experience in clinical, psychological and legal management<br />

of vulnerable groups. The median age of screened subjects<br />

was 32 years, range 12-74, 75.2% were males, 52% came<br />

from Sub-Saharan Africa, 18% from Eastern-Europe, 13% from<br />

Indo-Pakistan Area, 7% from Northern-Africa and the remaining<br />

10% from others countries. Positive subjects ware further<br />

investigated at two third-level liver units. Results: One-hundred<br />

sixteen migrants (9.6%) were HBsAg positive, 490 (40.4%)<br />

HBsAg negative/anti-HBc positive and 606 (50%) sero-negative<br />

for both. All positive subjects ignored their serological<br />

status. A high HBsAg sero-prevalence was found in migrants<br />

from sub-saharan Africa (13.8%) and an intermediate one in<br />

those from Eastern Europe (6%), Northern Africa (3.4%) and<br />

Indo-Pakistan area (2.8%). A logistic regression analysis identified<br />

as factors independently associated with HBV infection<br />

(both HBsAg positive and HBsAg negative/anti-HBc positive)<br />

the male gender (p=0.002), the sub-Saharan African origin<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 963A<br />

of infectious diseases: only 4 (5.1%) were HBeAg positive and<br />

2 (2.5%) anti-Delta positive. According the EASL criteria, 58<br />

(73.4%) were considered asymptomatic carriers, 17 (21.5%)<br />

with chronic hepatitis, 2 with liver cirrhosis and 2 with liver<br />

cirrhosis plus HCC. Conclusions: The HBsAg sero-prevalence is<br />

very high in the immigrants investigated, particularly in those<br />

from Sub-Saharan Africa who are introducing HBV-genotype E<br />

in Italy. The data suggest a stronger action of the Healthcare<br />

Authorities to limit the diffusion of HBV infection in Italy by<br />

supporting screenings, educational programs, HBV vaccination<br />

and treatments in favor of migrant populations.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nicola Coppola, Loredana<br />

Alessio, Luciano Gualdieri, Mariantonietta Pisaturo, Caterina Sagnelli, Nunzio<br />

Caprio, Carmine Minichini, Mario Starace, Giuseppe Signoriello, Giuseppe<br />

Pasquale, Evangelista Sagnelli<br />

1544<br />

Four years real-life experience with antiviral adherence<br />

in chronic hepatitis B infection in a tertiary university<br />

hospital<br />

Rodrigo M. Abreu 1,2 , Leda C. Bassit 3 , Sijia Tao 3 , Yong Jiang 3 ,<br />

Denise Paranaguá-Vezozzo 1 , Raymond F. Schinazi 3 , Flair J. Carrilho<br />

1 , Suzane K. Ono 1 ; 1 Gastroenterology, University of São<br />

Paulo School of Medicine, São Paulo, Brazil; 2 Division of Pharmacy,<br />

Clinical Hospital, University of São Paulo School of Medicine,<br />

São Paulo, Brazil; 3 Pediatrics, Emory University School of<br />

Medicine, and Veterans Affairs Medical Center, São Paulo, Brazil<br />

BACKGROUND: Evidence shows that antiviral treatment for<br />

chronic hepatitis B (CHB) infection suppress viral load, but does<br />

not eradicate hepatitis B virus (HBV). Among the factors directly<br />

linked to therapeutic success is adherence to treatment. The purpose<br />

of this study was to evaluate the causes of non-response<br />

to antiviral treatment in chronic HBV infection. METHODS:<br />

A prospective cohort study with CHB treated with adefovir,<br />

entecavir, lamivudine and / or tenofovir was performed in<br />

a Brazilian reference tertiary hospital. Treatment adherence<br />

was evaluated by a validate questionnaire named CEAT-HBV<br />

within December 2010 and August 2011. HBV drug resistance<br />

mutation and single-dose pharmacokinetics were determined<br />

by PCR sequencing and LC-MS respectively. The IRB approved<br />

the research study. RESULTS: Of the 183 individuals enrolled,<br />

CEAT-HBV identified 104 (57%) on HBV treatment adherence.<br />

Among the 79/183 (43%) individuals with non-adherence<br />

to antiviral treatment, 53/79 (67.1%) were more frequently<br />

viral load positive. The frequencies of HBV genotypes were A<br />

(14%), A1 (34%), A2 (3%), C (18%) and D (14%). However,<br />

38% (70/183) had positive viral loads suggesting non-response<br />

to antiviral treatment. In the adherence group (n = 14)<br />

with positive viral load, 9 (64%) individuals had HBV with drug<br />

resistance mutations. In the positive viral load and non-adherence<br />

group (n = 40), 33 (83%) individuals had drug resistance<br />

HBV mutations. Sixteen out of 70 HBV positive plasma samples<br />

could not be amplified. The main factors involved with<br />

non-response to antiviral treatment were drug resistance mutations<br />

in HBV (51%), non-adherence without drug resistance<br />

mutations in HBV (37%), short treatment duration (8%) and<br />

non-determined (4%). The single-dose pharmacokinetics was<br />

used to confirm antiviral non-adherence. Two years after the<br />

first evaluation, the CEAT-HBV showed that 106/143 (74.1%)<br />

of individuals were on treatment adherent. However, 21.2%<br />

(40/183) individuals could not be evaluated and excluded.<br />

The main reasons for exclusion were death (20/183), 10 out<br />

20 deaths due to hepatocellular carcinoma, loss to follow up<br />

(11/183) and others (9/183). It is worth noting that these 10<br />

dead persons were adherent to antiviral treatment and had<br />

undetectable HBV DNA. Conclusion: Non-adherence was high<br />

among CHB individuals receiving antiviral treatment. Drug<br />

resistance mutations and non-adherence to anti-HBV therapy<br />

were the main cause of treatment failure in CHB individuals.<br />

This study demonstrates the urgency to implement more effective<br />

procedures to improve adherence to antiviral treatment for<br />

chronic hepatitis B.<br />

Disclosures:<br />

Raymond F. Schinazi - Board Membership: Cocrystal Pharma, Inc.; Stock Shareholder:<br />

Cocrystal Pharma, Inc.<br />

The following authors have nothing to disclose: Rodrigo M. Abreu, Leda C. Bassit,<br />

Sijia Tao, Yong Jiang, Denise Paranaguá-Vezozzo, Flair J. Carrilho, Suzane<br />

K. Ono<br />

1545<br />

Factors Associated with the Development of Hepatocellular<br />

Carcinoma after HBeAg Seroclearance<br />

James Fung, Ching-Lung Lai, Danny Wong, Wai-Kay Seto, Man-<br />

Fung Yuen; University of Hong Kong, Hong Kong, China<br />

Background: The role of core promoter mutation, precore mutation,<br />

and genotype in the development of hepatocellular carcinoma<br />

(HCC) in patients after HBeAg seroclearance is unclear.<br />

Methods: The CHB E-Seroclearance Study (C.H.E.S.S.) cohort<br />

follows up a large population of patients with documented<br />

HBeAg seroclearance. Surveillance for HCC was performed at<br />

regular intervals with alpha-fetoprotein and ultrasonography.<br />

This substudy includes patients who had precore and core promoter<br />

mutation and genotype determined at 3 to 24 months<br />

after the time of HBeAg seroclearance. Results: A total of 293<br />

CHB subjects were included with a median follow up of 124<br />

months (range, 15 to 254) after HBeAg seroclearance, totaling<br />

3,200 patient-years. The median age of HBeAg seroclearance<br />

was 34 years, of which 164 (56%) were male. The presence<br />

of core promoter and precore mutation was observed in 181<br />

(61.8%) and 249 (85.0%) patients respectively. The majority of<br />

patients were of genotype C (88.1%). The cumulative rates of<br />

HCC at 5, 10 and 20 years after HBeAg seroclearance were<br />

3.5%, 6.9%, and 8.8% respectively. There were a total of 19<br />

cases of HCC, with a median time to HCC development (from<br />

the time of HBeAg seroclearance) of 57 months (range, 17<br />

to 192). The median age of HCC development was 50 years<br />

(range, 33 to 72). Males had a significantly higher rate of HCC<br />

compared to females (12.5% vs 3.8% at 20 years respectively,<br />

p=0.037). The median age of HBeAg seroclearance was significantly<br />

higher for those who subsequently developed HCC<br />

compared to those without HCC (43 vs 34 years respectively,<br />

p


964A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Wai-Kay Seto - Advisory Committees or Review Panels: Gilead Science; Speaking<br />

and Teaching: Bristol-Myers Squibb<br />

Man-Fung Yuen - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Bristol-Myers Squibb, Pfizer, GlaxoSmithKline, Bristol-Myers Squibb, Pfizer,<br />

GlaxoSmithKline, Bristol-Myers Squibb, Pfizer, GlaxoSmithKline, Bristol-Myers<br />

Squibb, Pfizer; Grant/Research Support: Roche, Bristol-Myers Squibb,<br />

GlaxoSmithKline, Gilead Science, Roche, Bristol-Myers Squibb, GlaxoSmith-<br />

Kline, Gilead Science, Roche, Bristol-Myers Squibb, GlaxoSmithKline, Gilead<br />

Science, Roche, Bristol-Myers Squibb, GlaxoSmithKline, Gilead Science<br />

The following authors have nothing to disclose: James Fung, Danny Wong<br />

1546<br />

Assessing the impact of hepatitis B immune globulin<br />

(HBIG) on responses to the hepatitis B vaccine during<br />

co-administration<br />

Qingwen Cui 1 , Iryna Zubkova 1 , Trudy Murphy 2 , Sarah F. Schillie 2 ,<br />

Marian E. Major 1 ; 1 CBER/FDA, Alexandria, VA; 2 Division of Viral<br />

Hepatitis, Centers for Disease Control and Prevention, Atlanta, GA<br />

Background: Administration of hepatitis B vaccine (HepB) and<br />

hepatitis B immune globulin (HBIG) is recommended for infants<br />

born to hepatitis B virus (HBV) carrier mothers and for exposed<br />

persons with no known HBV protection. The recommended prophylaxis<br />

for infants typically consists of HepB and HBIG at birth<br />

followed by vaccination at 1 and 6 months; recommendations<br />

specify administering HepB and HBIG via the intramuscular<br />

(i/m) route at different sites. However, there are documented<br />

instances of HepB and HBIG being administered at the same<br />

site (i.e., in the same limb) and the impact of this on immune<br />

responses to the vaccine remains unanswered. Our goal was<br />

to address this issue through a series of <strong>studies</strong> immunizing<br />

mice with a licensed HepB (GSK Engerix-B) alone or in combination<br />

with HBIG (Nabi-HB, Biotest Pharmaceuticals). BALB/c<br />

mice have been routinely used for in vivo potency assays for<br />

HepB and provide a viable model to study responses. Methods:<br />

Six week old mice (15/group) were immunized i/m with<br />

HepB and/or HBIG at a 1:1 v/v ratio at 0, 4, and 16 weeks.<br />

Groups consisted of: 1) HepB alone 2) HepB in one leg, HBIG<br />

in a different leg (HepB boost at 4 and 16 wks) 3) HepB and<br />

HBIG at the same site in the same leg (HepB boost at 4 and 16<br />

wks) 4) HepB and HBIG mixed prior to administration (O/N<br />

+4C) (HepB boost at 4 and 16 wks) 5) HBIG alone (no HepB<br />

boost) 6) Untreated. Animals were bled at 2, 6 and 26 weeks<br />

and sera were assessed for anti-HBs titers using a commercial<br />

assay. Results: At week 26 we observed 100% seroconversion<br />

(>10mIU/mL) in all groups that received HepB, regardless<br />

of how HepB was administered with HBIG. However, we<br />

observed significantly lower anti-HBs titers at 26 weeks when<br />

HBIG and HepB were administered in the same limb (Gp3)<br />

or after incubation overnight (Gp4) compared to HepB alone<br />

(Gp1) (p=0.01 and p=0.02, respectively). At week 2 post<br />

inoculation, prior to the development of responses to vaccine,<br />

circulating levels of HBIG were lower in Gp3 mice (p=0.07;<br />

87% seropositive) and Gp4 mice (p=0.001; 66% seropositive)<br />

that received HepB and HBIG in the same limb compared to<br />

those that received HBIG and HepB in different limbs (Gp2).<br />

We observed similar results following inoculation of 3 week<br />

old mice to simulate neonates. Conclusions: Administration of<br />

HepB and HBIG in the same limb does not impact the long-term<br />

response to HepB in either adult or young mice. However,<br />

this incorrect method of administration can impact the initial<br />

levels of circulating HBIG prior to the development of adaptive<br />

immune responses which in some individuals could result in no<br />

protective antibody levels during the first weeks after exposure<br />

to HBV.<br />

Disclosures:<br />

The following authors have nothing to disclose: Qingwen Cui, Iryna Zubkova,<br />

Trudy Murphy, Sarah F. Schillie, Marian E. Major<br />

1547<br />

End-of-therapy Serum Gradient of Hepatitis B Surface<br />

Antigen Predicts Off-therapy Durability after Discontinuation<br />

of Nucleos(t)ide Analogues<br />

Yao-Chun Hsu 1,2 , Chun-Ying Wu 3 , Chi-Yang Chang 1 , Ming-Shiang<br />

Wu 7 , Jia-Horng Kao 4 , Tzeng-Huey Yang 5 , Hsu-Wei Hung 8 ,<br />

Jaw-Town Lin 6 ; 1 Division of Gastroenterology, E-Da Hospital, Kaohsiung,<br />

Taiwan; 2 School of Medicine, I-Shou University, Kaohsiung<br />

City, Taiwan; 3 Division of Gastroenterology, Taichung Veterans<br />

General Hospital, Taichung City, Taiwan; 4 Graduate Institute of<br />

Clinical Medicine, National Taiwan University, Taipei City, Taiwan;<br />

5 Division of Gastroenterology, Lotung Poh-Ai Hospital, Yilan<br />

County, Taiwan; 6 School of Medicine, Fu Jen Catholic University,<br />

New Taipei City, Taiwan; 7 Department of Internal Medicine,<br />

National Taiwan University, Taipei City, Taiwan; 8 Taipei Institute<br />

of Pathology, Taipei City, Taiwan<br />

Background and Aims: Relapse of chronic hepatitis B (CHB)<br />

is common but remains unpredictable after discontinuation of<br />

nucleos(t)ide analogues (NAs). This multicenter prospective<br />

research aimed to explore the association between serum gradient<br />

of hepatitis B surface antigen (HBsAg) and off-therapy<br />

durability following NA cessation. Methods: From July 2011<br />

through December 2014, 228 consecutive CHB patients who<br />

were about to discontinue NAs were screened from 3 teaching<br />

hospitals in Taiwan. Those who had achieved undetectable<br />

viral DNA after a minimum of 3-year therapy were included.<br />

Participants were monitored until February 2015 for occurrence<br />

of clinical relapse, stringently defined as viral DNA >2,000IU/<br />

mL plus ALT >2 folds upper normal limit for ^3 months, and<br />

viral relapse, simply defined as viral DNA>2,000 IU/mL. Incidences<br />

of relapse were evaluated by the Kaplan Meier method<br />

and risk predictors determined by the Cox proportional hazard<br />

analysis. Results: Among 158 eligible patients observed for a<br />

mean duration of 14.6 (range, 3-42.2) months, clinical and<br />

viral relapses occurred in 44 and 103 patients, respectively,<br />

corresponding to 2-year cumulative incidences of 46.2% (95%<br />

CI, 35.2-58.8%) and 81.6% (95% CI, 72.9-88.8%), respectively.<br />

In 119 patients with negative HBeAg at NA cessation,<br />

end-of-therapy HBsAg was associated with both clinical and<br />

viral relapse in a dose-response manner (Ptrend=0.02 for clinical<br />

and 0.0005 for virological relapse). Patients whose serum<br />

HBsAg fell below 100 IU/mL did not develop clinical relapse<br />

but could experience viral relapse with a cumulative incidence<br />

of 33.4% (95% CI, 14.6-64.7%) at 2 years. In those with<br />

HBsAg


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 965A<br />

Disclosures:<br />

Yao-Chun Hsu - Speaking and Teaching: GlaxoSmithKline, Novartis, Bristol-Myers<br />

Squibb Company, Harvester Trading Company<br />

The following authors have nothing to disclose: Chun-Ying Wu, Chi-Yang Chang,<br />

Ming-Shiang Wu, Jia-Horng Kao, Tzeng-Huey Yang, Hsu-Wei Hung, Jaw-Town<br />

Lin<br />

1548<br />

Liver fibrosis progression, new-onset metabolic syndrome<br />

and risk of hepatocellular carcinoma in chronic<br />

hepatitis B – an 8-year prospective cohort study of<br />

1,178 patients with paired transient elastography<br />

examination<br />

Grace LH Wong, Henry Lik-Yuen Chan, Angel ML Chim, Vincent<br />

W. Wong; Institute of Digestive Disease, The Chinese University of<br />

Hong Kong, Shatin, Hong Kong<br />

Background: New-onset metabolic syndrome (nMES) was<br />

recently found to increase the risk of liver fibrosis progression<br />

in patients with chronic hepatitis B (CHB). We aimed to investigate<br />

the effect of nMES on the risk of hepatocellular carcinoma<br />

(HCC) in CHB patients. Methods: 1,466 CHB patients<br />

underwent liver stiffness measurement (LSM) by transient elastography<br />

in 2006-2008; 1,178 patients had 2 nd LSM in 2010-<br />

2012. Liver fibrosis progression was defined as an increase in<br />

LSM ≥30% at the second assessment. Patients were prospectively<br />

followed up with regular ultrasonographic examinations<br />

and 6-monthly alpha-fetoprotein for HCC surveillance since the<br />

first LSM. The impact of liver fibrosis progression and nMES<br />

on HCC development was evaluated. Results: At 1 st LSM, the<br />

mean age of these 1,178 patients was 46±11 years, 63%<br />

were males, body-mass index (BMI) 23.2±3.3kg/m 2 , serum<br />

alanine aminotransferase (ALT) 63±40IU/l, 26% HBeAg positive,<br />

HBV DNA 4.5±2.0logIU/ml, and LSM 8.0±3.6kPa. At<br />

2 nd LSM, the mean BMI was 23.2±3.5kg/m 2 , ALT 38±35IU/l,<br />

HBV DNA 2.5±1.8logIU/ml, and LSM was 6.2±3.7kPa; 44%<br />

had received antiviral treatments. Metabolic syndrome was<br />

diagnosed in 153 (13%) and 350 (30%) patients at 1 st and<br />

2 nd LSM; 230 (20%) and 33 (3%) patients had nMES and<br />

resolved MES respectively. After an interval of 44±7 months,<br />

155 (13%) patients developed liver fibrosis progression. At<br />

93±9 months from 1 st LSM, 28 patients (2.4%) developed HCC<br />

and 5 patients (0.4%) died. Liver fibrosis progression but not<br />

nMES was the independent risk factor of HCC. The risk of HCC<br />

in patients with or without liver fibrosis progression at 3 years<br />

after 2 nd LSM was 4.5% (95% CI 1.1%-9.9%) and 1.6% (95%<br />

CI 1.0%-2.2%) respectively (P=0.008). The 3-year risk of HCC<br />

in patients with nMES, compared to those with no MES at both<br />

time points, was 1.3% (95% CI 0.1%-2.5%) vs. 2.3% (95% CI<br />

1.3%-3.3%) respectively (P=0.28). Conclusions: Liver fibrosis<br />

progression but not nMES increases the risk of HCC in CHB<br />

patients.<br />

Cumulative probability of HCC in patients with or without liver<br />

fibrosis progression and new-onset MES.<br />

Disclosures:<br />

Grace LH Wong - Advisory Committees or Review Panels: Otsuka, Gilead;<br />

Speaking and Teaching: Echosens, Furui, Gilead, Janssen, Bristol-Myers Squibb,<br />

Otsuka, Abbvie<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

Vincent W. Wong - Advisory Committees or Review Panels: AbbVie, Gilead,<br />

Janssen; Consulting: Merck, NovaMedica; Speaking and Teaching: Gilead,<br />

Echosens<br />

The following authors have nothing to disclose: Angel ML Chim<br />

1549<br />

High Rates of Surface Antigen Loss and Low Rates of<br />

Seroreversion in Asian Chronic Hepatitis B Patients<br />

Treated with Oral Antiviral Therapy: Community-Based<br />

Real World Outcomes<br />

Robert J. Wong 1 , My T. Nguyen 2 , Huy N. Trinh 3 , Andrew Huynh 2 ,<br />

Mytop Ly 2 , Huy Nguyen 3 , Khanh Nguyen 3 , Jenny D. Yang 3 , Ruel<br />

Garcia 3 , Brian S. Levitt 3 , Eduardo DaSilvera 3 , Robert Gish 4 ; 1 Gastroenterology<br />

and Hepatology, Alameda Health System - Highland<br />

Hospital, Oakland, CA; 2 Silicon Valley Research Institute, San<br />

Jose, CA; 3 San Jose Gastroenterology, San Jose, CA; 4 Gastroenterology<br />

and Hepatology, Stanford University Medical Center,<br />

Stanford, CA<br />

Background:Chronic hepatitis B virus (HBV) infection is a global<br />

epidemic with over 240 million persons chronically infected<br />

with the majority from Asian-Pacific regions. The most definitive<br />

endpoint of antiviral therapy, surface antigen (HBsAg) loss, is<br />

rare. It is not clear if HBsAg loss correlates well with baseline<br />

patient characteristics, undetectable virus and normalization<br />

of alanine aminotrasferase (ALT) on treatment, or risk of seroreversion<br />

or detectable virus after stopping therapy. Aim:To<br />

evaluate rates of HBsAg loss, baseline correlates of HBsAg<br />

loss, normalization of ALT and undetectable virus, and rates<br />

of HBsAg seroreversion or re-emergence of detectable virus<br />

among chronic HBV patients. Methods:A retrospective cohort<br />

study evaluated 949 adults (age >18) with chronic HBV on<br />

oral antiviral therapy at a large community gastroenterology<br />

clinic from June 1997 to May 2015. The majority (99.6%)<br />

were Asian ethnicity. Rates of HBsAg loss, normalization of<br />

ALT, achieving undetectable virus, and developing surface<br />

antibody (anti-HBs) were stratified by HBeAg status. Following<br />

HBsAg loss, rates of HBsAg seroreversion or re-emergence of<br />

detectable virus were analyzed. Multivariate logistic regression<br />

models evaluated for baseline and on-treatment predictors of<br />

HBsAg loss. Results:With a mean follow up of 86.6 months<br />

and mean duration of treatment of 81.5 months, overall rate<br />

of HBsAg loss was 4.74% (n=45), with similar rates between<br />

HBeAg positive and negative patients (Table). Among HBsAg<br />

loss patients, 30.9% developed anti-HBs, 91.1% achieved<br />

undetectable virus, and 64.4% normalized ALT. Only 2.22%<br />

(n=1) who achieved HBsAg loss had detectable virus, whereas<br />

13.3% (n=6) of HBsAg loss patients had seroreversion of positive<br />

HBsAg within 1-17 months. All 6 patients who seroreverted<br />

had sustained undetectable virus. While there was a<br />

trend towards higher odds of HBsAg loss among men (OR<br />

2.05, 95% CI 0.92-4.59, p=0.08), no significant associations<br />

between baseline characteristics (age, ALT, platelets, HBeAg,<br />

FIB-4 score, APRI score) and HBsAg loss were seen. Conclusion:Among<br />

a large community-based gastroenterology clinic,<br />

treatment of chronic HBV achieved high rates of HBsAg loss<br />

among Asian patients. While 13.3% of patients experienced<br />

seroreversion, all of these patients remained viral load undetectable,<br />

demonstrating the sustainability of HBsAg loss in most<br />

patients and stable suppression of HBV viral load.


966A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Huy N. Trinh - Advisory Committees or Review Panels: BMS, Gilead; Grant/<br />

Research Support: BMS, Gilead; Speaking and Teaching: BMS, Gilead; Stock<br />

Shareholder: Gilead<br />

Robert Gish - Advisory Committees or Review Panels: Gilead, AbbVie, Arrowhead;<br />

Consulting: Eiger, Isis, Genentech; Speaking and Teaching: Gilead, Abb-<br />

Vie; Stock Shareholder: Arrowhead<br />

The following authors have nothing to disclose: Robert J. Wong, My T. Nguyen,<br />

Andrew Huynh, Mytop Ly, Huy Nguyen, Khanh Nguyen, Jenny D. Yang, Ruel<br />

Garcia, Brian S. Levitt, Eduardo DaSilvera<br />

1550<br />

The role of hepatitis B core-related antigen in predicting<br />

HBV reactivation among HBsAg-negative, anti-HBc-positive<br />

individuals undergoing immunosuppressive therapy<br />

Wai-Kay Seto, Danny Wong, Sze Hang Kevin Liu, James Fung,<br />

Ching-Lung Lai, Man-Fung Yuen; Medicine, The University of Hong<br />

Kong, Queen Mary Hospital, Hong Kong, Hong Kong<br />

Background: Hepatitis B core-related antigen (HBcrAg) detects<br />

an identical amino-acid sequence shared by hepatitis B e<br />

antigen, hepatitis B core antigen and 22kDa precore protein.<br />

Serum HBcrAg correlates with disease activity of chronic hepatitis<br />

B (CHB), and can be detected in a considerable proportion<br />

of CHB patients after hepatitis B surface antigen (HBsAg)<br />

seroclearance. Its role in predicting HBV reactivation among<br />

HBsAg-negative, antibody to hepatitis B core antigen (anti-HBc)<br />

positive has not been explored. Methods: We retrieved stored<br />

plasma samples of HBsAg-negative, anti-HBc positive, patients<br />

with undetectable HBV DNA at baseline who were enrolled into<br />

two prospective <strong>studies</strong>, which aimed at investigating the rate<br />

of HBV reactivation during rituximab-containing chemotherapy<br />

and hematopoietic stem-cell transplantation (HSCT) (ClinicalTrials.gov<br />

identifier NCT01502397 and NCT01481647<br />

respectively). HBV reactivation was defined as detectable HBV<br />

DNA (≥10 IU/mL). Serum HBcrAg (lower limit of detection<br />

100 U/mL) was measured using a chemiluminscent enzyme<br />

immunoassay (Fujirebio Inc, Tokyo, Japan) at baseline and<br />

at every 3 months up to the date of last follow-up. Results:<br />

We included 131 HBsAg-negative, anti-HBc positive patients<br />

(rituximab/HSCT 47.3%/52.7%), with a mean age of 57.4<br />

(±15.1) years and a mean follow-up duration of 53.3 (±35.0)<br />

weeks. 32 patients developed HBV reactivation (rituximab/<br />

HSCT 56.3%/43.7%); the 2-year cumulative rate of HBV<br />

reactivation was 39.2%. 21 patients (16.0%) and 25 patients<br />

(19.1%) had detectable serum HBcrAg at either baseline or<br />

any time point before reactivation respectively. The median<br />

detectable baseline HBcrAg level was 400 (range 200-5,300)<br />

U/mL. Baseline serum HBcrAg positivity, compared to baseline<br />

HBcrAg-negative patients, had a significantly higher 2-year<br />

cumulative rate of HBV reactivation (75.8% versus 29.9%<br />

respectively, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 967A<br />

come of HBsAg loss/seroconversion, with a PPV of 83%. NCP<br />

on treatment was associated with a NPV of 69%. Discussion<br />

The measurement of the loss or gain of epitope recognition on<br />

TDF therapy revealed selection pressures directed against the<br />

HBsAg. Detection of the HBsAg CP by epitope mapping may<br />

be a potential viral biomarker, possibly reflecting emerging<br />

anti-HBs selection pressure. These findings may help individualising<br />

patient therapy, though further validation in expanded<br />

datasets and with other HBV genotypes is needed.<br />

Disclosures:<br />

Julianne Bayliss - Grant/Research Support: Gilead Inc.<br />

Peter A. Revill - Grant/Research Support: Gilead Sciences<br />

Thomas Leary - Employment: Abbott Laboratories<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

Kathryn M. Kitrinos - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

Stephen Locarnini - Consulting: Gilead, Arrowhead; Employment: Melbourne<br />

Health<br />

The following authors have nothing to disclose: Renae Walsh, Rachel Hammond,<br />

Lilly Yuen, Howard Thomas<br />

1552<br />

Comparison between spontaneous and post-nucleos(t)<br />

ide analogue HBsAg seroclearance in chronic HBV<br />

infection<br />

Yi-Cheng Chen, Rachel Wen-Juei Jeng, Rong-Nan Chien, Chia-<br />

Ming Chu, Yun -Fan Liaw; Liver Research Unit, Chang Gung<br />

Memorial Hospital, Taoyuan, Taiwan<br />

Background and Aims The prognosis following spontaneous<br />

HBsAg seroclearance in patients with chronic hepatitis B is<br />

excellent, except in those with cirrhosis or concurrent hepatitis<br />

C or D virus (HCV or HDV) infection. HBsAg seroclearance<br />

after nucleos(t)ide analogue (NUC) treatment is associated<br />

with favorable clinical outcomes. Whether there is difference<br />

between spontaneous and post-NUC HBsAg seroclearance is<br />

unknown. Methods Eight-eight patients with HBsAg seroclearance<br />

after NUC treatment (group A) and 88 age (±1 year)<br />

and gender-matched patients with spontaneous HBsAg seroclearance<br />

(group B) were included for analysis. Patients with<br />

coinfection of HCV or HDV, exposure to interferon and organ<br />

transplantation have been excluded. HBsAg seroclearance<br />

was defined as persistent absence of HBsAg for at least 12<br />

months and until last visit. Results The median age at HBsAg<br />

seroclearance was 49 years and 78% were males. The proportion<br />

of parenchymal fatty change at HBsAg seroclearance was<br />

comparable in both groups (p=1.000). The rate of cirrhosis<br />

when HBsAg seroclearance occurred was significantly higher<br />

in group A (21.6% vs. 4.5%, p=0.002). During a median<br />

follow-up period of 67.8 months in group A and 82.2 months<br />

in group B, one hepatocellular carcinoma (HCC) developed in<br />

each group with an estimated annual incidence of 0.2% and<br />

0.16%, respectively (p=1.000). The cumulative incidence of<br />

HBsAg seroconversion at 1 year, 3 years and 5 years after<br />

HBsAg seroclearance was 37.6% and 22%, 60.2% and<br />

39.9%, 67.1% and 53.5% in group A and group B, respectively<br />

(p=0.003 at 5 years). Conclusions There was significant<br />

higher rate of cirrhosis in patients with NUC treatment when<br />

HBsAg seroclearance occurred. Even in low incidence, HCC<br />

could develop in patients after HBsAg seroclearance. Long-term<br />

follow-up is needed to elucidate this untoward event.<br />

Disclosures:<br />

Yun -Fan Liaw - Advisory Committees or Review Panels: Roche; Grant/Research<br />

Support: Roche<br />

The following authors have nothing to disclose: Yi-Cheng Chen, Rachel Wen-Juei<br />

Jeng, Rong-Nan Chien, Chia-Ming Chu<br />

1553<br />

Different DNA methylation patterns by viral status in<br />

liver cancer<br />

Min-Ae Song 2,4 , Sandi Kwee 3 , Maarit Tiirikainen 2 , Gordon S.<br />

Okimoto 2 , Brenda Y. Hernandez 2 , Linda L. Wong 2,1 , Naoky CS<br />

C. Tsai 2 , Herbert Yu 2 ; 1 Surgery, Univ of Hawaii School of Medicine,<br />

Honolulu, HI; 2 Cancer Center, University of Hawaii, Honolulu,<br />

HI; 3 PET research, Queens Medical Center, Honolulu, HI;<br />

4 College of Public Health, Ohio State University, Columbus, OH<br />

Hepatocellular carcinoma (HCC) is one of the most common<br />

human cancers with high fatality. There is limited understanding<br />

of genome-wide differences in DNA methylation by viral<br />

status in HCC. Methods:We performed genome-wide methylation<br />

profiling in a total of 33 HCC tumor samples, including<br />

10 HBV-HCC, 13 HCV-HCC, and 10 non-viral HCC, using the<br />

Illumina HumanMethylation450 BeadChip. Results:Thirty-nine<br />

loci were found significantly different in methylation among<br />

the three groups (p


968A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Sandi Kwee - Grant/Research Support: Bayer Healthcare<br />

Linda L. Wong - Speaking and Teaching: Bayer, Bayer<br />

Naoky CS C. Tsai - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, AbbVie; Consulting: BMS, Gilead; Grant/Research Support: BMS,<br />

Gilead, AbbVie; Speaking and Teaching: Bristol-Myers Squibb, Gilead, Salix,<br />

Bayer, AbbVie<br />

The following authors have nothing to disclose: Min-Ae Song, Maarit Tiirikainen,<br />

Gordon S. Okimoto, Brenda Y. Hernandez, Herbert Yu<br />

ng/mL) in this study population, suggesting an opportunity to<br />

improve the current HCC screening for AFP-negative HCC. To<br />

validate the urine DNA test for the early detection of primary<br />

and recurrent HCC, an on-going blinded study was conducted<br />

from a population of HBV-infected cirrhosis patients, with or<br />

without previous HCC. All patients were monitored for severe<br />

liver diseases including HCC. Urine samples were collected<br />

at visits with a hepatologist and were barcoded and tested<br />

for three DNA markers. After a course of 6 to 18 months of<br />

follow-ups, four HCC cases, including three recurrent cases,<br />

were detected by MRI. More significantly, urine DNA markers<br />

detected three of four HCC cases at 6 and 9 months prior to<br />

the time of MRI diagnosis. In conclusion, HCC DNA markers<br />

can be detected in urine of patients with HCC by short-amplicon,<br />

PCR-based assays. Furthermore, we have demonstrated<br />

that in combination with serum AFP, a urine DNA test detected<br />

at least 30% more HCC in an open-labeled study as compared<br />

to serum AFP alone, and therefore, this test has the potential<br />

to detect HCC prior to MRI imaging in a blinded pilot study.<br />

Su, et. al., J Mol Diag, 6, 101-107. (2004) Lin, et al., J Mol<br />

Diag 13: 474-484 (2011) Jain, et al., Hepatology Research,<br />

(2014), doi: 10.1111/hepr.12449<br />

Disclosures:<br />

Hie-Won Hann - Advisory Committees or Review Panels: Gilead; Grant/Research<br />

Support: Gilead, Bristol-Myers Squibb<br />

Surbhi Jain - Employment: JBS Science, Inc<br />

Wei Song - Stock Shareholder: JBS Science Inc<br />

Ying-Hsiu Su - Advisory Committees or Review Panels: JBS Science Inc<br />

The following authors have nothing to disclose: Selena Lin, Sooji Yi, Ting-Tsung<br />

Chang, Chi-Tan Hu<br />

1554<br />

Detection of genetic and epigenetic DNA markers in<br />

urine for the early detection of primary and recurrent<br />

hepatocellular carcinoma<br />

Hie-Won Hann 1 , Surbhi Jain 2 , Selena Lin 6 , Sooji Yi 1 , Ting-Tsung<br />

Chang 3 , Chi-Tan Hu 4 , Wei Song 2 , Ying-Hsiu Su 5 ; 1 Liver Disease<br />

Prevention Center, Division of Gastroenterology and Hepatology,<br />

Thomas Jefferson University Hospital, Philadelphia, PA; 2 JBS Science,<br />

Inc, Doylestown, PA; 3 National Cheng Kung Universtiy Medical<br />

College, Tainan, Taiwan; 4 Buddhist Tzu Chi General Hospital<br />

and Tzu Chi University, Hualien, Taiwan; 5 The Baruch S. Blumberg<br />

Institute, Doylestown, PA; 6 Drexel University College of Medicine,<br />

Philadelphia, PA<br />

The purpose of this study is to explore the potential of urine<br />

DNA biomarkers for the early detection of primary and recurrent<br />

hepatocellular carcinoma (HCC). HCC is an aggressive<br />

disease with a 5-year survival rate of 26%, in early-stage cancers,<br />

and a mere 2% in later stages, with an approximate<br />

50% recurrence rate in the first 2 years of treatment. The most<br />

commonly used screening biomarker is serum alpha-fetoprotein<br />

(AFP), which detects only 40-60% of cases. We have<br />

previously shown that urine contains fragmented, circulation-derived,<br />

cell-free DNA that can be used for detection of<br />

cancer-related DNA markers, if the tumor is present (1). In<br />

order to detect circulation-derived, cell-free DNA markers in<br />

urine, we have developed short amplicon (~50 bp) PCR-based<br />

assays for mutations in TP53 (249T) (2) and for aberrant DNA<br />

methylation in GSTP1 (mGSTP1) and RASSF1A (mRASSF1A<br />

(3)). In addition, we have shown that the AUROC of three urine<br />

DNA markers, mGSTP1, mRASSF1A, and TP53 249T mutations,<br />

in combination with AFP, is 0.98 in detecting primary<br />

HCC (n=74) from non HCC [(cirrhosis (n=45) and hepatitis<br />

(n=42)] patients by a logistic regression-based combination<br />

algorithm. Furthermore, these 3 DNA markers detected 35 of<br />

the 46 (76%) HCC samples that were AFP negative (< 20<br />

1555<br />

Prevalence and serological features of chronic hepatitis<br />

B patients with concurrent HBsAg and anti-HBs in the<br />

Hepatitis B Research Network (HBRN).<br />

William M. Lee 1 , Yona K. Cloonan 2 , Kathleen B. Schwarz 3 , Doan<br />

Y. Dao 1 , Anna S. Lok 4 ; 1 Division of Digestive and Liver Diseases,<br />

University of Texas Southwestern Medical Center at Dallas, Dallas,<br />

TX; 2 Epidemiology, University of Pittsburgh, Pittsburgh, PA; 3 Pediatrics,<br />

Johns Hopkins University School of Medicine, Baltimore, MD;<br />

4 Gastroenterology, University of Michigan, Ann Arbor, MI<br />

Background: Some patients with chronic HBV infection have<br />

concurrent HBsAg and anti-HBs. We examined the HBRN<br />

cohort to determine the prevalence of anti-HBs in participants<br />

with chronic HBV infection, and compared demographic and<br />

clinical factors between participants with and without anti-HBs.<br />

Methods: Data from 1402 (1115 adults and 287 pediatric)<br />

participants with chronic HBV infection enrolled in the HBRN,<br />

who had been tested for both HBsAg and anti-HBs within 1<br />

year of enrollment were included. 139 participants belonging<br />

to specifically targeted groups (e.g., pregnant women) were<br />

excluded from prevalence estimates. Characteristics of anti-HBs<br />

positive and anti-HBs negative participants were compared.<br />

Results: 93 participants were anti-HBs positive, overall prevalence<br />

of concurrent HBsAg and anti-HBs was 6.6% (7.7%<br />

pediatric and 6.3% adult). Age, sex, race and birthplace did<br />

not differ significantly between anti-HBs positive and anti-HBs<br />

negative participants. Similarly, there were no significant differences<br />

in mode of transmission, estimated years of infection,<br />

past HBV treatment, prevalence of HBeAg, HBV genotypes,<br />

AST, ALT or other biochemical markers of liver injury[SB1] in<br />

the 2 groups. HBsAg levels were significantly lower in anti-HBs<br />

positive participants: median 3.0 vs. 3.6 log 10<br />

IU/mL (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 969A<br />

noted in pediatric and adult groups. Conclusion: Concurrent<br />

HBsAg and anti-HBs is not uncommon in persons with chronic<br />

HBV infection. The prevalence of anti-HBs is similar for adults<br />

and children, and is not associated with markers of disease<br />

activity or a specific clinical phenotype. Anti-HBs positive participants<br />

had lower HBsAg titers but similar HBV DNA levels<br />

suggesting a role for anti-HBs in decreasing sub-virion particles<br />

or in immune control of HBV infection. Further <strong>studies</strong> are<br />

planned to examine surface gene variants and anti-HBs titers in<br />

this [SB2] population.<br />

Comparison Data between anti-HBs pos and neg HBsAg positive<br />

individuals<br />

Disclosures:<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

Kathleen B. Schwarz - Consulting: Novartis; Grant/Research Support: Bristol-Myers<br />

Squibb, Gilead, Roche/Genentech, Synageva, Vertex, Roche<br />

Anna S. Lok - Advisory Committees or Review Panels: Gilead, MYR, Tekmira;<br />

Consulting: GSK, Merck; Grant/Research Support: AbbVie, BMS, Gilead, Idenix<br />

The following authors have nothing to disclose: Yona K. Cloonan, Doan Y. Dao<br />

1556<br />

Clinical Outcomes of Patients with Private vs. Public<br />

Drug Coverage for Treatment of Chronic Hepatitis B<br />

Aman V. Arya, Mayur Brahmania, Matthew Kowgier, Hemant<br />

Shah, Orlando Cerocchi, David K. Wong, Jordan J. Feld, Harry<br />

L. Janssen; Gastroenterology, University of Toronto, Toronto, ON,<br />

Canada<br />

Background Canada has universal health coverage for basic<br />

medical care; however, prescription drugs are covered by a<br />

private insurance (if eligible) or through an exceptional access<br />

program (EAP) for those without insurance. Coverage for<br />

chronic hepatitis B treatment (CHB) is restricted by the EAP.<br />

Aim To compare clinical outcomes between individuals covered<br />

by EAP and those covered by private insurance for the<br />

management of CHB. Methods A retrospective chart audit was<br />

conducted of adult patients with CHB who were on treatment<br />

with a nucleos(t)ide analogue at a large tertiary care liver clinic<br />

from April 2010 to March 2015. Rates of HBeAg seroconversion,<br />

HBsAg loss, time to undetectable HBV DNA and ALT normalization,<br />

HBV DNA breakthrough, and ALT flares (defined as<br />

5xULN) on therapy were compared. Results 588 patients (285<br />

EAP, 303 non-EAP) met study inclusion criteria. There were 409<br />

males (70%) and the mean age was 52 (range = 19-89). 78%<br />

were Asian, 15% Caucasian and 4% African. 144 patients<br />

had cirrhosis (25%), The mean age of EAP patients receiving<br />

treatment was 56 years compared with 48 for patients with<br />

private insurance (p < 0.001). EAP patients were less likely<br />

to be born in Canada (2% vs. 10%, p < 0.001), more likely<br />

to have cirrhosis (29% vs 20%, p=0.02), and were less likely<br />

to be employed (49% vs. 77%, p < 0.001). LAM, TDF and<br />

ETV were used in 64%, 22%, 3% of EAP patients respectively,<br />

compared to 29%, 39%, 21% in patients with private insurance<br />

(p < 0.001). HBeAg seroconversion occurred in 42% of<br />

EAP patients compared to 46% of private insurance patients<br />

(p=0.89). HBsAg loss occurred in 5% versus 7% (p=0.39) and<br />

the mean time to HBV DNA negativity was 5 months (range:<br />

0-46) versus 5 months (range: 0-54) (p=0.17). Time to ALT<br />

normalization was 3 months and not different between the two<br />

groups (p=0.55). HBV DNA breakthrough leading to a change<br />

in therapy was more common among EAP patients (38% vs.<br />

22%, p


970A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

onymous (NS), genotype specific viral substitutions from wild<br />

type (WT) and entropy scores, were calculated across the<br />

whole genome (WG) and within specific structural (Envelope;<br />

Polymerase; precore-core, and; HBx) and regulatory regions<br />

(negative regulatory element- basal core promoter NRE-BCP)<br />

and correlated with baseline clinical and virological data and<br />

serological response at W192. Results 94 patients were successfully<br />

analysed (median age 32yrs; 75% male; genotype<br />

A/B/C/D 26%/23%/23%/28%; ALT 139 IU/mL; 24% cirrhosis;<br />

HBV DNA 8.43 log 10<br />

IU/mL; HBsAg 50,793 IU/mL;<br />

HBeAg 1,990 PEIU/mL; 77% TDF therapy W0-W48, 23%<br />

ADV therapy W0-W48). Median WG coverage, substitutions<br />

from WT and entropy scores are listed in Table 1. Entropy<br />

scores were similar across the genome, but three-fold greater<br />

in genotype B/C infection as compared to A/D (p< 0.0001).<br />

An increase in WG viral diversity was associated with lower<br />

baseline HBV DNA, HBeAg and HBsAg titres, independent<br />

of genotype (p< 0.0001 for all). Higher baseline viral diversity,<br />

specifically in BCP-NRE, was also associated with a lower<br />

likelihood of achieving HBeAg SC (p=0.04) and HBsAg loss<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 971A<br />

1559<br />

Genotype-Specific Prevalence and Importance of Naturally<br />

Occurring Precore-, Basal Core Promoter- and<br />

preS-Mutations in a Large European Study Cohort of<br />

Patients Chronically Infected with Hepatitis B Virus<br />

Lisa Sommer 1 , Kai-Henrik Peiffer 1 , Julia Dietz 1 , Simone Susser 1 ,<br />

Joerg Petersen 2 , Peter Buggisch 2 , Markus Cornberg 3 , Stefan<br />

Mauss 4 , Hartwig Klinker 5 , Martin F. Sprinzl 6 , Florian van Bömmel<br />

7 , Eberhard Hildt 8 , Caterina Berkowski 1 , Dany Perner 1 , Sandra<br />

Passmann 1 , Stefan Zeuzem 1 , Christoph Sarrazin 1 ; 1 Medical<br />

Clinic 1, J. W. Goethe University Hospital, Frankfurt, Germany;<br />

2 IFI-Institut an der Asklepiosklinik St. Georg, Hamburg, Germany;<br />

3 Gastroenterology, Hepatology and Endocrinology, Hannover<br />

Medical School, Hannover, Germany; 4 Center for HIV and Hepatogastroenterology,<br />

Düsseldorf, Germany; 5 Internal Medicine II,<br />

University of Wuerzburg Medical Center, Würzburg, Germany;<br />

6 I. Medizinische Klinik und Poliklinik, Universitätsmedizin der<br />

Johannes Gutenberg-Universität Mainz, Mainz, Germany; 7 University<br />

Hospital Leipzig, Leipzig, Germany; 8 Paul-Ehrlich-Institut,<br />

Langen, Germany<br />

Background/Aims In several <strong>studies</strong> from Asia mutations in<br />

HBV basal core promoter (BCP) and in the preS region are<br />

associated with the development of liver fibrosis, cirrhosis and/<br />

or hepatocellular carcinoma. For precore (PC) mutations data<br />

are controversially discussed. In EU/US only limited data about<br />

the significance of these mutations in chronically infected HBV<br />

patients (pts.) is available. To establish prognostic markers<br />

the genotype-specific prevalence of these mutations and their<br />

importance for progression of disease were examined in a<br />

large European study cohort of pts. infected with HBV genotypes<br />

(gt) A to E. Methods In the prospective, longitudinal<br />

ALBATROS study HBsAg carriers without antiviral therapy will<br />

be followed for more than 10 years. In the present interim<br />

analysis baseline (BL) sera of 340 pts. (gtA: 81, gtB: 28, gtC:<br />

16, gtD: 192, gtE: 23) were examined. Progression of infection<br />

was controlled up to 5 years follow-up (FU). BCP/PC and<br />

preS regions were amplified via nested PCR and analyzed<br />

by population-based sequencing. Results BCP double mutation<br />

A1762T/G1764A was found frequently in our population<br />

(203/340; 60%) with the highest prevalence in gtE (91%),<br />

with moderate frequency (56-69%) in gtA, gtC and gtD and<br />

with the lowest prevalence in gtB (29%). The PC G1896A<br />

mutation was detected in 42% (142/340) with the highest<br />

prevalence in gtB, gtC and gtD (71, 56 and 54%) and the<br />

lowest prevalence in gtA and gtE (10 and 9%). In addition, the<br />

PCG1896A/G1899A double mutation was found with moderate<br />

frequency in gtD, gtB and gtC (27, 18 and 13%) but only<br />

infrequently in gtA (3%) and was absent in gtE (0%). Deletions<br />

in the preS region varying in length from 1aa to 41aa occurred<br />

in 8% of investigated patients with the highest prevalence in<br />

gtE (40%) and gtC (29%). Seven patients (gtA: 2; gtD: 4; gtE:<br />

1) started antiviral therapy after 1-4 years of FU. BCP double<br />

mutation was observed in 6/7 and PC double mutation in<br />

1/7 patients at BL. Conclusion BCP and PC mutations were<br />

found frequently in our study cohort in a HBV genotype-specific<br />

pattern. Deletions in the preS region were found with a lower<br />

prevalence but were observed predominantly in HBV gtE and<br />

gtC. Interestingly, 6/7 HBsAg carriers who had to start treatment<br />

during FU were positive for the BCP double mutation at<br />

BL, whereas PC mutations were found only infrequently in these<br />

patients. Therefore, BCP and PC mutations might be of clinical<br />

interest as potential prognostic markers in HBsAg carriers.<br />

Disclosures:<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

Peter Buggisch - Advisory Committees or Review Panels: Janssen, AbbVie, BMS;<br />

Speaking and Teaching: Roche, MSD, Gilead, Merz Pharma<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

Stefan Mauss - Advisory Committees or Review Panels: BMS, AbbVie, Janssen,<br />

ViiV, Gilead; Speaking and Teaching: BMS, AbbVie, Janssen, Gilead, MSD<br />

Hartwig Klinker - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Janssen, Hexal; Grant/Research Support: Gilead, Abbott, AbbVie, Janssen,<br />

Novartis, Bristol-Myers Squibb; Speaking and Teaching: Gilead, Bristol-Myers<br />

Squibb, Merck Sharp & Dohme<br />

Martin F. Sprinzl - Grant/Research Support: GILEAD<br />

Florian van Bömmel - Advisory Committees or Review Panels: Gilead Sciences<br />

Inc., Roche Pharmaceutics; Grant/Research Support: Biopredictive; Speaking<br />

and Teaching: Bristol-Myers Squibb, Janssen, Roche Pharmaceutics, Gilead Sciences<br />

Inc., Abbvie<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

The following authors have nothing to disclose: Lisa Sommer, Kai-Henrik Peiffer,<br />

Julia Dietz, Simone Susser, Eberhard Hildt, Caterina Berkowski, Dany Perner,<br />

Sandra Passmann<br />

1560<br />

Individualizing Hepatitis B Prophylaxis in Liver Transplant<br />

(LT) Recipients: Diminished Need for Hepatitis B<br />

Immunoglobulin (HBIG)<br />

Kavita Radhakrishnan 1 , Aileen Chi 2 , David J. Quan 2 , John P.<br />

Roberts 3 , Norah Terrault 2 ; 1 Medicine, UCSF, San Francisco, CA;<br />

2 Transplant Hepatology, UCSF, San Francisco, CA; 3 Transplant<br />

Surgery, UCSF, San Francisco, CA<br />

Background & Aims: HBIG has been central to the prevention<br />

of hepatitis B virus (HBV) recurrence post-LT for two decades.<br />

However, its high cost and need for IV or IM administration<br />

have prompted HBIG minimization strategies. Elimination of<br />

HBIG and use of antiviral therapy (AVT) alone is associated<br />

with HBsAg recurrence in ~20% at 2 yrs (Fung, Gastroenterology.<br />

2011). Achieving a higher rate of HBsAg negativity is<br />

highly desirable. We report our experience with an individualized<br />

protocol for HBIG prophylaxis. Methods: All HBsAg-positive<br />

patients undergoing LT from 2009-2014 were included.<br />

Patients at high risk for recurrence [defined as HBV DNA>100<br />

IU/mL at time of LT (33%, 2 acute), prior AVT resistance (33%),<br />

HDV+ (22%), or HIV+ (12%)] received HBIG 5K-10K U in anhepatic<br />

phase & daily X 5 days, and every 4-12 wks to maintain<br />

trough anti-HBs >100 U/L. In contrast, low risk patients [all<br />

others] received HBIG 5K U in anhepatic phase & daily x 5<br />

days only. Both groups received ATV therapy indefinitely. HBV<br />

recurrence was defined as HBsAg positivity post-LT. Results: Of<br />

79 LT recipients, 63% and 37% met criteria for high and low<br />

risk protocols. Pre-LT characteristics and post-LT prophylaxis<br />

are shown in Table. Median (IQR) follow-up was 2.9 (1.3-4.4)<br />

years. The 1 and 3-yr cumulative incidences of HBsAg positivity<br />

post-LT were 2.0% (CI 95<br />

: 0.3 -13.4) and 5.4% (CI 95<br />

: 1.3-<br />

20.7) in the high risk vs. 0% (upper CI 95<br />

: 11.9) in the low risk<br />

groups (p=0.28). Only those who developed metastatic HCC<br />

became HBsAg-positive (n=2). All patients were HBV DNA<br />

negative post-LT, except one with recurrent HCC. Post-LT survival<br />

was 91% and 84% at 3 and 5 years respectively, with no<br />

difference between prophylaxis groups (p=0.55). Conclusion:<br />

In patients without HIV and HDV who have HBV DNA levels<br />


972A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

is highly effective in preventing HBV recurrence. With current<br />

preferred antivirals (high barrier to resistance), more patients<br />

are likely to have such a profile and utilizing an individualized<br />

approach to HBIG prophylaxis offers a better balance between<br />

prophylaxis costs and prevention of HBV recurrence.<br />

ULN: 575.0 vs 127.9; ALT


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 973A<br />

ily was detected by Western blot, and the DNA binding level of<br />

hepatocyte nuclear factor 4α(HNF4α) was detected by EMSA.<br />

Summary: Luteolin decreased the levels of HBsAg and HBeAg<br />

and the HBV DNA load both in the cultured medium and in<br />

serum, and it also suppressed the HBV DNA RI and the expression<br />

of HBsAg and HBcAg, which demonstrated that luteolin<br />

can effectively inhibit HBV replication both in vitro and in vivo.<br />

In addition, luteolin effectively downregulated the expression of<br />

HNF4α and decreased its binding level with HBV preC/C promoter<br />

in HepG2.2.15 cells. The extracellular signal-regulated<br />

kinase(ERK) and c-Jun N-terminal kinase(JNK) signal pathways,<br />

but not the p38 signal pathway in the MAPK family, were<br />

activated by luteolin, which can be blocked by their inhibitors,<br />

respectively. However, it was the ERK inhibitor(U0126) rather<br />

than the JNK inhibitor(SP600125) can attenuate the down-regulation<br />

of HNF4α by luteolin and restore the protein level of<br />

HNF4α. In addition, blocking the ERK pathway with inhibitor<br />

can cripple the inhibition of HBV replication by luteolin, while<br />

blocking the JNK pathway do not affect the inhibitory effect<br />

of luteolin on HBV replication. Conclusions: Luteolin inhibits<br />

HBV replication through down-regulating HNF4α expressin via<br />

activating ERK signal pathway. Keywords: luteolin; hepatitis B<br />

virus; replication; hepatocyte nuclear factor 4α; extracellular<br />

signal-regulated kinase<br />

Disclosures:<br />

The following authors have nothing to disclose: Yunhong Nong, Ying Shi, Miao<br />

Liu, Libo Yan, Yong Lin, Lang Bai, Hong Tang<br />

1563<br />

Redefining the Natural History of Chronic Hepatitis B:<br />

Using Viral Diversity to Classify Disease and Predict<br />

Serological Response to Nucleos(t)ide Analogue (NA)<br />

Therapy<br />

Julianne Bayliss 1 , Gillian Rosenberg 1 , Lilly Yuen 1 , Anuj Gaggar 2 ,<br />

Kathryn M. Kitrinos 2 , Mani Subramanian 2 , Edward J. Gane 3 ,<br />

Henry Lik-Yuen Chan 4 , Rachel Hammond 1 , Scott Bowden 1 , Peter<br />

A. Revill 1 , Stephen Locarnini 1 , Alex J. Thompson 5 ; 1 Molecular<br />

R&D, VIDRL, Melbourne, VIC, Australia; 2 Gilead Sciences, Foster<br />

City, CA; 3 New Zealand Liver Transplant Unit, Auckland City<br />

Hospital, Auckland, New Zealand; 4 Medicine and Therapeutics,<br />

Chinese University of Hong Kong, Hong Kong, Hong Kong; 5 Gastroenterology,<br />

St Vincent’s Hospital, Melbourne, VIC, Australia<br />

Background The presence of basal core promoter (BCP) and<br />

precore (PC) mutations in immune tolerant (IT) patients with<br />

chronic hepatitis B (CHB) suggests a transition between IT<br />

and immune clearance (IC) disease. Here, next generation<br />

sequencing (NGS) was used to analyse the impact of baseline<br />

viral diversity on CHB natural history and its association<br />

with baseline serology and response to NA treatment. Methods<br />

GS-US-203-0101 and GS-US-174-0103 evaluated tenofovir<br />

DF (TDF) containing regimens in IT and IC patients, respectively.<br />

Only genotype B/C patients were included (B= 83;<br />

C= 102). Patients were classified as True IT (ALT 20-40-80IU/mL). Viral diversity, defined as non-synonymous<br />

(NS), genotype specific viral substitutions from wild<br />

type (WT) and entropy scores across the whole genome (WG),<br />

were calculated using Illumina MiSeq NGS and correlated<br />

with baseline clinical and virological data and W192 serological<br />

response. Results 185 patients were analysed (baseline:<br />

median age 32yrs, 55% male; median ALT 29 IU/mL, HBV<br />

DNA 8.38 log 10<br />

IU/mL, HBsAg 51,590 IU/mL, HBeAg 3,358<br />

PEIU/mL). Increased viral diversity was associated with lower<br />

HBV DNA (p= 0.001), HBeAg (p< 0.0001) and HBsAg (p<<br />

0.0001) titres and higher ALT levels (p< 0.0001). Although<br />

HBV DNA, HBeAg and HBsAg titres did not differ significantly<br />

between True IT and Late IT patients (Table 1), a significantly<br />

higher percentage of Late IT patients had BCP and PC mutations<br />

(p= 0.001 and p< 0.0001, respectively). Late IT patients<br />

also had more NS substitutions (p= 0.001), higher entropy<br />

scores (p= 0.002), and were more likely to achieve HBeAg<br />

seroconversion by W192 compared to True IT patients (True<br />

IT= 3%, Late IT= 14%, p< 0.0001). Discussion There is a subset<br />

of IT CHB patients with a virological profile similar to IC<br />

patients. This subset of patients may be transitioning between<br />

IT and IC disease and are more likely to respond favourably<br />

to TDF therapy.<br />

Disclosures:<br />

Julianne Bayliss - Grant/Research Support: Gilead Inc.<br />

Gillian Rosenberg - Grant/Research Support: Gilead Sciences<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Kathryn M. Kitrinos - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Mani Subramanian - Employment: Human Genome Sciences, Human Genome<br />

Sciences, Human Genome Sciences, Human Genome Sciences; Stock Shareholder:<br />

Human Genome Sciences, Human Genome Sciences, Human Genome<br />

Sciences, Human Genome Sciences<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

Peter A. Revill - Grant/Research Support: Gilead Sciences<br />

Stephen Locarnini - Consulting: Gilead, Arrowhead; Employment: Melbourne<br />

Health<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

The following authors have nothing to disclose: Lilly Yuen, Rachel Hammond,<br />

Scott Bowden<br />

1564<br />

Angiopoetin-like protein 2 independently associated<br />

with liver inflammation and fibrosis in patients chronically<br />

infected with hepatitis B virus<br />

Hong Zhao, Yong-Qiong Deng, Gui-Qiang Wang; Department of<br />

infectious disease, Peking University First Hospital, Beijing, China<br />

Background: Guidelines recommend antiviral therapy for<br />

HBeAg(+)patients with HBV DNA >=20000IU/ml or HBeAg(-)<br />

patients with HBVDNA >=2000IU/ml, and alanine aminotransferase<br />

(ALT) levels more than 2 times of upper limit of normal<br />

(ULN). For patients who don’t meet the criteria above, therapy<br />

decision depends on hepatic histology, an invasive process.<br />

The available non-invasive assessments could only predict the<br />

stages of liver fibrosis. Therefore, a non-invasive method or<br />

biomarker predicting moderate and more inflammation or significant<br />

fibrosis is urgently needed. Angiopoetin-like protein 2<br />

(Angptl2) is a mediator of chronic inflammation contributing to<br />

extracellular matrix remodeling. Aim: The aim of the study was<br />

to explore the predicting value of serum angptl2 for moderate


974A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and more inflammation or significant fibrosis. Methods: One<br />

hundred and seventy-three patients with chronic hepatitis B<br />

virus (HBV) infection, whose treatment decisions depended on<br />

liver biopsies, were enrolled in the study (NCT01962155 and<br />

ChiCTR-DDT-13003724). The levels of serum angptl2 were<br />

detected by Human ANGPTL2 Assay kit (Immuno-Biological<br />

Laboratories Co., Japan). Liver biopsies were performed in<br />

all patients. Histological Activity Index (HAI) and Liver fibrosis<br />

stage were assessed according to Ishak criteria independently<br />

by 2 pathologists. Results: 1) Serum Angptl2 levels were significantly<br />

associated with not only HAI scores (p=0.000, spearman<br />

rho=0.327) but also the liver fibrosis stages (P=0.000,<br />

spearman rho=0.318). 2) Patients with moderate and more<br />

inflammation or significant fibrosis (HAI>=5 or F>=3)who<br />

should be treated immediatelyhad higher serum angptl2 (5.76<br />

± 2.62 vs 4.31±1.82 ug/ml), AST, GGT and lower platelet<br />

counts, HBsAg. Multivariate analysis confirmed that only<br />

serum angptl2 level could independently predict urgent antiviral<br />

therapy. 3) Angptl2 showed areas under the receiver<br />

operating characteristics curve of 0.69 (95% CI: 0.60, 0.76)<br />

for indicating moderate and more inflammation or significant<br />

fibrosis. Leave-one-out cross-validation showed 64% cross-validation<br />

grouped cases correctly classified. Using a cutoff value<br />

of ≥7.36ug/ml, moderate and more inflammation or significant<br />

fibrosis could be correctly identified with 91.76% specificity<br />

and 72.0% PPV. Similarly, a lower angptl2 level (3.95 ug/<br />

ml) could excluded urgent therapy with 78.41% specificity<br />

and 70.8% PPV. Conclusion: Serum angptl2, a novel single<br />

biomarker, independently predict moderate and more inflammation<br />

or significant fibrosis, saving half of patients from liver<br />

biopsies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hong Zhao, Yong-Qiong Deng,<br />

Gui-Qiang Wang<br />

1565<br />

PreS mutations analyzed by deep sequencing are<br />

correlated with the progression of liver diseases and<br />

HBsAg production.<br />

Yuichiro Suzuki, Shinya Maekawa, Nobutoshi Komatsu, Mitsuaki<br />

Sato, Masaru Muraoka, Shuya Matsuda, Mika Miura, Yasuhiro<br />

Nakayama, Taisuke Inoue, Minoru Sakamoto, Nobuyuki Enomoto;<br />

University of Yamanashi, Chuo, Japan<br />

Background and Aim: Recently, it is considered that achieving<br />

HBsAg seroclearance is the ultimate therapeutic goal in<br />

the treatment of chronic hepatitis B since serum HBsAg titer<br />

is suspected to reflect HBV-cccDNA titer in the liver. On the<br />

other hand, preS region of HBV genome responsible for HBsAg<br />

production was reported to be often mutated in advanced liver<br />

disease by direct sequencing. However, the interrelationship<br />

among HBsAg, preS mutation and liver disease progression is<br />

complicated and remains unclear. In the present study, to clarify<br />

the interrelationship among HBsAg titer, preS mutation and<br />

disease progression, we performed deep sequencing study<br />

for preS mutations. Methods: A total of 96 patients including<br />

32 inactive carriers (IC), 28 with chronic hepatitis (CH) and<br />

36 with cirrhosis or hepatocellular carcinoma (LC/HCC) were<br />

analyzed. All of inactive carriers were HBeAg negative, and<br />

22 patients were with HCC. They were all nucleotide analogue<br />

naïve. Deep sequencing was performed targeting 522nt in the<br />

preS region from patient serum, and proportion of preS1 and<br />

preS2 start codon mutations, preS1 and preS2 deletions was<br />

determined in each patient. Results: In IC group, CH group,<br />

and LC/HCC group, median HBsAg was respectively 688 IU/<br />

ml, 2075 IU/ml, and 1065IU/ml. HBsAg level


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 975A<br />

ent association of the genotype 5 isolates with region/country<br />

of birth. Conclusion: This study demonstrates the increasing<br />

proportion of HDV infections which were acquired overseas<br />

prior to migration and compares phylogenetic analysis with<br />

country of birth data from surveillance notifications. Detection<br />

of genotype 5 strains (predominantly found in African patients)<br />

are increasing in Australia, reflecting changing migration patterns.<br />

Understanding the genotypes and country of origin of<br />

HDV may play a role in the diagnosis and management of an<br />

infection that is poorly recognized.<br />

Disclosures:<br />

Stephen Locarnini - Consulting: Gilead, Arrowhead; Employment: Melbourne<br />

Health<br />

The following authors have nothing to disclose: Kathy Jackson, Margaret Littlejohn,<br />

Lilly Yuen, Benjamin C. Cowie, Jennifer H. MacLachlan, Scott Bowden<br />

1567<br />

Kinetics of Serum Hepatitis B Surface Antigen During<br />

the First Year of Tenofovir Treatment in Patients With<br />

Chronic Hepatitis B<br />

Choong Kyun Noh, Soon Sun Kim, Sun Young Park, Joohan Park,<br />

Hyo Jung Cho, Jae Youn Cheong, Sung Won Cho; Ajou University<br />

School of Medicine, Suwon, Korea (the Republic of)<br />

Aims The predictive role of quantitative hepatitis B surface antigen<br />

(HBsAg) levels for HBsAg or hepatitis B e antigen (HBeAg)<br />

loss/seroconversion in patients receiving potent oral antiviral<br />

therapy is suggested. However, data on patients with genotype<br />

C hepatitis B virus is limited. We evaluated the HBsAg<br />

kinetics in chronic hepatitis B patients treated with tenofovir<br />

and correlation with treatment response. We also compared<br />

HBsAg kinetics with those of chronic hepatitis B patients<br />

treated with entecavir. Methods Serial HBsAg levels were<br />

determined in sera from 66 tenofovir (29 HBeAg-positive and<br />

37 HBeAg-negative) and 101 entecavir (59 HBeAg-positive<br />

and 42 HBeAg-negative)-treated chronic hepatitis B patients at<br />

baseline, 3, 6 and 12 months during treatment. Results During<br />

12 months of tenofovir treatment, the mean HBsAg levels did<br />

not significantly decrease in both HBeAg-positive and negative<br />

patients (3.40, 3.24, 3.26 and 3.30 log 10<br />

IU/ml and 3.31,<br />

3.39, 3.37 and 3.34 log 10<br />

IU/ml at baseline, 3, 6, 12 months<br />

of treatment, respectively). Only two HBeAg-positive patients<br />

exhibited a significant (≥ 1.0 log) decline in HBsAg levels<br />

until 12 months. However, decline of HBsAg levels between<br />

baseline and 3 months of treatment showed a tendency to be<br />

larger in HBeAg-positive patients than those in HBeAg-negative<br />

patients (0.09 log 10<br />

IU/ml decrement vs. 0.07 log 10<br />

IU/<br />

ml increment, P = 0.66). The on-treatment HBsAg levels did<br />

not correlate with HBeAg loss/seroconversion at 12 months in<br />

HBeAg-positive tenofovir-treated patients. Changes of HBsAg<br />

levels during tenofovir and entecavir were not significantly different<br />

in both HBeAg-positive and HBeAg-negative patients.<br />

Conclusions Decline of HBsAg levels during the 12 months<br />

of tenofovir treatment was insignificant in chronic hepatitis<br />

B patients with genotype C. Long-term oral antiviral therapy<br />

will be needed in majority of chronic hepatitis B patients even<br />

though antiviral therapies are very potent.<br />

Disclosures:<br />

The following authors have nothing to disclose: Choong Kyun Noh, Soon Sun<br />

Kim, Sun Young Park, Joohan Park, Hyo Jung Cho, Jae Youn Cheong, Sung<br />

Won Cho<br />

1568<br />

The PAGE-B Score Stratifies Chronic Hepatitis B Patients<br />

with Compensated Cirrhosis at High Risk of Hepatocellular<br />

Carcinoma Development with Good Accuracy<br />

Willem Pieter Brouwer 1 , Bettina E. Hansen 1 , Elena Raffetti 2 ,<br />

Francesco Donato 2 , Giovanna Fattovich 3 ; 1 Gastroenterology &<br />

Hepatology, Erasmus Medical Center Rotterdam, Rotterdam, Netherlands;<br />

2 Epidemiology and Public Health, Institute of Hygiene,<br />

Brescia, Italy; 3 Medicine, Azienda Ospedaliera Universitaria Integrata,<br />

Verona, Italy<br />

Background & aims. The PAGE-B score has been associated<br />

with HCC development in Caucasian patients treated with<br />

potent nucleos(t)ide analogues. We aimed to assess the performance<br />

of this marker in a cohort of untreated Western chronic<br />

hepatitis B (CHB) patients with Child Pugh A cirrhosis. Methods.<br />

In this retrospective cohort study conducted at 9 tertiary care<br />

centres in Europe, CHB patients with biopsy-proven, compensated<br />

cirrhosis and available laboratory measurements were<br />

included. The PAGE-B, CU-HCC, GAG-HCC and REACH-B<br />

scores were calculated and associated with the occurrence<br />

of liver failure, HCC development, transplant-free survival or<br />

any clinical event (combined endpoint including all the respective<br />

events). Results. Of 134 patients, mean age at inclusion<br />

was 47±13 years, 70% was HBeAg-negative and 87% male.<br />

In total, 16 developed an HCC, 21 died or underwent liver<br />

transplantation and 41 developed any clinical event during<br />

a median follow-up of 6.2 years (IQR 3.1-8.6). The 5-year<br />

(95%CI) HCC-free and transplant-free survival was 90% (84-<br />

95) and 87% (81-93), respectively. The PAGE-B score was<br />

significantly associated with HCC development (HR 1.34,<br />

95%CI:1.2-1.6, p


976A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1569<br />

Health care disparity in delivering optimal care to<br />

chronic hepatitis B pregnant mothers<br />

Rasham Mittal, Amandeep K. Sahota; Kaiser Permanente Los<br />

Angeles Medical Center, Los Angeles, CA<br />

Background Roughly 24,000 chronic hepatitis B (CHB) pregnant<br />

women give birth in United States every year. Since perinatal<br />

acquisition of CHB contributes significantly to prevalence<br />

of CHB, it becomes imperative to provide optimal antenatal<br />

care to these mothers. Study Aim: - Determine if health care<br />

disparities exist in antenatal care of CHB pregnant mothers at<br />

a large integrated health care system in United States. Methods:<br />

Retrospective study (January 2005- June 2010) at a large<br />

integrated health care system in Southern California. All CHB<br />

pregnant mothers identified using the regional perinatal database.<br />

Inclusion criteria: age > 18 years and positive serum<br />

HBsAg (hepatitis B surface antigen). They were excluded if<br />

tested negative for HBsAg; tested positive for preliminary<br />

enzyme-linked immunosorbent assay (ELISA) HBsAg but had<br />

non-reactive HBsAg neutralization confirmatory antibody assay<br />

(false positive); termination of pregnancy or lost to follow up.<br />

Only first live-birth pregnancy was considered for CHB mothers<br />

who delivered more than once during the study timeline. Chart<br />

review was performed to determine maternal antenatal hepatitis<br />

Be antigen (HBeAg) serology, antenatal hepatitis B virus (HBV)<br />

viral load and specialist consultation visits. Results: A total of<br />

462 CHB pregnant mothers were identified. The mean maternal<br />

age in years was 32.5 ± 4.8 (range 18-45, median 33).<br />

Highest prevalence of CHB was noted in Asian Pacific Islander<br />

mothers (78.8%), followed by non-Hispanic white (9.7%), Hispanic<br />

white (5.8%), non-Hispanic black (5.2%) and Hispanic<br />

black (0.4%). Antenatal hepatitis Be antigen (HBeAg) serology<br />

testing was performed in 274 (59.3%) and hepatitis B virus<br />

(HBV) viral load was tested in 156 (33.8%) mother only. Only<br />

93 (20.1%) mothers had a specialist consultation during their<br />

pregnancy. Among these 93 mothers, 90 consulted gastroenterologist<br />

and 3 hepatologist. Additionally 104 (26.2%) infants<br />

born to these CHB mothers had no HBsAg serological testing<br />

at 9-18 months follow up. Conclusion: - Significant health care<br />

disparities in standard care of CHB pregnant mothers identified<br />

at a large integrated health care system in United States.<br />

Disclosures:<br />

The following authors have nothing to disclose: Rasham Mittal, Amandeep K.<br />

Sahota<br />

1570<br />

Ultrasound predicts steatosis in patients with chronic<br />

hepatitis B<br />

Erin Kelly 2,1 , Vickie A. Feldstein 3 , Rebecca Hudock 1 , Marion G.<br />

Peters 1 ; 1 Hepatology, University of California San Francisco, San<br />

Francisco, CA; 2 Gastroenterology and Hepatology, University of<br />

Ottawa, Ottawa, ON, Canada; 3 Radiology, UCSF, San Francisco,<br />

CA<br />

Inflammation and fibrosis may impair the ability of ultrasound<br />

to identify steatosis in patients with chronic hepatitis B (CHB).<br />

We determined the accuracy of ultrasound (US) in grading<br />

steatosis in patients with CHB compared to liver biopsy, and<br />

examined clinical factors associated with steatosis. This was a<br />

single-center, retrospective study of all non-transplanted CHB<br />

patients undergoing US and same day liver biopsies from<br />

2004-2014. Steatosis was graded by US as none, mild, moderate<br />

or severe. Liver histology graded steatosis (0:


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 977A<br />

greatest risk. Here we investigate whether a qHBsAg threshold<br />

(


978A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1573<br />

Application of Highly Sensitive Chemiluminescent<br />

Enzyme Immunoassay for Hepatitis B Surface Antigen to<br />

Detect Occult HBV infection: an Appropriate Method as<br />

Pre-Transfusion Testing.<br />

Takako Inoue 1 , Noriyo Ochi 1 , Takaaki Goto 1 , Shintaro Ogawa 2 ,<br />

Noboru Shinkai 3 , Kumiko Ohne 1 , Yukio Wakimoto 1 , Yasuhito<br />

Tanaka 2,1 ; 1 Clinical Laboratory, Nagoya City University Hospital,<br />

Nagoya, Japan; 2 Department of Virology & Liver unit, Nagoya City<br />

University Graduate School of Medical Sciences, Nagoya, Japan;<br />

3 Department of Gastroenterology and Metabolism, Nagoya City<br />

University Graduate School of Medical Sciences, Nagoya, Japan<br />

Background: We developed highly sensitive chemiluminescent<br />

enzyme immunoassay (CLEIA) for detection of hepatitis B<br />

surface antigen (HBsAg). As recently reported, this assay uses<br />

for quantitative HBsAg detection by combining monoclonal<br />

antibodies, each specific for a different epitope of the antigen,<br />

and employing an improved conjugation technique. Here,<br />

we demonstrate that highly sensitive HBsAg CLEIA (Lumipulse<br />

HBsAg-HQ) is an appropriate and precise way as pre-transfusion<br />

testing to detect occult HBV viremia. Methods: In Japan,<br />

two quantitative HBsAg detection systems are available: Architect<br />

HBsAg-QT (Abbott Japan) (detection range, 50 to 250,000<br />

mIU/mL) and HISCL HBsAg (Sysmex) (detection range, 30 to<br />

2,500,000 mIU/mL). The sensitivity of highly sensitive HBsAg<br />

CLEIA (Lumipulse HBsAg-HQ) (Fujirebio, Inc.) (5150,000,000<br />

mIU/mL) is approximately 6-fold higher than that of HISCL<br />

HBsAg (30 mIU/mL). In this study, the performance of Lumipulse<br />

HBsAg-HQ was compared with that of HISCL HBsAg.<br />

This study protocol was approved by the appropriate institutional<br />

ethics review committees. Results: As pre-transfusion<br />

testing, 4,733 serum samples were measured HBsAg by HISCL<br />

HBsAg in our hospital between July 1st 2013 and February<br />

28th 2015. Of the 4,733 samples, 117 samples (2.47%) were<br />

HBsAg seronegative but showed low concentration of HBsAg<br />

(10 to 20 mIU/mL) as determined by HISCL HBsAg. Of the<br />

117 samples, 17 (14.5%) were positive for anti-hepatitis B<br />

core antigen (anti-HBc), and 15 of the 17 samples were available<br />

for additional assay. They were measured by Lumipulse<br />

HBsAg-HQ retrospectively and 3 (20%) were detectable for<br />

HBsAg. Two samples were from chronic hepatitis B carriers<br />

with HBsAg seroclearance and one was from an adult T-cell<br />

leukemia/lymphoma patient who was diagnosed as HBV reactivation.<br />

Their HBsAg concentrations by Lumipulse HBsAg-HQ<br />

were 8.0, 38.5 and 7.2 mIU/mL, respectively. Conclusion: As<br />

pre-transfusion testing, Lumipulse HBsAg-HQ is very appropriate<br />

and precise system for detection of occult HBV viremia as<br />

well as HBV reactivation at early phase.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Takako Inoue, Noriyo Ochi,<br />

Takaaki Goto, Shintaro Ogawa, Noboru Shinkai, Kumiko Ohne, Yukio<br />

Wakimoto<br />

1574<br />

Collaborative Public Health Efforts to Test and Link Foreign-Born<br />

Persons with Chronic Hepatitis B Virus Infection<br />

to Care<br />

Aaron M. Harris, Ben Schoenbachler, Gilberto Ramirez, Claudia<br />

Vellozzi, Geoff Beckett; Division of Viral Hepatitis, Centers for<br />

Disease Control and Prevention, Atlanta, GA<br />

Background: Hepatitis B virus (HBV) infection continues to<br />

be a public health threat. An estimated 1 million persons are<br />

chronically infected with HBV in the United States. Of these,<br />

70% are foreign-born and more than two-thirds are unaware<br />

of their infection. We launched a public health initiative among<br />

foreign-born persons for HBV testing and linkage to care.<br />

Methods: Nine U.S. sites participated in the initiative during<br />

October 2012–September 2014. Sites partnered with healthcare<br />

centers and community-based organizations for recruitment<br />

of foreign-born persons. Blood samples were collected<br />

and tested for hepatitis B surface antigen (HBsAg). Information<br />

regarding risk factors, follow-up counseling, specialty referrals,<br />

and medical appointment attendance was collected. Results:<br />

Of 23,144 individuals tested across all sites, 1,317 (5.7%)<br />

were HBsAg-positive. Of these, median age was 47 years,<br />

91% had at least one risk factor for HBV infection, 85% were<br />

provided counseling, 83% were referred to care, and 46%<br />

attended a first medical visit. The proportion of HBsAg-positive<br />

persons by region of origin included: Africa (9.7%), Western<br />

Pacific (6.4%), Eastern Mediterranean (5.2%), South-East Asia<br />

(4.9%), South America (2.4%), Eastern Europe (2.3%), and<br />

North America (0.9%). Conclusions: Community-based HBV<br />

testing initiatives can identify substantial numbers of persons<br />

with chronic HBV infection. However, strategies are needed to<br />

improve linkage to HBV-directed medical care for foreign-born<br />

individuals living with chronic HBV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Aaron M. Harris, Ben Schoenbachler,<br />

Gilberto Ramirez, Claudia Vellozzi, Geoff Beckett<br />

1575<br />

Hepatitis B viral load in dried blood spots: a validation<br />

study in Zambia<br />

Michael J. Vinikoor 1,2 , Samuel Zürcher 3 , Kalo Musukuma 4,2 , Obert<br />

Kachuwaire 2 , Andri Rauch 3 , Benjamin Chi 5 , Meri Gorgievski 3 ,<br />

Marcel Zwahlen 3 , Gilles Wandeler 3 ; 1 Medicine, University of<br />

North Carolina at Chapel Hill, Lusaka, Zambia; 2 Centre for Infectious<br />

Disease Research in Zambia, Lusaka, Zambia; 3 University of<br />

Bern, Bern, Switzerland; 4 University of Zambia, Lusaka, Zambia;<br />

5 Obstetrics & Gynecology, University of North Carolina at Chapel<br />

Hill, Chapel Hill, NC<br />

Study Purpose: Access to hepatitis B viral load (VL) testing is<br />

poor in sub-Saharan Africa (SSA) due to economic and logistical<br />

reasons. In a laboratory in Lusaka, Zambia, we compared<br />

hepatitis B VLs between plasma samples and dried blood spots<br />

(DBS), a potential alternative specimen type for such settings.<br />

Methods: Paired plasma and DBS samples from HIV-hepatitis B<br />

virus (HBV) co-infected Zambian adults were analyzed for HBV<br />

VL using COBAS AmpliPrep/COBS Taqman HBV test version<br />

2.0 and genotype by direct sequencing. We used Bland-Altman<br />

analysis to compare VLs between DBS and plasma and by<br />

HBV genotype. Logistic regression analysis was conducted to<br />

assess the probability of an undetectable DBS result by plasma<br />

VL. Results: Among 68 participants, median age was 34 years,<br />

61.8% were men, and plasma HBV VLs ranged from 170,000,000 IU/ml. Among sequenced viruses, 28 were<br />

genotype A1 and 27 were genotype E. Bland-Altman plots<br />

suggested strong agreement between DBS and plasma VLs (see<br />

Figure). On average VLs were 1.59 log 10<br />

IU/ml (95% CI: 1.48<br />

to 1.70) lower in DBS compared to plasma. The probability of<br />

an undetectable DBS result at a plasma VL of ≥2,000 IU/ml<br />

was 1.8% (95% CI: 0.5-6.6) and 0.2% (95% CI: 0.03-1.7) at<br />

≥20,000 IU/ml. Conclusions: We demonstrated the feasibility<br />

and performance of HBV VL testing with DBS in a Zambian<br />

laboratory and observed a strong agreement between sample<br />

types and high sensitivity in DBS at plasma VL ≥2,000 IU/ml.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 979A<br />

As HBV treatment expands, DBS could increase access to HBV<br />

VL testing in SSA settings.<br />

Bland-Altman plot of the agreement between HBV viral loads in<br />

plasma and dried blood spots<br />

Disclosures:<br />

Andri Rauch - Advisory Committees or Review Panels: Gilead Sciences, Abbvie,<br />

MSD, Janssen Cilag<br />

The following authors have nothing to disclose: Michael J. Vinikoor, Samuel<br />

Zürcher, Kalo Musukuma, Obert Kachuwaire, Benjamin Chi, Meri Gorgievski,<br />

Marcel Zwahlen, Gilles Wandeler<br />

1576<br />

Utility of quantitative hepatitis B surface antigen (qHBsAg)<br />

testing for predicting maternal viremia associated<br />

with mother to child transmission of HBV in a multiethnic<br />

cohort of pregnant chronic hepatitis B (CHB) carriers<br />

in Canada<br />

Golasa Samadi Kochaksaraei 1 , Carmen L. Charlton 2 , Stephen<br />

Congly 1 , Trudy Matwiy 1 , Eliana Castillo 1 , Steven R Martin 3 , Carla<br />

S. Coffin 1 ; 1 Medicine, University of Calgary, Calgary, AB, Canada;<br />

2 Provincial Laboratory for Public Health (ProvLab), University<br />

of Alberta Hospital, Edmonton, AB, Canada; 3 Pediatrics, Alberta<br />

Children’s Hospital, University of Calgary, Calgary, AB, Canada<br />

Background and Aims: Mother-to-child transmission of hepatitis<br />

B despite complete passive-active immunoprophylaxis is linked<br />

to high maternal viremia. Available data from our clinic and<br />

others suggest that anti-HBV therapy (i..e, Tenofovir, TDF) in<br />

highly viremic (HBV DNA ≥7 log IU/ml) mothers is safe and<br />

can reduce HBV DNA levels to below the threshold associated<br />

with vaccine failure. Quantitative HBsAg (qHBsAg) monitoring<br />

is increasingly used in management of chronic hepatitis B<br />

(CHB), but there is limited data in pregnancy. We conducted<br />

a prospective observational study to determine whether qHBsAg<br />

can be used as a valid surrogate marker of HBV DNA.<br />

Methods: CHB pregnant patients were recruited from a hepatology<br />

outpatient practice or an obstetrics internal medicine<br />

clinic. Demographics and laboratory data (i.e., HBeAg and<br />

ALT), HBV DNA and quantitative HBsAg were assessed in the<br />

second-third trimester. Statistical analysis was performed by<br />

Spearman’s rank correlation and student’s t-test. Results: In 96<br />

pregnant CHB patients enrolled to date (of which 4 received<br />

TDF therapy), median age 32 years (IQR 28-35), 66% Asian,<br />

23% African, and 11% other (Hispanic or Caucasian). Overall,<br />

22% (21/96) were HBeAg (+), median ALT was 18.5<br />

U/L (IQR 14-31), median HBV DNA was 2.46 log IU/ml (IQR<br />

1.44-3.54). In 69/96 with qHBsAg testing to date, there was a<br />

statistically significant difference in qHBsAg in HBeAg positive<br />

vs. HBeAg negative CHB patients (3.58 log IU/ml [IQR 2.76-<br />

4.28] vs 3.38 log IU/ml [IQR 2.84-3.77], p=0.007) and HBV<br />

DNA (2.57 log IU/ml [IQR 0-7.89] in HBeAg positive vs 2.1<br />

log IU/ml [IQR 1.56-2.92] in HBeAg negative, p=0.0001).<br />

In HBeAg positive patients, there was a significant correlation<br />

between qHBsAg titer and HBV DNA level (r=0.508, p=0.013)<br />

whereas no significant correlation was noted in HBeAg negative<br />

group (r=0.105, p=0.39). In receiver operating characteristic<br />

(ROC) analysis, the optimal qHBsAg cut-off values<br />

for predicting HBV DNA level believed to be associated with<br />

immunoprophylaxis failure (i.e., HBV DNA ≥7 log IU/ml) was<br />

3.58 log IU/ml (accuracy 85.5%, sensitivity 87.5%, specificity<br />

80%). Conclusion: Quantitative HBsAg positively correlates<br />

with serum HBV DNA level in HBeAg positive CHB pregnant<br />

patients and could also be used as a marker for high maternal<br />

viremia, and the need for anti-HBV NA therapy to prevent HBV<br />

immunoprophylaxis failure.<br />

Disclosures:<br />

Carla S. Coffin - Advisory Committees or Review Panels: Janssen, GSK; Grant/<br />

Research Support: BMS, Gilead Sciences, Roche<br />

The following authors have nothing to disclose: Golasa Samadi Kochaksaraei,<br />

Carmen L. Charlton, Stephen Congly, Trudy Matwiy, Eliana Castillo, Steven R<br />

Martin<br />

1577<br />

Nationwide Prospective Survey in Japan to Clarify the<br />

Incidence of Viral Reactivation in Patients with Prior<br />

HBV Infection Receiving Immunosuppressive Agents<br />

Masamitsu Nakao 1 , Nobuaki Nakayama 1 , Yoshihito Uchida 1 ,<br />

Masashi Mizokami 2 , Satoshi Mochida 1 ; 1 Gastroenterology<br />

& Hepatology, Saitama Medical University, Saitama, Japan;<br />

2 Research Center for Hepatitis and Immunology, National Center<br />

for Global Health and Medicine, Ichikawa, Japan<br />

Aim: More than 20% of the population of individuals over 50<br />

years of age in Japan have a history of transient HBV infection.<br />

In such patients, serum HBV-DNA may become detectable following<br />

immunosuppressive therapies, potentially associated<br />

with the development of severe hepatitis. While HBV reactivation<br />

is known to occur at a high incidence in patients receiving<br />

rituximab, the incidence in those with immunosuppressive<br />

agents are yet to be investigated. We conducted 2 types<br />

nationwide survey to determine the incidence of HBV reactivation.<br />

Methods: A total of 420 patients from 59 institutions, who<br />

tested negative for serum HBs-antigen, but positive for serum<br />

anti-HBc and/or anti-HBs antibodies were enrolled, and were<br />

givin immunosuppressive agents such as glucocorticoids (≥0.5<br />

mg/kg of prednisolone), methotrexate and/or biologic agents.<br />

Among them, 131 patiens received the therapies following<br />

the enrollement (Part-1), and the remaining 289 patients were<br />

enrolled at 6 months or later after the initiation of the therapy<br />

when HBV reactivation has not yet developed (Part-2). In all<br />

patients serum HBV-DNA levels were measured evry 1 month,<br />

and the incidences if HBV reactivation, defined as the increase<br />

of serum HBV-DNA levels of 2.1 Log copies/mL or more, were<br />

evaluated by Kaplan-Meier method; within 6 months after the<br />

initiation of therapies in patients in Part-1 study, and later than<br />

6 months in those enrolled in both Part-1 and Part-2 <strong>studies</strong>.<br />

Results: HBV reactivation was found in 5 patients in Part-1<br />

study during mediam periods for 38 (1-62) months, and in 5<br />

patients in Part-2 study after the initiarion of therapies from 7<br />

to 341 months. HBV reactivation occurred even in those who<br />

received monotherapy with methotrexate or glucocorticoid as<br />

well as the therapies with biologic agents. These patients were<br />

given entecavir at least within 1 month after HBV reactivation,<br />

and liver injuries did not occur in all of them. Kaplan-Meier


980A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

analysis revealed that the cumulative incidences of HBV reactivation<br />

were 0.8%, 3.2% and 3.2% at 1, 3 and 6 months,<br />

respectively, after the intiation of therapies through Part-1<br />

study. Also, the analysis thorugh both Part-1 and Part-2 <strong>studies</strong><br />

demonstrated that the increase of incidence later than 6<br />

months were 0%, 0.5%, 1.5% and 1.5% at 12, 24, 36 and 48<br />

months, respectively. Conclusion: HBV reactivation developed<br />

mainly within 6 months after initiation of immunosuppressive<br />

therapies, and the risk was reduced then later. Monitoring of<br />

HBV reactivation by serum HBV-DNA measurement is recommended<br />

to be done especially within 6 months after initiation<br />

of immonusuppresive therapies and/or alteration of agents for<br />

the therapies.<br />

Disclosures:<br />

Yoshihito Uchida - Patent Held/Filed: SRL Inc.<br />

Satoshi Mochida - Grant/Research Support: Chugai, MSD, Tioray Medical,<br />

BMS; Speaking and Teaching: MSD, Toray Medical, BMS, Tanabe Mitsubishi<br />

The following authors have nothing to disclose: Masamitsu Nakao, Nobuaki<br />

Nakayama, Masashi Mizokami<br />

1578<br />

HBV-related cirrhosis in the Chronic Hepatitis Cohort<br />

Study (CHeCS)<br />

Stuart C. Gordon 2,3 , Loralee B. Rupp 1 , Joseph A. Boscarino 4 ,<br />

Mark A. Schmidt 5 , Connie M. Trinacty 6 , Lois Lamerato 7 , Nancy<br />

Oja-Tebbe 7 , Mei Lu 7 ; 1 Center for Health Policy & Health Services<br />

Research, Henry Ford Health System, Detroit, MI; 2 Division of Gastroenterology<br />

and Hepatology, Henry Ford Health System, Detroit,<br />

MI; 3 School of Medicine, Wayne State University, Detroit, MI;<br />

4 Center for Health Research, Geisinger Health System, Danville,<br />

PA; 5 Center for Health Research, Kaiser Permanente Northwest,<br />

Portland, OR; 6 Center for Health Research, Kaiser Permanente<br />

Hawaii, Honolulu, HI; 7 Public Health Sciences, Henry Ford Health<br />

System, Detroit, MI<br />

PURPOSE: The overall spectrum of liver disease among Americans<br />

with chronic hepatitis B (CHB) infection remains unknown,<br />

and staging liver biopsy has become increasingly uncommon.<br />

We used four different methods, including liver biopsy, to<br />

identify potential cirrhosis among CHB patients enrolled in the<br />

Chronic Hepatitis Cohort Study (CHeCS), an ongoing observational<br />

study at four large, integrated health systems in the<br />

United States. METHODS: We included patients who met predefined<br />

inclusion criteria for CHB whose case status had been<br />

confirmed through chart abstraction. We excluded patients<br />

coinfected with hepatitis C and liver transplant recipients. Data<br />

were collected through 2012 from liver biopsy reports, lab<br />

results, and diagnosis/procedure codes contained in the electronic<br />

health record. We calculated FIB-4 scores based on most<br />

recently available aminotransferase and platelet count values<br />

collected when a patient was not on antiviral therapy. Cirrhosis<br />

was identified via four methods: (1) liver biopsy reports, (2)<br />

presence of a diagnosis code for cirrhosis (ICD-9 diagnosis<br />

codes 571.2 and 571.5), (3) presence of a diagnosis or procedure<br />

code for end stage liver disease (ESLD), and (4) FIB-4<br />

score >5.17, a cut-off value previously validated as predictive<br />

of cirrhosis in this cohort. We compared the extent to which<br />

cirrhosis was indicated by each method individually and by<br />

multiple methods. RESULTS: Among the 2,731 CHB patients,<br />

24% were 60 years old, 54% were male, and<br />

53% were Asian. Median follow-up time was 6.0 years. Of the<br />

total, 564 (21%) patients had at least one biopsy report during<br />

follow-up, and of those, 85 (15% of those who had a biopsy)<br />

had a recent biopsy less than two years old. Overall, cirrhosis<br />

was indicated for 520 (19%) patients by at least one of the<br />

four methods: by liver biopsy for 66 patients (2% of patients<br />

overall, and 12% among those with any liver biopsy report);<br />

by FIB-4 score for 281 (10%); by diagnosis code for cirrhosis<br />

for 318 (12%); and by diagnosis or procedure code for ESLD<br />

for 216 (8%), 145 of whom had a diagnosis code for both cirrhosis<br />

and ESLD. Of the 281 patients with FIB-4 score >5.17,<br />

only 26 (9%) had been diagnosed with cirrhosis via biopsy,<br />

135 (48%) via ICD-9 code for cirrhosis, and 113 (40%) via<br />

diagnosis or procedure code indicating ESLD. CONCLUSION:<br />

An estimated 19% of CHB patients among our cohort had<br />

indication of cirrhosis at some point in their follow-up through<br />

2012. The use of additional parameters, including a previously<br />

validated biomarker as a surrogate for more advanced liver<br />

disease, potentially identifies additional unrecognized cirrhosis<br />

in untreated CHB patients.<br />

Disclosures:<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

The following authors have nothing to disclose: Loralee B. Rupp, Joseph A. Boscarino,<br />

Mark A. Schmidt, Connie M. Trinacty, Lois Lamerato, Nancy Oja-Tebbe,<br />

Mei Lu<br />

1579<br />

HBsAg transferring from mother to infant through placenta<br />

account for non-response to hepatitis B vaccine in<br />

infant from HBsAg(+) mother<br />

Jing Wang 1 , Jinfeng Liu 2 , Yingli He 2 , Dongfang Jin 3 , Yuan Yang 2 ,<br />

Li Jin 1 , Taotao Yan 1 , Furong Cao 1 , Shulin Zhang 2 , Yingren Zhao 2 ,<br />

Tianyan Chen 2 ; 1 Xi’an Jiaotong University, Xi’an, China; 2 The First<br />

Affiliated Hospital of Medical College, Xi’an Jiaotong University,<br />

Xi’an, Shaanxi Province, China, Xi’an, China; 3 Shaanxi Kangfu<br />

Hospital, Xi’an, China<br />

Background: It is common for infants from HBsAg(+) mothers<br />

to fail to mounting protective immunity after standard hepatitis<br />

B vaccination and then develop infection. Maternal HBV<br />

markers crossing placenta may lead to neonatal temporal HBV<br />

markers positivity rather than infection, whether these markers<br />

affecting infantile HB vaccine response remains unclear. Methods:<br />

A total of 328 HBsAg(+) mothers and their offspring were<br />

enrolled. All infants received combined immunoprophylaxis<br />

and then followed up to 12 months. HBsAg, anti-HBs titer and<br />

HBV DNA load were assayed. Immunohistochemical staining<br />

for HBsAg in placenta was employed to explore HBsAg transferring.<br />

Results: Firstly, 22 (6.7%) were non-responders to HB<br />

vaccine were identified from those 328 infants. By stastical<br />

analysis, we found that: 1. HBsAg(+) newborns showed higher<br />

risk of non-response than HBsAg(−) infants ( 13.0% vs. 5.0%,<br />

P = 0.016). 2. Infants from high HBsAg titer mothers displayed<br />

higher risk of HBsAg positive at birth than those from low<br />

titer mothers (45.3% vs. 2.8%, P < 0.001). 3. HBsAg titer in<br />

mothers of HBsAg(+) newborns was much higher than mothers<br />

of HBsAg(−) newborn (P < 0.001). All those data supporting<br />

HBsAg can be transferring through placenta. Moreover,<br />

our hypothesis was reinforced by immunostaining with specific<br />

antibody against HBsAg, showing placenta of HBsAg(+)<br />

infants demonstrated higher prevalence (87.5% vs. 30.8%,<br />

P = 0.024) and stronger immunostaining (P = 0.008) than<br />

placenta of HBsAg(−) infants. Conclusion: HBsAg can transfer<br />

though placenta from mother to infant and the neonatal HBsAg<br />

positivity was associated with non-response to HB vaccine.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jing Wang, Jinfeng Liu, Yingli<br />

He, Dongfang Jin, Yuan Yang, Li Jin, Taotao Yan, Furong Cao, Shulin Zhang,<br />

Yingren Zhao, Tianyan Chen


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 981A<br />

1580<br />

Aldehyde dehydrogenases-2 rs671 polymorphism is<br />

associated with an increased risk of hepatocellular carcinoma<br />

in patients with hepatitis B virus<br />

Jing Sun 1 , Hongqin Xu 1 , Xiuting He 1 , Jingyun Wang 1 , Xiaomei<br />

Wang 1 , Bin Gao 2 , Junqi Niu 1 , Yanhang Gao 2 ; 1 The first hospital<br />

of Jilin University, Changchun, China; 2 National Institutes of Health<br />

NIH/NIAAA, Rockville, MD<br />

Background and aims: Hepatocellular carcinoma (HCC) is<br />

the second-most common cancer in China and accounts for<br />

51% of the deaths from liver cancer worldwide. Up to 80% of<br />

HCC cases in China are attributable to hepatitis B virus (HBV),<br />

but the exact pathogenesis is still not completely understood.<br />

Aldehyde dehydrogenase 2(ALDH2) is well known about<br />

its role on detoxifying aldehydes in ethanol metabolism. An<br />

ALDH2 inactivating mutation is the most common single point<br />

mutation in humans, mostly found in East Asians. There is a<br />

good geographical correlation between the prevalence of the<br />

ALDH2 gene mutation and HBV infection, but the relationship<br />

between these two factors is unclear. Methods: In the present<br />

study, we investigated the association between ALDH2 (rs671,<br />

rs10849970 and rs886205) polymorphisms, chronic alcohol<br />

consumption, and the risk of HBV-related HCC among Han<br />

populations in Northeast China in a case-control study (2012-<br />

2015). A total of 930 subjects were included in the study,<br />

including 102 cases with HBV-related chronic hepatitis, 264<br />

cases with HBV-related cirrhosis, 281 cases with HBV-related<br />

HCC, and 283 healthy individuals as controls. The odds ratios<br />

(OR) and corresponding 95% confidence intervals (CI) were<br />

calculated using logistic regression models. Results: ALDH2<br />

rs671 includes three genotypes, GG (wild type), AA (homozygous<br />

mutation), GA (heterozygous mutation). The distribution<br />

difference of genotype frequencies of ALDH2 rs671 in the four<br />

groups we studied was statistically significant (P


982A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

between selected parameters of lipid metabolism and hepatic<br />

steatosis measured using CAP option of transient elastography<br />

(FibroScan), serum HBV-DNA and HBsAg levels during treatment<br />

with tenofovir (TDF), after switching from any antiHBV<br />

nucleos(t)ide analogue (NUC). Methods: The study included<br />

45 HBsAg-positive and HBeAg-negative pts, who within last<br />

3 months have been switched to TDF due to the lack of efficacy<br />

or intolerance to prior treatment. We evaluated the level<br />

of total cholesterol, low-density lipoprotein (LDL), high-density<br />

lipoprotein (HDL), triglycerides (TG), HBV-DNA and HBsAg<br />

levels. Liver steatosis and fibrosis were assessed FibroScan with<br />

CAP option. In statistical analysis parametric Pearson correlation<br />

and Rang Spearman correlation were used. Results were<br />

considered statistically significant if p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 983A<br />

and 6.15-log copies/mL (4.0-above 8.1) among HBeAg-negative<br />

patients. Among HBeAg positive and negative patients,<br />

the percent ratios of cases whose HBV DNA concentration<br />

fell below 3-log copies/mL were 29.0/60.7 at 3 months,<br />

58.1/96.4 at 6 months, 87.1/100 at 12 months, 90.3/100<br />

at 18 months, and 90.3/100 at 24 months of therapy. The percent<br />

ratios of cases whose HBV DNA levels were undetectable<br />

were 0/7.1 at 3 months, 3.2/17.9 at 6 months, 12.9/23.1<br />

at 12 months, 9.7/53.6 at 18 months, and 25.8/71.4 at 24<br />

months of therapy. (2) A strong correlation of HBsAg values<br />

between the Architect and Lumipulse assays was observed in<br />

both HBeAg-positive and HBeAg-negative patients (r=0.88 and<br />

0.93, respectively). However, more than 1-log reduction of<br />

HBsAg values (IU/mL in the Architect and cutoff index [COI] in<br />

the Lumipulse) after 24 months of therapy was observed in only<br />

11.9% of samples by either assay. Conclusion: HBsAg values<br />

measured by the Architect and Lumipulse assays were directly<br />

correlated. However, HBsAg values decreased slowly during<br />

the 2-year ETV therapy whereas HBV DNA concentrations<br />

decreased more than 3-log copies/mL in all patients. Thus,<br />

neither of these assays are effective predictors of the efficacy<br />

of short-term ETV therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kazuma Shinkai, Hisashi Ishida,<br />

Ryousuke Kiyota, Taku Tashiro, Kentarou Nakagawa, Keisuke Fukutomi, Akio<br />

Ishihara, Tetsuya Iwasaki, Ryuichiro Iwasaki, Kumiko Nishio, Yuko Sakakibara,<br />

Takuya Yamada, Shoichi Nakazuru, Eiji Mita<br />

qHBsAg: -0.33 log IU/mL vs. -0.06 log IU/mL, p=0.02; Figure).<br />

The area under the curve for a prediction model based<br />

on week 12 for HBcrAg, ALT, and qHBsAg was 0.70 (CI-95%<br />

0.57-0.83, p=0.005). Conclusion. On-treatment HBcrAg level<br />

decline in the first 12 weeks is associated with response to<br />

peginterferon in HBeAg-negative CHB patients. HBcrAg levels<br />

are strongly correlated with HBV DNA.<br />

Disclosures:<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

Andre Boonstra - Grant/Research Support: BMS, Janssen Pharmaceutics, Merck,<br />

Roche, Gilead<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Margo J. van Campenhout, Willem<br />

Pieter Brouwer, Vincent Rijckborst, Bettina E. Hansen<br />

1585<br />

Hepatitis B Core-Related Antigen Level Decline in the<br />

First 12 Weeks of Peginterferon Treatment is Associated<br />

with Response in HBeAg-Negative Chronic Hepatitis B<br />

Margo J. van Campenhout 1 , Willem Pieter Brouwer 1 , Vincent Rijckborst<br />

1 , Robert J. de Knegt 1 , Andre Boonstra 1 , Harry L. Janssen 2,1 ,<br />

Bettina E. Hansen 1,3 ; 1 Gastroenterology & Hepatology, Erasmus<br />

MC University Medical Center Rotterdam, Rotterdam, Netherlands;<br />

2 Toronto Centre for Liver Disease, Toronto Western and General<br />

Hospital, University Health Network, Toronto, ON, Canada; 3 Public<br />

Health, Erasmus MC University Medical Center Rotterdam, Rotterdam,<br />

Netherlands<br />

Background. Markers for prediction of response in HBeAg-negative<br />

chronic hepatitis B (CHB) patients are scarce. Hepatitis B<br />

core-related antigen (HBcrAg), which is a new serum marker<br />

for the combined measure of HBcAg, HBeAg, and p22cr,<br />

may be useful for response prediction as it correlates with<br />

cccDNA in HBeAg-negative patients. Methods. We studied<br />

133 HBeAg-negative CHB patients treated with 48 weeks of<br />

peginterferon alfa-2a +/- ribavirin. Serum HBcrAg levels were<br />

measured at baseline and at week 12, and we assessed the<br />

correlation with sustained response (SR; ALT normalization<br />

& HBV DNA


984A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

was lower in patients with F4-F6 (n=27) compared to those<br />

with F0-F3 (7.6 vs. 8.1 log U/mL, p=0.02), and declined with<br />

increasing fibrosis stage (Figure). HBcrAg levels were independently<br />

associated with advanced fibrosis when adjusting<br />

for age and serum ALT (OR 0.5, CI-95% 0.3-0.8, p=0.003).<br />

In HBeAg-negative patients, fibrosis stage was available for<br />

all patients. The mean HBcrAg was 5.5 (SD 0.1), median HAI<br />

score 5 (IQR 3), and median Ishak fibrosis stage 3 (IQR 2).<br />

Genotypes A/B/C/D/other were present in 13/1/2/81/3%.<br />

In these patients, log HBcrAg was not correlated with Ishak<br />

fibrosis score (r s<br />

=-0.1, p=0.59), similar to fibrosis correlation<br />

to qHBsAg (r s<br />

=-0.1, p=0.19), and HBV DNA (r s<br />

=0.2, 0.07).<br />

When adjusting for age and serum ALT, HBcrAg levels were<br />

not associated with advanced fibrosis (OR 0.9, CI-95% 0.5-<br />

1.6, p=0.83) and did not change with increasing fibrosis<br />

stage (Figure). For both HBeAg-positive and HBeAg-negative<br />

patients, variation of qHBsAg levels across fibrosis stages<br />

was comparable to that of HBcrAg levels (Figure). Conclusion.<br />

In this study, serum HBcrAg levels were inversely correlated<br />

with Ishak fibrosis stage in HBeAg-positive CHB, but not in<br />

HBeAg-negative CHB.<br />

obtainable mutation analysis in 2 centers. Of 177 HBeAg(+)<br />

patients, 6(3.39%) were genotype A, 65(36.72%) genotype<br />

B, 104(58.76%) genotype C, 1(0.57%) genotype B/C, and<br />

1(0.57%) genotype C/D. Of 90 HBeAg(-) patients, 1(1.11%)<br />

were genotype A, 50(55.56%) genotype B, 37(41.11%) genotype<br />

C and 2(2.22%) genotype D. When compared to HBeAg(-)<br />

patients, HBeAg(+) patients were significantly younger in mean<br />

age (37.93 vs. 44.40; P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 985A<br />

able on post-delivery flares in patients treated with tenofovir<br />

(TDF) during pregnancy. Aim: Cross-sectional observational,<br />

single centre study to compare the frequency of post-delivery<br />

flares in untreated CHB mothers and in CHB patients<br />

requiring therapy during pregnancy (due to liver disease or<br />

prevent HBV transmission) and evaluate the importance virological<br />

(HBV DNA), serological (HBeAg and HBsAg levels)<br />

and immunological markers (IP10 levels) during pregnancy<br />

(2 nd trimester) on predicting post-delivery flares. Patients: 297<br />

CHB patients (median age 31.1years, 16% HBeAg+) were<br />

assessed at 2 nd pregnancy trimester (1 st visit median gestation<br />

week 26); 35 patients were treated with TDF prior to pregnancy,<br />

39 patients started TDF from gestation week 28 to prevent<br />

HBV transmission and 223 untreated patients (all HBV<br />

DNA38IU/l and increase by two-fold pregnancy level. 36<br />

patients stopped therapy shortly after delivery. Results: Treated<br />

patients had higher frequency of flares than untreated mothers<br />

(51% vs. 26%,p200pg/ml) (OR<br />

4.6, 1.8-11.4), HBV DNA (>10000IU/ml) (OR 9.2, 4.9-16.9),<br />

HBsAg (>20000IU/ml) (OR 5.5, 2.5-12.2) and higher HBeAg<br />

(>1000S/CO) levels (OR 5.7, 0.5-60) at 2 nd pregnancy trimester.<br />

While only high IP10 was predictive of HF in untreated<br />

mothers (OR 4.2, 1.5-10.2); higher HBV DNA, HBsAg, HBeAg<br />

and IP10 levels and younger age ( Methods: Based on Taiwan National Health Insurance<br />

Research Database, 176,120 patients diagnosed with RA or<br />

psoriasis were screened for eligibility. Only those having past<br />

HBV infection were screened. After excluding those having<br />

other hepatitis, taking antiviral drugs for viral hepatitis or those<br />

with malignant diseases, we enrolled a total of 354 eligible<br />

patients who received TNF-α antagonists (biologics cohort)<br />

and matched them 1: 2 with 708 patients who received nbD-<br />

MARDs alone by baseline characteristics and propensity scores<br />

(DMARDs cohort). Both cohorts were followed up for the occurrence<br />

of HBV hepatitis flare after starting TNF-α antagonists.<br />

Results: The 5-year cumulative incidences of HBV hepatitis flare<br />

were 19.3% (95% confidence interval [CI] 13.8-24.9%) and<br />

12.9% (95% CI 9.2-16.6%) for the Biologics and DMARDs<br />

cohorts, respectively (P=0.004). On modified Cox proportional<br />

hazards analysis, patients in the biologics cohort had<br />

significantly higher risk of HBV hepatitis flare (adjusted HR<br />

2.10, 95% CI 1.3-3.3, P Conclusion:<br />

Biologics use is associated with significantly higher risk of<br />

HBV hepatitis flare than DMARDs in RA or psoriasis patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chun-Ying Wu, Yi-Ju Chen, Jaw-<br />

Town Lin<br />

1590<br />

The PAGE-B Score Accurately Predicts Clinical Outcome<br />

and Outperforms Other Biomarkers over 15 Years of<br />

Follow-up in a Diverse Cohort of Chronic Hepatitis B<br />

Patients<br />

Willem Pieter Brouwer 1 , Adriaan J. van der Meer 1 , Andre Boonstra<br />

1 , Elisabeth P. Plompen 1 , Suzan D. Pas 1 , Robert J. de Knegt 1 ,<br />

Robert A. de Man 1 , Fiebo J. ten Kate 1 , Harry L. Janssen 2,1 , Bettina<br />

E. Hansen 1 ; 1 Gastroenterology & Hepatology, Erasmus Medical<br />

Center Rotterdam, Rotterdam, Netherlands; 2 Gastroenterology and<br />

Hepatology, UHN Liver Clinic, Toronto, ON, Canada<br />

Background & aims. Multiple non-invasive markers have been<br />

associated with hepatocellular carcinoma (HCC) development<br />

in chronic hepatitis B (CHB) patients. We aimed to compare the<br />

prognostic performance of these markers and to assess whether<br />

liver histology could improve this performance. Methods. Liver<br />

biopsies from consecutive treatment-naïve CHB patients were<br />

scored by a single experienced hepato-pathologist. Laboratory<br />

values at the time of biopsy were used to calculate the PAGE-B,<br />

REACH-B, GAG-HCC, and CU-HCC scores. A clinical event<br />

was defined as a combined endpoint including HCC development,<br />

liver failure, liver transplantation, and all-cause mortality.<br />

Event data was obtained from national database registries.<br />

Results. Of 557 patients, mean age at biopsy was 35±13<br />

years, 47/31/19% was Caucasian/Asian/African, 63%<br />

received antiviral therapy (AVT) after liver biopsy and 113<br />

(20%) had advanced fibrosis (F3/F4). Forty patients developed<br />

a clinical event within a median follow-up of 10.1 years<br />

(IQR 5.7-15.9, complete follow-up 93%). The PAGE-B score<br />

predicted any clinical outcome (C-statistic [C] 0.87, 95%CI:<br />

0.81-0.92), HCC development (C: 0.89) and reduced transplant-free<br />

survival (C: 0.84) with good accuracy, also when<br />

stratified by ethnicity, AVT after biopsy or advanced fibrosis.<br />

The REACH-B, GAG-HCC, and CU-HCC predicted clinical outcome<br />

less accurate (C: 0.70, 0.82, and 0.73, respectively).


986A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The event prediction with the PAGE-B improved modestly with<br />

the Ishak fibrosis score (C: 0.88), but the net reclassification<br />

improvement appeared not clinically significant. Conclusion.<br />

The PAGE-B score showed the best performance to assess the<br />

likelihood to develop a clinical event among a diverse CHB<br />

population over 15 years of follow-up. Additional liver histologic<br />

characteristics did not appear to improve this event risk<br />

assessment.<br />

27.6%, 12.1%, 4.0% and 1.4% for patients with LC aged ≥<br />

50 years, those with LC aged < 50 years, those without LC<br />

aged ≥ 50 years, and those without LC aged < 50 years,<br />

respectively. HBV DNA levels and quantitative hepatitis B surface<br />

antigen (qHBsAg) levels were independent predictors for<br />

inactive carriers in patients without LC. The 5- year cumulative<br />

incidence of an inactive carrier was 67.9% in non-LC patients<br />

with both low qHBsAg and HBV DNA levels, while the disease<br />

progression rate was 2.2 %. Conclusions: HBeAg negative<br />

patients without LC can be monitored for becoming an inactive<br />

carrier when both HBV DNA levels and qHBsAg levels are<br />

low, as the risk of disease progression is low while incidence<br />

of inactive carrier is high.<br />

Cumulative incidence of inactive carriers according to HBV DNA<br />

and qHBsAg levels in patients without cirrhosis. A-D represent low<br />

HBV DNA/low qHBsAg, low HBV DNA/high qHBsAg, high HBV<br />

DNA/low qHBsAg and high HBV DNA/high qHBsAg, respectively.<br />

Disclosures:<br />

Adriaan J. van der Meer - Consulting: Gilead; Speaking and Teaching: MSD,<br />

Gilead<br />

Andre Boonstra - Grant/Research Support: BMS, Janssen Pharmaceutics, Merck,<br />

Roche, Gilead<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

Robert A. de Man - Advisory Committees or Review Panels: Norgine; Grant/<br />

Research Support: Biotest<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Willem Pieter Brouwer, Elisabeth<br />

P. Plompen, Suzan D. Pas, Fiebo J. ten Kate, Bettina E. Hansen<br />

1591<br />

Prediction of Clinical Outcomes in Hepatitis B E Antigen<br />

Negative Chronic Hepatitis B Patients with Elevated<br />

Hepatitis B Virus DNA Levels<br />

Jem Ma Ahn, Dong Hyun Sinn, Geum-Youn Gwak, Yong Han<br />

Paik, Moon Seok Choi, Joon Hyeok Lee, Kwang Cheol Koh, Seung<br />

Woon Paik; Department of Medicine, Samsung Medical Center,<br />

Sungkyunkwan University School of Medicine, Seoul, Korea (the<br />

Republic of)<br />

We investigated whether long-term clinical outcomes such as<br />

disease progression or inactive hepatitis B virus (HBV) carrier<br />

can be predicted by baseline factors in hepatitis B e antigen<br />

(HBeAg)-negative HBV infected patients with elevated viral<br />

loads. A retrospective cohort of 527 HBeAg-negative chronic<br />

HBV infected patients with elevated viral loads (HBV DNA ≥<br />

2,000 IU/ml) were assessed for disease progression defined<br />

by development of hepatocellular carcinoma or cirrhosis (LC)<br />

complications, as well as becoming an inactive carrier, defined<br />

by a decrease of HBV DNA < 2,000 IU/ml for ≥ 6 months in<br />

absence of antiviral therapy. During a median 3.6 years of<br />

follow-up, disease progression was detected in 46 patients,<br />

and 79 patients became inactive carriers. Old age and cirrhosis<br />

were independent predictors for disease progression.<br />

The 5-year cumulative incidence of disease progression were<br />

Disclosures:<br />

The following authors have nothing to disclose: Jem Ma Ahn, Dong Hyun Sinn,<br />

Geum-Youn Gwak, Yong Han Paik, Moon Seok Choi, Joon Hyeok Lee, Kwang<br />

Cheol Koh, Seung Woon Paik<br />

1592<br />

The Role of Surface Antibody in Hepatitis B Virus Reactivation<br />

in Patients with Resolved Infection: A Systematic<br />

Review and Meta-Analysis<br />

Sonali Paul 1 , Aaron Dickstein 1 , Akriti P. Saxena 1 , Norma Terrin 1 ,<br />

Kathleen Viveiros 1 , Ethan M. Balk 2 , John B. Wong 1 ; 1 Tufts Medical<br />

Center, Cambridge, MA; 2 Brown University School of Public<br />

Health, Providence, RI<br />

Background: Patients with resolved hepatitis B virus (HBV) and<br />

hematological malignancies are at risk for HBV reactivation.<br />

Data about whether hepatitis B surface antibody (HBsAb+)<br />

reduces the risk of reactivation are conflicting and limited.<br />

This systematic review with meta-analysis assesses whether the<br />

presence of HBsAb+ reduces the risk of HBV reactivation in<br />

patients with resolved HBV infection (negative surface antigen,<br />

HBsAg-, and positive core antibody, HBcAb+) receiving<br />

chemotherapy for hematological malignancies. Methods: We<br />

included English language <strong>studies</strong> through March 31, 2015<br />

that examined reactivation in patients with resolved HBV<br />

infection receiving chemotherapy for all hematological malignancies<br />

from MEDLINE, Web of Science, Cochrane Central<br />

Register of Controlled Trials, TOXNET, and Scopus. The pri-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 987A<br />

mary outcome was HBV reactivation defined as an increase<br />

in HBV DNA levels from baseline or re-emergence of HBsAg+<br />

when previously negative. The odds ratio (OR) of reactivation<br />

with versus without HBsAb+ was estimated with random-effects<br />

model meta-analyses. Results: Fourteen observational <strong>studies</strong><br />

with 836 patients (range 8-354 with 7 patients receiving antiviral<br />

prophylaxis) met inclusion criteria. There were 9 prospective<br />

and 5 retrospective <strong>studies</strong>; 11 were from Asia and 3 from<br />

Italy; 10 included only lymphoma and 4 had mixed hematologic<br />

malignancies including leukemia and multiple myeloma.<br />

When reported, the median patient age across <strong>studies</strong> was 62<br />

years (range 18-90) with 334 males and 261 females. Chemotherapy<br />

regimens varied across <strong>studies</strong> with rituximab in 8<br />

<strong>studies</strong>. The overall pooled OR was 0.20 (95% CI 0.11–0.35;<br />

with no heterogeneity, I 2 =0%) indicating a protective effect<br />

of HBsAb+ on the risk of reactivation versus HBcAb+ alone.<br />

The absolute risk of reactivation with HBcAb+ alone was 17%<br />

(95% CI 10-25%). Similar results were found when the analysis<br />

was limited to rituximab-based chemotherapy (OR 0.17, 95%<br />

CI 0.08–0.34) and lymphoma (OR 0.17, 95% CI 0.09–0.32).<br />

Sensitivity analysis found a significant protective effect of<br />

HBsAb+ in 9 prospective <strong>studies</strong> but not in 5 retrospective <strong>studies</strong>.<br />

The ORs continued to remain significant when analyzed by<br />

geographic region (Asia OR 0.22, 95% CI 0.12–0.40; Italy<br />

OR 0.07, 95% CI 0.01–0.54). Conclusions: In patients with<br />

resolved HBV receiving chemotherapy for hematologic tumors,<br />

HBsAb+ decreases the risk of reactivation. Our results support<br />

the need for future <strong>studies</strong> examining the effect of HBsAb titers<br />

and booster vaccinations prior to chemotherapy in this patient<br />

population.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sonali Paul, Aaron Dickstein,<br />

Akriti P. Saxena, Norma Terrin, Kathleen Viveiros, Ethan M. Balk, John B. Wong<br />

and the death certification system was performed. Cox proportional<br />

hazards models were used to estimate the multivariate-adjusted<br />

hazard ratio of developing LC and HCC. Results:<br />

Incidence rates of LC per 100,000 person-years were 341.2,<br />

565.5, 654.1, 920.0, 1477.9 for serum anti-HBc levels of<br />


988A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

D. At baseline, serum levels of HBV DNA, HBsAg and HBV<br />

RNA did not differ in patients with and without subsequent<br />

HBeAg SC. On treatment, HBV DNA and HBsAg levels did not<br />

differ between patients with and without HBeAg SC (Table).<br />

However, HBeAg positive patients achieving HBeAg SC had<br />

rapid HBV RNA declines by week 12 with most below LLOD by<br />

week 24. In contrast, HBeAg positive patients without SC had<br />

significantly higher levels of HBV RNA at all time points tested<br />

during treatment (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 989A<br />

tive samples. Out of which 280 (48.4%) samples were found<br />

to be HBV DNA positive and genotyping was done for 228<br />

samples. The frequency distribution of genotypes were genotype<br />

D=214 (93.8%), genotype C=12 ( 5.2%), Genotype<br />

I=2(0.8%). Genotype D (93.8%) was found to be predominant<br />

genotype followed by C (5.2%) in the North eastern states.<br />

Conclusion:- HBV prevalence rate in North eastern states was<br />

found to be 5.9%. Genotype D (93.8%) was found to be predominant<br />

genotype followed by C (5.2%) in the North eastern<br />

states. This analysis of population-based data provides<br />

evidence of the prevalence of HBV infection in North eastern<br />

states. This important finding adds to the existing data of HBV<br />

prevalence in India. These results are relevant to public health<br />

policy makers and highlight the importance of promoting hepatitis<br />

B vaccination programs in the endemic areas of high HBV<br />

prevalence.<br />

Disclosures:<br />

The following authors have nothing to disclose: Premashis Kar, Vijay Karra<br />

1597<br />

Performance of the new cobas ® HBV Assay on the<br />

cobas ® 6800 system in comparison to different<br />

TaqMan ® assays<br />

Benjamin Maasoumy 1 , Kevin Luk 2 , Birgit Bremer 1 , Patrick Lehmann<br />

1 , Ed G. Marins 2 , Veronique Michel-Treil 3 , Ekaterina Gelman<br />

3 , Merlin Njoya 2 , Jesse A. Canchola 2 , Christian O. Simon 4 ,<br />

Heiner Wedemeyer 1 ; 1 Gastroenterology, Hepatology and Endocrinology,<br />

Hannover Medical School, Hannover, Germany; 2 Roche<br />

Molecular Systems Inc., Roche Diagnostics, Pleasanton, CA;<br />

3 Covance Central Laboratory Services SA, Geneva, Switzerland;<br />

4 Roche Diagnostics International, Ltd, Rotkreuz, Switzerland<br />

Background: Management of patients with hepatitis B (HBV)<br />

critically relies on robust and sensitive HBV DNA assays. Various<br />

commercial HBV DNA assays that are based on amplifying<br />

and detecting HBV DNA have received CE Mark or FDA<br />

Approval. In 2014, the new cobas ® 6800/8800 systems were<br />

introduced and are now used in routine diagnostic practice.<br />

For this platform, a new HBV viral load test was developed.<br />

We here aimed to investigate the performance of this new<br />

HBV DNA assay in a routine diagnostic setting. Methods: We<br />

evaluated the performance of the new cobas ® HBV test on the<br />

cobas ® 6800 system in comparison to the version 2.0 of the<br />

COBAS ® AmpliPrep/ COBAS ® TaqMan ® HBV test with automated<br />

nucleic acid extraction (HBV TaqMan ® v2) and to the<br />

manual extraction COBAS ® TaqMan ® HBV test For Use With<br />

The High Pure System (HBV HPS). Up to 336 HBV WHO 3 rd<br />

International Standard dilutions as well as 285 clinical samples<br />

were tested each with all the three assays. Results: The lower<br />

limits of detection were 2.7 IU/ml, 9.6 IU/ml and 6.2 IU/<br />

ml for cobas ® HBV, HBV TaqMan ® v2 and HBV HPS assays,<br />

respectively, based on up to 48 samples tested for each dilution<br />

(2000 IU/mL, 200 IU/mL, 60 IU/ml, 20 IU/mL, 10 IU/<br />

mL, 5 IU/mL, and 2.5 IU/mL). Cobas ® HBV detected 48/48<br />

samples with a nominal titer of 5 IU/mL and 40/48 samples<br />

with a titer of 2.5 IU/mL. The assay was highly reproducible<br />

and linear even in samples with very low viremia with an average<br />

log10 IU/mL difference between observed and nominal<br />

titers of 0.04 (respective values were 0.05 and 0.18 for the<br />

HBV TaqMan ® v2 and the HBV HPS assays). Parallel testing<br />

of clinical samples showed largely comparable values across<br />

different viral loads with minimally lower values revealed by<br />

cobas ® HBV compared to HBV TaqMan ® v2 (mean of paired<br />

difference -0.099 log10 IU/mL; 95% CI -0.132; -0.065), while<br />

no significant difference was observed compared to the HBV<br />

HPS (mean of paired difference -0.011 log10 IU/mL; 95% CI<br />

-0.044; +0.023). Classification of samples at critical treatment<br />

decision cut-offs showed an overall percentage of agreement<br />

between the cobas ® HBV and the HBV TaqMan ® v2 assays<br />

of 94.4% for both the 2000 IU/mL and 20000 IU/mL thresholds<br />

while the respective values were 95.4% and 97.3% for<br />

cobas ® HBV vs. HBV HPS. Conclusions: The new cobas ® HBV<br />

test demonstrated an improved sensitivity compared to current<br />

TaqMan ® HBV assays, however, the clinical meaning of very<br />

low HBV viremia detected with new assay needs to be determined.<br />

Quantitative HBV DNA values are highly comparable<br />

between the assays, which demonstrates that the cobas ® HBV<br />

test can be reliably used in clinical practice to guide treatment<br />

decisions.<br />

Disclosures:<br />

Benjamin Maasoumy - Advisory Committees or Review Panels: Abbott Molecular,<br />

Janssen-Cilaq; Grant/Research Support: Abbott Molecular; Speaking and Teaching:<br />

MSD, Roche Diagnostics, Roche Pharma, Janssen-Cilaq, Fujirebio, BMS<br />

Ed G. Marins - Employment: Roche Molecular Systems<br />

Merlin Njoya - Employment: Roche<br />

Jesse A. Canchola - Employment: Roche Molecular Diagnostics Inc.<br />

Christian O. Simon - Employment: Roche Diagnostics International<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

The following authors have nothing to disclose: Kevin Luk, Birgit Bremer, Patrick<br />

Lehmann, Veronique Michel-Treil, Ekaterina Gelman<br />

1598<br />

Seroprevalence and Clinical Features of Hepatitis D<br />

Virus (HDV) Infection in a North American Cohort<br />

Norah Terrault 1 , Marc G. Ghany 2 , Chaeryon Kang 3 , Stewart<br />

Cooper 4 , Adrian M. Di Bisceglie 5 , Michael W. Fried 6 , Steven<br />

H. Belle 3 , Jay H. Hoofnagle 2 ; 1 UCSF, San Francisco, CA; 2 NIH-<br />

NIDDK, Bethesda, MD; 3 University of Pittsburgh, Pittsburgh, PA;<br />

4 California Pacific Medical Center, San Francisco, CA; 5 St. Louis<br />

University, St. Louis, MO; 6 University of North Carolina, Chapel<br />

Hill, NC<br />

Background Hepatitis D virus (HDV) infection is estimated to<br />

affect ~4-5% (15 million) of persons with chronic hepatitis B<br />

virus (HBV) worldwide. The burden and clinical characteristics<br />

of HDV infection in North America is unknown. Aim: To determine<br />

the prevalence and clinical features of HDV in a well-characterized<br />

cohort of adults in North America with chronic HBV.<br />

Methods: All HBsAg-positive adults in the HBRN cohort with<br />

serum samples at enrollment were tested for total anti-HDV<br />

using a commercially available assay (DiaSorin, Italy). Prevalence<br />

of HDV was examined with respect to demographic and<br />

clinical features at enrollment. Results: Of 1165 adults tested,<br />

30 (2.6%) were anti-HDV positive and 4 (0.3%) equivocal. For<br />

analysis of clinical features, 14 with HCV (13 HDV-negative;<br />

one triple-infected HBV/HCV/HDV) were excluded. Prevalence<br />

of anti-HDV positivity was similar in men (2.3%) and women<br />

(2.7%) and among whites (3.2%), blacks (4.9%) and Asians<br />

(1.8%) (p=0.08). Prevalence of HDV did not differ significantly<br />

by age (median 42 vs. 50 yrs) or presumed route of infection.<br />

Prevalence differed by genotype (p=0.001) being highest in<br />

participants with genotypes D (10.7%) and E (8.0%) which<br />

was also reflected in the differing prevalence of HDV by region<br />

of birth (p


990A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ALT or AST levels, platelet counts, HBsAg levels, or HBeAg<br />

prevalence were similar between HDV positive and negative<br />

participants. Conclusions: HDV coinfection was infrequent<br />

(~3%) among adults in this North American HBV cohort. A distinctive<br />

risk profile for acquisition was not apparent, highlighting<br />

the challenges in using a risk-based screening algorithm to<br />

identify HDV-infected adults.<br />

Laboratory/Viral Features<br />

*median (range)<br />

Disclosures:<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

Adrian M. Di Bisceglie - Advisory Committees or Review Panels: Gilead, AbbVie,<br />

Novartis, Bayer, BTG; Grant/Research Support: Gilead, AbbVie<br />

Michael W. Fried - Consulting: Merck, Abbvie, Janssen, Bristol Myers Squibb,<br />

Gilead; Grant/Research Support: Merck, AbbVie, Janssen, Bristol Myers Squibb,<br />

Gilead; Patent Held/Filed: HCCPlex<br />

Steven H. Belle - Grant/Research Support: Rottapharm!Madaus<br />

The following authors have nothing to disclose: Marc G. Ghany, Chaeryon Kang,<br />

Stewart Cooper, Jay H. Hoofnagle<br />

1599<br />

Comparative Study for Anti-Hepatitis B Surface Antigen<br />

Titers Based on Two Measurement Methods: Using<br />

Monoclonal Antibodies Isolated from Hepatitis B Vaccinated<br />

Japanese Recipients.<br />

Takako Inoue 1 , Shuko Murakami 2 , Susumu Tsutsumi 2 , Kumiko<br />

Ohne 1 , Kazuto Tajiri 3 , Hiroyuki Kishi 4 , Shintaro Ogawa 2 , Noboru<br />

Shinkai 5 , Takaaki Goto 1 , Yukio Wakimoto 1 , Yasuhito Tanaka 2,1 ;<br />

1 Clinical Laboratory, Nagoya City University Hospital, Nagoya,<br />

Japan; 2 Department of Virology & Liver unit, Nagoya City University<br />

Graduate School of Medical Sciences, Nagoya, Japan;<br />

3 Department of Immunology, University of Toyama, Toyama, Japan;<br />

4 Department of Immunology, Graduate School of Medicine and<br />

Pharmaceutical Sciences, University of Toyama, Toyama, Japan;<br />

5 Department of Gastroenterology and Metabolism, Nagoya City<br />

University Graduate School of Medical Sciences, Nagoya, Japan<br />

Background: Since anti-hepatitis B surface antigen (anti-HBs)<br />

titers are various depending on measurement methods, we<br />

compared two different methods to quantity anti-HBs titers in<br />

sera and HBs monoclonal antibodies. Methods: 1) The measurement<br />

of anti-HBs was compared using either Lumipulse<br />

G1200 (Fujirebio Inc.) or Architect i2000SR (Abbott Japan).<br />

Previously, we isolated human monoclonal antibodies (mAbs)<br />

against HBV from Japanese healthy volunteers who had been<br />

immunized with a genotype C recombinant HBV vaccine<br />

(Biimugen, Kaketsuken), using a cell-microarray system. A following<br />

report revealed that among these mAbs, HB0116 and<br />

HB0478, recognize the first N-terminal peptide loop within<br />

the “a” determinant and have HBV-neutralizing activities. In<br />

this study, HB0116, HB0478 and four other mAbs were used.<br />

2) The sera from 182 HBV-resolved patients in our hospital,<br />

who were negative for hepatitis B surface antigen (HBsAg)<br />

but positive for anti-hepatitis B core antigen (anti-HBc) and/or<br />

anti-HBs, were obtained. This study protocol was approved by<br />

the appropriate institutional ethics review committees. Results:<br />

1) Measuring 2 mAbs (HB0116 and HB0478) with HBV-neutralizing<br />

activities, the anti-HBs titers of HB0116 (1.0 μg/mL)<br />

were comparable (440.7 mIU/mL by Architect and 689.3<br />

mIU/mL by Lumipulse), whereas those of HB0478 (1.0 mg/<br />

mL) were much lower by Architect than by Lumipulse (42.6<br />

mIU/mL by Architect and 818.6 mIU/mL by Lumipulse). Of<br />

the other 4 mAbs without HBV-neutralizing activities, equal<br />

titers were observed for one; two mAbs had less anti-HBs titers<br />

by Architect; and one was below the cut-off index (


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 991A<br />

(95%CI 1.01 − 1.09), p=0.0018), AFP > 20 ng/mL (HR 3.2<br />

(95%CI 1.01 − 9.99), p=0.049), WFA(+)-M2BP (per 1 COI)<br />

(HR 1.3 (95%CI 1.01 – 1.63), p=0.039) as independent risk<br />

factors for the development of HCC. Conclusion: WFA(+)-M2BP<br />

can be applied as a simple and reliable surrogate marker for<br />

the risk of HCC development, in addition to liver fibrosis stage<br />

as determined by liver biopsy.<br />

Disclosures:<br />

Seigo Abiru - Grant/Research Support: CHUGAI PHARMACEUTICAL CO.,LTD<br />

The following authors have nothing to disclose: Kazumi Yamasaki, Shigemune<br />

Bekki, Yuki Kugiyama, Shinjiro Uchida, Akira Saeki, Satoru Hashimoto, Shinya<br />

Nagaoka, Atsumasa Komori, Hiroshi Yatsuhashi<br />

1601<br />

Identification of a non-invasive model based on serum<br />

levels of miR-122 and miR-222 to predict severe fibrosis<br />

and cirrhosis in patients with chronic hepatitis B<br />

Kevin Appourchaux 5,2 , Emilie Estrabaud 5,2 , Matthieu Resche-Rigon<br />

3,4 , Martine Lapalus 5,2 , Michelle Martinot-Peignoux 5,2 , Nathalie<br />

Boyer 2 , Michel Vidaud 6 , Pierre Bedossa 7,1 , Patrick Marcellin 5,2 ,<br />

Tarik Asselah 5,2 ; 1 INSERM, Paris, France; 2 Hepatology, Beaujon<br />

Hospital, Clichy, France; 3 Service de Biostatistique et information<br />

médicale, Saint-Louis Hospital, Paris, France; 4 INSERM UMR1153,<br />

Paris, France; 5 INSERM UMR1149, Paris, France; 6 INSERM<br />

UMR745, Paris, France; 7 Service d’Anatomie Pathologique, Beaujon<br />

Hospital, Clichy, France<br />

Background and aims Patients with chronic hepatitis B (CHB)<br />

are at high risk to develop cirrhosis and hepatocellular carcinoma<br />

(HCC). Staging fibrosis is mandatory, for both the prognosis<br />

and the need for treatment. The liver-enriched miR-122<br />

has been suggested to inhibit HBV replication. MicroRNAs are<br />

increasingly investigated as biomarkers because of their high<br />

stability. We aimed to identify miRNAs differentially expressed<br />

during fibrosis in patients with CHB. Patients and Methods A<br />

total of 103 patients with CHB were consecutively enrolled. All<br />

the patients had at least one biopsy to determine the stage of<br />

fibrosis (METAVIR scoring system) and no HCC. 85,4% of the<br />

patients were males, the mean age was 41.8 years. The mean<br />

HBV DNA level was 5.74 logUI/mL and 60.2% of the patients<br />

were negative for HBe antigen. Serums and biopsies were<br />

available for respectively 88 and 83 patients (69 paired liver/<br />

serum). Among patients with available serums 2.3%, 26.4%,<br />

27.6%, 24.1% and 19.6% had respectively F0, F1, F2, F3<br />

and F4 ; and among patients with available biopsies, 2.4%,<br />

27.7%, 27.7%, 20.5% and 21.7% had respectively F0, F1,<br />

F2, F3 and F4. The expression of miR-27b, -29a, -92a, -122,<br />

-146a, -222 and -224 which are related to liver fibrosis was<br />

assessed by RT-qPCR. Results A reduced expression of hepatic<br />

miR-122 (p=0.004) and miR-27b (p=0.038) was observed in<br />

patients with severe fibrosis and cirrhosis (F3-F4) compared<br />

to those with no, mild and moderate fibrosis (F0-F1-F2). An<br />

increased expression of hepatic miR-222 (p=0,018) and<br />

miR-224 (p=0.0001) was observed in patients with F3-F4<br />

compared to those with F0-F1-F2. A reduced expression of<br />

circulating miR-122 (p=0.049), miR-92a (p=0.031), and miR-<br />

29a (p=0.048) was observed in patients with F3-F4 compared<br />

to those with F0-F1-F2. An increased expression of circulating<br />

miR-146a (p=0.015) and miR-222 (p=0.040) was observed<br />

in patients with F3-F4 compared to those with F0-F1-F2. The<br />

multivariate analysis of clinical data and circulating miRNAs,<br />

allowed us to build a model combining circulating miR-122<br />

and miR-222, platelets count and alkaline phosphatase (ALP).<br />

The AUC of the model was 0,86 while FIB-4 and APRI had<br />

an AUC of 0,81 and 0,70, respectively. Conclusions Hepatic<br />

miR-27b, -122, -222 and -224 were differentially expressed in<br />

patients with F0-F1-F2 and in those with F3-F4. The increased<br />

expression of miR-222 in patients with F3-F4 might suggest<br />

a premalignant condition to HCC. The model combining the<br />

assessment of the following non-invasive biomarkers, circulating<br />

miR-122 and miR-222, platelets count and ALP was more<br />

accurate than FIB-4 and APRI to distinguish patients with F3-F4<br />

from those with F0-F1-F2.<br />

Disclosures:<br />

Nathalie Boyer - Board Membership: MSD, JANSSEN, Gilead, Abbvie; Speaking<br />

and Teaching: BMS<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

The following authors have nothing to disclose: Kevin Appourchaux, Emilie<br />

Estrabaud, Matthieu Resche-Rigon, Martine Lapalus, Michelle Martinot-Peignoux,<br />

Michel Vidaud, Pierre Bedossa<br />

1602<br />

Interleukin-2 Receptor and Transforming Growth Factor<br />

Alpha Independently Predict Significant Fibrosis in<br />

Chronic Hepatitis B Patients with Normal or Mildly Elevated<br />

Alanine Aminotransferase<br />

Gui-Qiang Wang, Yong-Qiong Deng, Hong Zhao; Department of<br />

infectious disease, Peking University First Hospital, Beijing, China<br />

Background: Patients with normal or mildly elevated serum alanine<br />

aminotransferase (ALT) are not guaranteed to be free from<br />

liver damage. Invasive liver biopsy is the current gold standard<br />

for assessing histological liver injury. There is an increasing<br />

demand for developing a noninvasive marker or index for diagnosis<br />

of liver inflammation and fibrosis in patients with normal<br />

or mildly elevated ALT.Aim: The aim of this study was to investigate<br />

the association of circulating cytokines and chemokines<br />

with liver inflammation and fibrosis, and to deeply assess their<br />

diagnostic value in CHB patients with ALT less than 2 times of<br />

upper limit of normal(ULN).Methods: Two hundred and twenty<br />

seven CHB patients with qualified biopsy were prospectively<br />

enrolled. There were 151 patients with ALT≥2×ULN and 76<br />

persons with ALT 2×ULN. All patients underwent liver biopsy.<br />

The serum levels of cytokines and chemokines were determined<br />

by simultaneous multianalyte detection methods. Histological<br />

Activity Index (HAI) and Liver fibrosis stage were assessed<br />

according to Ishak criteria.Results: Patients with moderate and<br />

more inflammation showed significantly higher levels of CXCL-<br />

11, CXCL-10 and interlerkin-2 receptor(IL-2 R)than the opposite<br />

(P0.001). Patients with significant fibrosis had higher levels of<br />

interleukin-8(IL-8) (P=0.027), Transforming growth factor alpha<br />

(TGF-a) (P=0.011), IL-2R (P=0.002) and CXCL-11(P=0.032)<br />

than no significant fibrosis. In 151 patients with ALT≤2×ULN,<br />

31.8% and 29.1% showed moderate and more inflammation<br />

and significant fibrosis respectively. Multivariate analysis indicated<br />

CXCL-11 independently associated with moderate and<br />

more inflammation, and TGF-a, L-2R with significant fibrosis<br />

in patients with ALT≤2×ULN. Based on CXCL-11, TGF-a, L-2R<br />

and clinical parameters, we developed inflammation-index<br />

and fib-index which showed areas under the receiver operating<br />

characteristics curve (AUROC) of 0.75 (95% CI 0.66-<br />

0.84) for moderate and more inflammation, and 0.82(95% CI<br />

0.75, 0.90) for significant fibrosis respectively in patients with<br />

ALT≤2×ULN. Leave-one out cross-validation showed 74.5%<br />

and 81.3% cases correctly classified by inflammation-index<br />

and fib-index. Compared to existing scores, fib-index was<br />

significantly superior to aspartate aminotransferase (AST) to<br />

platelet ratio index (APRI) and FIB-4 score for significant fibro-


992A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

sis.Conclusion: IL-2R and TGF-a were independent predictors<br />

for significant fibrosis in CHB patients with normal or mildly<br />

elevated ALT. An IL-2R and TGF-a based score, fib-index, was<br />

superior to APRI, FIB-4 for the diagnosis of significant fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gui-Qiang Wang, Yong-Qiong<br />

Deng, Hong Zhao<br />

1603<br />

Characterization of Nucleic acid testing reaction samples<br />

and Occult Hepatitis B infection Among blood<br />

donors in three cities of China<br />

Xiaomei Wang, Xiumei Chi, Ruihong Wu, Xiuzhu Gao, Junqi<br />

Niu; Hepatology, The First Hospital of Jilin University, Changchun,<br />

China<br />

Introduction. In China, HBV infection is common. With increasing<br />

voluntary blood donation and the still-prevalent infectious<br />

diseases in donors, we need to augment transfusion-transmitted<br />

infections testing before use. Our study was aimed to compare<br />

the seroprevalence of HBV, HCV and HIV among the<br />

donors of Chineses tested by ELISA and nucleic acid testing<br />

(NAT) . A variant of hepatitis B virus (HBV) having a specific<br />

mutation within the S gene has been found in occult HBV infection<br />

(OBI) . To know whether similar variants were involved in<br />

blood donors, we analyzed 18 blood samples HBV S region<br />

sequence. Materials and Methods. Enzyme linked immunosorbent<br />

assay (ELISA) was used for detection of HBsAg, HBeAg,<br />

anti-HBs, anti-HBe, anti-HBc,anti-HIV, and anti-HCV in all donor<br />

serum. The blood samples which were negative on ELISA were<br />

also subjected to NAT testing for HBV, HCV, and HIV. Nested<br />

PCR were used to amplify the S regions and products were<br />

sequenced and analyzed. Results. A total of 482370 donations<br />

were tested in studied, NAT yielded 156 HBV-DNA-positive<br />

donations in the HBsAg-negative (1:2365). NAT yielded 2<br />

HIV-RNA-positive donations were p24 Ag reactive. There were<br />

no NAT positive in anti-HCV negative donations. There were<br />

no escape mutations observed in the S gene. Conclusion. In<br />

regions with a high prevalence of HBV,there are likely to be<br />

a significant number of OBI donations that can be identified<br />

by NAT. Therefor, in order to provide safe blood, NAT should<br />

be applied in blood centers for testing HBV.In 18 occult HBV<br />

infections, HBV sequences lacked G145R and other escape<br />

mutations in S region, which may be caused by wild-type HBV<br />

strains.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xiaomei Wang, Xiumei Chi,<br />

Ruihong Wu, Xiuzhu Gao, Junqi Niu<br />

1604<br />

Baseline quantitative hepatitis B core antibodies can<br />

predict HBV DNA and HBsAg seroclearance in treatment-naïve<br />

HBeAg negative chronic hepatitis B patients<br />

Hui-Han Hu 1 , Jessica Liu 1 , Chia-Lin Chang 1 , Chin-Lan Jen 1 , Mei-<br />

Hsuan Lee 2 , Sheng-Nan Lu 3 , Li-Yu Wang 4 , San-Lin You 1 , Pei-Jer<br />

Chen 5,6 , Hwai-I Yang 1 , Chien-Jen Chen 1,7 ; 1 Genomics Research<br />

Center, Academia Sinica, Taipei, Taiwan; 2 Institute of Clinical<br />

Medicine, National Yang-Ming University, Taipei, Taiwan;<br />

3 Department of Gastroenterology, Chang-Gung Memorial Hospital,<br />

Kaohsiung, Taiwan; 4 MacKay College of Medicine, Taipei,<br />

Taiwan; 5 Division of Gastroenterology, Department of Internal<br />

Medicine, National Taiwan University Hospital, Taipei, Taiwan;<br />

6 National Taiwan University, Graduate Institute of Clinical Medicine,<br />

Taipei, Taiwan; 7 Graduate Institute of Epidemiology and<br />

Preventative Medicine, College of Public Health, National Taiwan<br />

University, Taipei, Taiwan<br />

Hepatitis B core antibody (anti-HBc) is one of the classical seromarkers<br />

for HBV infection. Several lines of evidence showed<br />

that among chronic hepatitis B (CHB) patients, quantitative<br />

anti-HBc levels were strongly correlated with serum ALT levels,<br />

suggesting that anti-HBc is a surrogate indicator of immune<br />

activation. In addition, among treated HBeAg-seropositive CHB<br />

patients, baseline quantitative anti-HBc levels were strong predictors<br />

for treatment responses including HBeAg seroconversion.<br />

Here, we investigated the impact of baseline quantitative<br />

anti-HBc levels on spontaneous HBV DNA and HBsAg seroclearance<br />

in treatment-naïve HBeAg negative CHB patients<br />

from the REVEAL-HBV cohort (n=2503). Baseline anti-HBc levels<br />

were assessed using a newly developed double-sandwich<br />

anti-HBc immunoassay. HBV DNA and HBsAg levels were longitudinally<br />

evaluated with repeated measurement in the longterm<br />

follow-up samples. The associations of baseline anti-HBc<br />

levels with HBV DNA and HBsAg seroclearance were assessed<br />

by multivariate-adjusted Cox proportional hazard regression<br />

models. The rate ratios (RR adj<br />

) with 95% confidence intervals<br />

(CI) were estimated. Baseline HBV DNA levels were significantly<br />

and positively associated with increasing anti-HBc levels.<br />

Mean log 10<br />

(anti-HBc) levels were 2.82, 3.64, 3.95, 4.14,<br />

and 4.29, respectively, for patients with HBV DNA levels of<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 993A<br />

The following authors have nothing to disclose: Hui-Han Hu, Jessica Liu, Chia-Lin<br />

Chang, Chin-Lan Jen, Mei-Hsuan Lee, Sheng-Nan Lu, Li-Yu Wang, San-Lin You,<br />

Hwai-I Yang, Chien-Jen Chen<br />

1605<br />

Risk assessment of hepatocellular carcinoma development<br />

using transient elastography vs. liver biopsy in<br />

chronic hepatitis B patients starting antiviral therapy<br />

Mi Na Kim 1 , Yeon Seok Seo 2 , Seung Up Kim 3 , Sang Gyune Kim 4 ,<br />

Soon Ho Um 2 , Kwang-Hyub Han 3 , Young Seok Kim 4 ; 1 Department<br />

of Internal Medicine, CHA Gangnam Medical Center, CHA<br />

University, Seoul, Korea (the Republic of); 2 Department of Internal<br />

Medicine, Korea University College of Medicine, Seoul, Korea (the<br />

Republic of); 3 Department of Internal Medicine, Yonsei University<br />

College of Medicine, Seoul, Korea (the Republic of); 4 Department<br />

of Internal Medicine, Soonchunhyang University Bucheon Hospital,<br />

Bucheon, Korea (the Republic of)<br />

Background/Aims: Liver stiffness (LS) assessed using transient<br />

elastography (TE) can assess the risk of developing hepatocellular<br />

carcinoma (HCC) in patients with chronic hepatitis B (CHB).<br />

We evaluated whether TE, when compared with histological<br />

data as a reference standard, can predict the risk of HCC<br />

development in CHB patients starting antiviral therapy. Methods:<br />

Database of 381 patients with CHB who underwent liver<br />

biopsy (LB) and TE starting antiviral therapy were reviewed. All<br />

patients underwent surveillance for HCC development using<br />

ultrasonography and alpha-fetoprotein. Results: During the<br />

median follow-up period of 48.1 (range 6.0-109.8) months,<br />

HCC developed in 34 (8.9%) patients. In patients with HCC<br />

development, age, proportion of diabetes mellitus, histological<br />

fibrosis stage, and LS value were significantly higher than those<br />

in patients without (all P


994A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

age-platelet index (API) for significant fibrosis (Metavir F2-4)<br />

in low-replicative (HBV DNA 4, 94.8 vs. 93.8%). At ROC-defined thresholds, APRI<br />

(≥0.33), AAR (≥0.93), FIB-4 (≥0.70) and API (>2) showed<br />

greater AUROCs for F2-4 diagnosis in low replicative (0.80,<br />

0.62, 0.81 and 0.71, respectively) vs. high-replicative patients<br />

(0.73, 0.52, 0.67 and 0.69, respectively). Conclusion: All 4<br />

biomarkers in both, low and high-replicative HBV demonstrate<br />

modest accuracy for fibrosis diagnosis at conventional cut-offs.<br />

Lowering the cut-offs may increase the diagnostic relevance of<br />

these biomarkers, particularly for APRI and FIB-4 in low-replicative<br />

disease.<br />

Disclosures:<br />

Faisal M. Sanai - Advisory Committees or Review Panels: Merck Sharpe Dohme,<br />

Bristol Myers Squibb, Janssen Pharmaceuticals, Gilead Sciences; Grant/<br />

Research Support: Roche Pharmaceuticals, Bristol Myers Squibb; Speaking and<br />

Teaching: Roche Pharmaceuticals, Janssen Pharmaceuticals, Gilead Sciences,<br />

Bayer Schering<br />

Abdulrahman A. Aljumah - Advisory Committees or Review Panels: Gilead;<br />

Speaking and Teaching: Bristol-Myers Squibb<br />

Ibrahim H. Altraif - Consulting: MSD, ABBVIE, BRISTOL MYERS; Grant/Research<br />

Support: ROCHE, JANSSEN; Speaking and Teaching: GSK<br />

The following authors have nothing to disclose: Taha Farah, Khalid Albeladi,<br />

Faisal Batwa, Yaser Dahlan, Mohammed A. Babatin, Hamad I. Al-ashgar,<br />

Hadeel Al-mana, Khaled Alsaad, Khalid A. Alswat, Khalid I. Bzeizi<br />

1608<br />

Barriers to obtaining appropriate Hepatitis B care<br />

among US Veterans<br />

Tatyana Kushner, David E. Kaplan, Marina Serper; Division of<br />

Gastroenterology and Hepatology, Hospital of the University of<br />

Pennsylvania, Philadelphia, PA<br />

Background: Studies have demonstrated poor follow-up and<br />

adherence to guidelines in chronic hepatitis B (HBV) care. We<br />

sought to identify patient and health system-specific risk factors<br />

for gaps in quality of HBV care within the VA. Methods:<br />

We performed a retrospective analysis of 125 US Veterans<br />

with chronic HBV within 4 VA sites in Pennsylvania and Delaware<br />

from 1999-2012. Patients with acute HBV were excluded<br />

(n=12). Demographic and clinical data were obtained from<br />

medical records. Process of care outcomes were: specialist consultation,<br />

appropriate lab testing, and liver imaging. Results:<br />

The patients were 98% male, 50% African American, with a<br />

mean age of 51 (SD=13). A total of 18% were disabled and<br />

72% had ≥1 psychiatric comorbidity (depression, anxiety, or<br />

PTSD), 52% had ≥1 psychiatric hospitalization within 2 years<br />

of HBSAg+ result. A total of 8% reported marijuana use, 10%<br />

IV drug use, and 39% alcohol use; 28% of new cases were<br />

diagnosed as inpatients. A total of 7% had cirrhosis and 6%<br />

had hepatic decompensation. The median ALT was 38 (IQR<br />

25, 68). A total of 85% of subjects had primary care follow-up,<br />

with a mean 4 primary care visits per year; 16% were seen<br />

by infectious disease (ID) and 41% by gastroenterology (GI)<br />

specialists. The median time to specialist consult was 39 days<br />

(IQR: 4 to 255), 71% of patients did not have a consult ordered<br />

within 12 months. The median time to HBV follow up testing<br />

was 21 months (IQR 10 to 47); 33% of patients had complete<br />

laboratory testing with HBV DNA, HBeAg, and HBeAb<br />

testing. The median time to liver imaging was 4 months (IQR<br />

0.5 months to 14); 36% had confirmatory imaging within an<br />

ultrasound within 12 months. In bivariate analyses, psychiatric<br />

hospitalization was associated lower likelihood of specialist<br />

consultation (OR 0.75, 95% CI 0.56-0.98, p=0.03). African<br />

American race was associated with less confirmatory testing<br />

(OR=0.25 95% CI, 0.09-0.77, p=0.02). Receiving care at an<br />

outpatient vs. hospital VA facility was associated with lower<br />

rates of liver imaging (OR 3.5, 95% CI 1.2-10.2, p=0.02) and<br />

associated with lower receipt of either specialist consultation,<br />

appropriate lab testing, or liver imaging (OR 0.29, 95% CI<br />

0.09-0.94, p=0.04). Conclusion: Veterans with chronic HBV<br />

are significantly affected by psychiatric comorbidity and substance<br />

abuse. Delays in timely specialist referral, appropriate<br />

lab testing, and imaging were noted among certain subgroups<br />

(psychiatric comorbidity, African American race, outpatient VA<br />

facility). Prospective <strong>studies</strong> should confirm these relationships<br />

so that interventions to educate providers and improve care<br />

among high risk groups can be appropriately targeted.<br />

Disclosures:<br />

David E. Kaplan - Grant/Research Support: Bayer Pharmaceuticals, Inovio Pharmaceuticals<br />

The following authors have nothing to disclose: Tatyana Kushner, Marina Serper<br />

1609<br />

The Usefulness of CLIF-SOFA Score for Diagnosis and<br />

Evaluating the Patients with HBV-related Acute-on-<br />

Chronic Liver Failure<br />

Zhihong Wan, Chen Li, Shaoli You, HongLing Liu, ShaoJie Xin,<br />

Bing Zhu; Liver Failure Treatment and Research Center, 302 hospital<br />

of PLA, Beijing, China<br />

Background and Aims: Acute-on-chronic liver failure (ACLF)<br />

is a unique clinical entity. The definitions of ACLF used differed<br />

between Eastern and Western countries. Recently, a<br />

large prospective observational study (CANONIC study) was<br />

performed by EASL Chronic Liver Failure (EASL-CLIF) consortium.<br />

However, most of the enrolled patients were alcoholic<br />

cirrhosis, the criteria are necessary to be validated in China,<br />

in which the majority of ACLF is caused by chronic hepatitis B<br />

infection. In this regard, our present study was aimed to investigate<br />

clinical characteristics and prognosis in patients of HBV-<br />

ACLF diagnosed by EASL-CLIF criteria. Methods: Consecutive<br />

patients with HBV-ACLF were enrolled. All of the patients met<br />

the Chinese diagnostic criteria. EASL-CLIF criteria were used to<br />

further identify the patients with HBV-ACLF (HBV-ACLF-EASL).<br />

CLIF-SOFA score was used to assess organ failure and define<br />

ACLF grades. The mortality of the patients was recorded at<br />

28 days. Results: Total of 167 patients with HBV-ACLF was<br />

enrolled. Among them, 139 (83.2%) had liver cirrhosis and<br />

133 (79.6%) had acute decompensation of cirrhosis (AD). The<br />

28-day mortality rate was 24.6% in these patients. Among<br />

167 patients, thirty-six (21.6%) met the EASL-CLIF diagnostic<br />

criteria (diagnosed as HBV-ACLF-EASL). The 28-day mortality<br />

rate of these patients was 61.1%, which was significant higher<br />

than those with HBV-ACLF (24.6%, P < 0.01), indicating that<br />

patients with HBV-ACLF-EASL had more severity of disease.<br />

Of patients with HBV-ACLF-EASL, nine (25%) had ACLF grade<br />

1, 24 (66.7%) had ACLF grade 2, and 3 (8.3%) had ACLF<br />

grade 3. The most frequent organ failures were liver (97.2%)


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 995A<br />

and coagulation failures (58.3%) followed by cerebral (13.9%)<br />

and kidney failures (11.1%). The mortality was 44.4%, 62.5%,<br />

and 100% in patients with ACLF grade 1, ACLF grade 2 and<br />

ACLF grade 3, respectively, which was significant higher in<br />

comparison with patients in CANONIC study in corresponding<br />

grade (22.1% in grade 1, 32% in grade 2 and 76.7%<br />

in grade 3, respectively). The results indicated that the EASL-<br />

CLIF criteria may be too strict for diagnosis HBV-related ACLF.<br />

Of note, there were 21.4% of HBV-ACLF patients underlying<br />

chronic HBV hepatitis had more than one organ failure (assessing<br />

by CLIF-SOFA), and the mortality was 66.7% among them.<br />

The results suggested that diagnosis of HBV-ACLF should not<br />

exclude those patients without cirrhosis. Conclusions: ESAL-CLIF<br />

diagnostic criteria may not be suitable for assessing the patients<br />

with HBV-related ACLF in our study. A large, prospective, and<br />

multicenter study is needed to further validate the criteria.<br />

Disclosures:<br />

The following authors have nothing to disclose: Zhihong Wan, Chen Li, Shaoli<br />

You, HongLing Liu, ShaoJie Xin, Bing Zhu<br />

1610<br />

Anti-platelet therapy reduces the incidence of hepatocellular<br />

carcinoma in patients with chronic hepatitis B<br />

Minjong Lee 2 , Jeong-Hoon Lee 2 , Sohee Oh 1 , YOUNG CHANG 2 ,<br />

Joon Yeul Nam 2 , Young Youn Cho 2 , Jeong-Ju Yoo 2 , Yuri Cho 2 ,<br />

Donghyeon Lee 2 , Eun Ju Cho 2 , Su Jong Yu 2 , Yoon Jun Kim 2 , Jung-<br />

Hwan Yoon 2 ; 1 Department of Biostatistics, Seoul Metropolitan<br />

Government Seoul National University Boramae Medical Center,<br />

Seoul, Korea (the Republic of); 2 Department of Internal Medicine<br />

and Liver Research Institute, Seoul National University College of<br />

Medicine, Seoul, Korea (the Republic of)<br />

Background: Platelets contribute to liver damage by promoting<br />

the intrahepatic accumulation of virus-specific CD8 T cells in<br />

mouse models of viral hepatitis. A recent study showed that<br />

the long-term use of the anti-platelet drugs in a mouse model<br />

of chronic hepatitis B can prevent hepatocarcinogenesis. We<br />

aimed to compare the incidence of hepatocellular carcinoma<br />

in patients not treated with anti-platelet drugs to those treated<br />

with anti-platelet drugs. Methods: We performed a retrospective<br />

analysis of data from 3,479 consecutive patients with<br />

chronic hepatitis B who had HBV DNA completely suppressed<br />

with antiviral treatment. The population was divided into two<br />

groups: Group 1(n=2,891) not treated with anti-platelet drugs;<br />

Group 2 (n=588) treated with aspirin and/or clopidgrel. Data<br />

were collected from patients analyzed by a multivariable Cox<br />

proportional hazards model for the entire cohort. Results:<br />

During the study period, hepatocellular carcinoma was developed<br />

in 229 patients (7.9%) in Group 1 and 15 (2.5%) in<br />

group 2. In multivariable analyses, Group 2 patients showed<br />

a significantly lower risk of HCC than that in Group 1 (hazard<br />

ratio [HR]=0.21, 95% confidence interval[CI]=0.11–0.39,<br />

P5 ×ULN) were excluded<br />

due to potential overestimation of fibrosis by TE. Results: Serum<br />

M30 levels were significantly increased in significant fibrotic<br />

patients compared with no/minor fibrosis or NC (259.9 [IQR:<br />

122.6-501.2] vs 70.6 [IQR: 41.3 - 105.3] U/L, P


996A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: ZhuJun Cao, Hui Wang, Xiaogang<br />

Xiang, FengDI Li, KeHui Liu, Weiliang Tang, Yuhan Liu, Lanyi Lin, Qing<br />

Guo, Qing Xie<br />

1612<br />

Statin and the risk of hepatocellular carcinoma and<br />

death in a hospital-based hepatitis B-infected population:<br />

a propensity-score landmark analysis<br />

John C Hsiang, Grace LH Wong, Vincent W. Wong, Henry Lik-<br />

Yuen Chan; Chinese University of Hong Kong, Hong Kong, Hong<br />

Kong<br />

BACKGROUND & AIMS The use of statin in hepatocellular carcinoma<br />

(HCC) and death prevention is still uncertain among<br />

hepatitis B-infected (HBV) patients. This study aimed to examine<br />

the effect of statin on HCC and death in a hospital-based HBV<br />

population METHODS We conducted a population study of<br />

HBV patients using the Hospital Authority registry database<br />

containing data of patients attending 43 public hospitals in<br />

Hong Kong. We defined statin use by landmark analysis<br />

to abrogate “immortal time bias” and propensity score (PS)<br />

weighting to minimise baseline confounders and “indication<br />

bias”. Multiple imputation analysis were performed for missing<br />

laboratory data. The weighted Cox regression analyses<br />

was performed for the risk of HCC (adjusting for competing<br />

mortality) and death. RESULTS A total of 73,499 patients, with<br />

a crude HCC incidence of 1.75 per 100 patient-years, were<br />

entered into the 2-year landmark analysis. After landmark analysis<br />

and PS weighting of baseline covariates, statin users had<br />

32% risk reduction in HCC (weighted sub-hazard ratio (SHR)<br />

0.68; 95% CI 0.48-0.97, p=0.033). In a PS weighted cohort<br />

of statin users and non-users, there was no difference in mortality<br />

compared statin users to non-users (weighted HR 0.92;<br />

0.76-1.11, p=0.386). In subgroup analysis, concurrent statin<br />

and nucleos(t)ide analogue (NA) use was associated with<br />

59% risk reduction in HCC (weighted SHR 0.41; 0.19-0.89,<br />

p=0.023) compared to NA use alone. CONCLUSION In this<br />

HBV cohort adjusted for confounders and biases, statin use<br />

is associated with reduced HCC risk by 32%. Additive HCC<br />

chemopreventive effect was seen with the concomitant use of<br />

NA and statin. Further prospective <strong>studies</strong> are warranted to<br />

investigate the potential use of statin in HBV-infected NA users.<br />

Weighted cumulative incidence of hepatocellular carcinoma<br />

(2-year landmark analysis, for a single multiple imputation dataset);<br />

p=0.033<br />

Disclosures:<br />

Grace LH Wong - Advisory Committees or Review Panels: Otsuka, Gilead;<br />

Speaking and Teaching: Echosens, Furui, Gilead, Janssen, Bristol-Myers Squibb,<br />

Otsuka, Abbvie<br />

Vincent W. Wong - Advisory Committees or Review Panels: AbbVie, Gilead,<br />

Janssen; Consulting: Merck, NovaMedica; Speaking and Teaching: Gilead,<br />

Echosens<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

The following authors have nothing to disclose: John C Hsiang<br />

1613<br />

Influence of the ethnic status in chronic hepatitis B<br />

patients: comparing the Netherlands, Belgium and Turkey<br />

Özgür M. Koc 1 , Sarah Konsten 1 , Geraldine Jansen 1 , Geert<br />

Robaeys 2,3 , Beytullah Yildirim 4 , Dirk Posthouwer 5 , Ger H. Koek 6 ;<br />

1 Faculty of Health, Medicine and Life Sciences, University Maastricht,<br />

Maastricht, Netherlands; 2 Department of Gastroenterology<br />

and Hepatology, East Limburg Hospital, Genk, Belgium; 3 Faculty<br />

of Medicine and Life Sciences, Limburg Clinical Research<br />

Program, Hasselt, Belgium; 4 Department of Gastroenterology,<br />

Ondokuz Mayis University, School of Medicine, Samsun, Turkey;<br />

5 Department of Internal Medicine, Infectious Diseases and Medical<br />

Microbiology, Maastricht University Medical Centre, Maastricht,<br />

Netherlands; 6 Department of Internal Medicine, Division of Gastroenterology<br />

and Hepatology, Maastricht University Medical Centre,<br />

Maastricht, Netherlands<br />

Background & Aims Hepatitis B virus (HBV) is a global threat<br />

affecting different ethnic groups. In some diseases, ethnicity<br />

is an essential predictor of disease outcome. Therefore, this<br />

study aimed to understand the influence of ethnic status on the<br />

natural history of chronic hepatitis B (CHB) infections in the<br />

Netherlands, Belgium and Turkey. Methods In this multicentre<br />

retrospective cohort study, 269 CHB patients from the Hepatology<br />

Outpatient Departments of three hospitals one in the<br />

Netherlands, Belgium and Turkey were included. Variables as<br />

baseline characteristics, route of transmission, laboratory characteristics<br />

and HBV endpoints were collected. Ethnicity was<br />

defined as country of birth. Chi-square and ANOVA tests were<br />

used for categorical and continuous variables, respectively.<br />

Results The CHB population in the Netherlands, Belgium and<br />

Turkey consisted of respectively 98, 68 and 103 patients. In<br />

the Netherlands 32.7 and 16.3% were respectively of Dutch<br />

and Chinese origin. In Belgium mainly Belgian (36.8%) and<br />

Turkish (33.8%) patients were included, whereas the population<br />

in Turkey consisted in 99% of patients of Turkish origin.<br />

The CHB population in the Netherlands, Belgium and Turkey<br />

differed significantly in the number of vertical (40.8, 17.6 and<br />

15.5% respectively, p= .000), sexual (16.3, 13.2 and 3.9%<br />

respectively, p= .013) and intra-familial transmission (1.0,<br />

11.8 and 26.2% respectively, p= .000). In addition, the countries<br />

showed significant differences in HBV endpoints such as<br />

the number of cirrhosis (11.2, 17.6 and 38.8% respectively,<br />

p= .000) and hepatocellular carcinoma (2.0, 2.9 and 11.7%<br />

respectively, p= .008). Subsequently, patients from Belgium<br />

and the Netherlands were categorized as of Dutch / Belgian<br />

(n=46), Turkish (n=27) and Chinese (n=18) origin. There was<br />

a significant difference in the number of vertical transmission<br />

(12.1, 22.2 and 83.3% respectively, p= .000), sexual transmission<br />

(27.6, 7.4 and 0.0% respectively, p= .007) and<br />

cirrhosis (8.6, 22.2 and 0.0% respectively, p= .045). Conclusions<br />

The Netherlands, Belgium and Turkey differed regarding<br />

ethnicity, routes of transmission and disease progression. Chinese<br />

patients were characterized by vertical transmission and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 997A<br />

absence of cirrhosis. In addition, Turkish and Dutch/Belgian<br />

patients had significantly more cirrhosis and acquired the disease<br />

by intra-familial and sexual transmission, respectively.This<br />

study therefore implies the importance of ethnicity as there is an<br />

association between ethnic groups, routes of transmission and<br />

disease progression.<br />

Disclosures:<br />

Geert Robaeys - Advisory Committees or Review Panels: Gilead; Speaking and<br />

Teaching: Merck, Janssens<br />

The following authors have nothing to disclose: Özgür M. Koc, Sarah Konsten,<br />

Geraldine Jansen, Beytullah Yildirim, Dirk Posthouwer, Ger H. Koek<br />

1614<br />

High-throughput and ultrasensitive assay for the quantification<br />

of HBV precore and basal core promoter<br />

mutants and their clinical significance<br />

Motokazu Mukaide, Kazumoto Murata, Masaya Sugiyama,<br />

Masashi Mizokami; National Center for Global Health and Medicine,<br />

Ichikawa, Japan<br />

Background/Aim: Precore (PC) G1896A and basal core promoter<br />

(BCP) A1762T/G1764A mutations of HBV is associated<br />

with HBeAg seroconversion. However, in the previous<br />

<strong>studies</strong>, the percentage of PC and BCP mutant were examined<br />

by semi-quantitative assay, and lacked sufficient sensitivity,<br />

which make difficult for precise prediction for the response to<br />

therapies in each patient. Therefore, our aim of this study is<br />

to develop a novel quantitative assay for HBV, with special<br />

reference to PC and BCP, using new generation nanoliter-sized<br />

droplet technology paired with digital PCR (ddPCR), and to<br />

investigate its clinical significances. Methods: We collected<br />

serum samples from 95 HBV patients [mean: 54 years, sex<br />

male 69%, ALT 45U/L(range 14 to 333), DNA 3.6 logIU /<br />

mL, HBeAg positive 27.4%, HBV genotypes (B: 11, C: 84) .<br />

Serum samples were stored at -80°C until use. Among them,<br />

57 patients were under treatment with nucleos(t)ide analogs<br />

and no patients were treated with IFN. ddPCR assays were<br />

performed using the QX100 ddPCRsystem. ddPCR primers and<br />

probes were designed on the basis of the conserved nature of<br />

these sequences using the Hepatitis Virus Database. Results:<br />

The assay using standard plasmid linearly detected HBV from<br />

2.5 to 10 5 copies, and the limit of detection was 0.005% of<br />

mutant (MT) in the presence of wild-type (WT). The total viral<br />

load (MT+WT) quantified by ddPCR showed a linear correlation<br />

with HBV DNA levels as measured by COBAS TaqMan<br />

HBV Test (range 2.1 to 9.1 log 10<br />

IU/mL: PC: R 2 =0:86, P


998A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

factors independently predict immune reactive phase, i.e. alanine<br />

aminotransferase (< 80 U/L, 0; ≥ 80 U/L, 1), HBV DNA<br />

(≥ 8 log IU/mL, 0; 6-8 log IU/mL, 2; < 6 log IU/mL, 3) and<br />

HBsAg (> 4.25 log IU/mL, 0; < 4.25 log IU/mL, 3); and a prediction<br />

score ranging from 0 to 7 was established. Cutoff values<br />

of 4 best discriminated immune reactive phase with area<br />

under the receiver operating characteristic curves (AUROC) of<br />

0.830 (0.794-0.862) in training cohort. By applying the cutoff<br />

value of 4, the score differentiates immune reactive phase with<br />

AUROC of 0.812 (0.757-0.859) and positive predictive value<br />

of 89% in validation cohort, respectively. Conclusion: A simply<br />

non-invasive diagnostic score constructed from routine laboratory<br />

parameters is accurate in differentiating immune phases<br />

in HBeAg-seropositive persons. Future validation is warranted.<br />

Disclosures:<br />

The following authors have nothing to disclose: Qing-Lei Zeng, Yan-Ling Fu, Jun<br />

Lv, Zu-Jiang Yu<br />

1617<br />

Postpartum hepatic flare after telbivudine withdrawal in<br />

pregnant women with chronic hepatitis B<br />

Jinfeng Liu 2 , Jing Wang 1 , Dongfang Jin 3 , Yingli He 2 , Li Jin 1 , Taotao<br />

Yan 1 , Furong Cao 1 , Yuan Yang 2 , Shulin Zhang 2 , Yingren Zhao 2 ,<br />

Tianyan Chen 2 ; 1 Xi’an Jiaotong University, Xi’an, China; 2 The First<br />

Affiliated Hospital of Medical College, Xi’an Jiaotong University,<br />

Xi’an, Shaanxi Province, China, Xi’an, China; 3 Shaanxi Kangfu<br />

Hospital, Xi’an, China<br />

Background: The efficacy of telbivudine for breaking vertical<br />

transmission of HBV has been established. Data on the risk of<br />

postpartum flare after telbivudine withdrawal and optimal duration<br />

of antiviral therapy after delivery are limited. Methods:<br />

We enrolled 312 women who received telbivudine beginning<br />

at week 24 or 28 of gestation and then followed up to 52<br />

weeks postpartum. Virological and biochemical parameters<br />

were assessed. Results: Of the 312 women enrolled, 26.3%<br />

had elevated serum alanine aminotransferase (ALT) during<br />

pregnancy. Telbivudine administration resulted in ALT normalization<br />

in 80.5% before delivery. Compared with women<br />

having a normal ALT level throughout pregnancy, those with<br />

elevated ALT had a significantly higher rate of ALT flare after<br />

telbivudine withdrawal (27.8% vs. 7.0%; P = 0.032). Multivariate<br />

analysis indicated only ALT elevation during pregnancy<br />

correlated with postpartum flare after telbivudine withdrawal.<br />

Those women with elevated ALT during pregnancy, continued<br />

antiviral treatment to 52 weeks postpartum, had a significantly<br />

higher HBeAg seroconversion rate (P = 0.008) and a notable<br />

decrease in HBsAg titers (P = 0.046). Conclusion: It is safe for<br />

the majority of women to withdraw telbivudine after delivery;<br />

whereas its withdrawal is not prudent for women with elevated<br />

ALT during pregnancy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jinfeng Liu, Jing Wang, Dongfang<br />

Jin, Yingli He, Li Jin, Taotao Yan, Furong Cao, Yuan Yang, Shulin Zhang, Yingren<br />

Zhao, Tianyan Chen<br />

1618<br />

Performance Assessment of Common Anti-hepatitis B<br />

Core Antigen Assays in Japan for Prevention of HBV<br />

Reactivation<br />

Takako Inoue 1 , Tomoya Iwase 1 , Satomi Kani 1 , Kumiko Ohne 1 ,<br />

Takaaki Goto 1 , Yukio Wakimoto 1 , Yasuhito Tanaka 2,1 ; 1 Clinical<br />

Laboratory, Nagoya City University Hospital, Nagoya, Japan;<br />

2 Department of Virology & Liver unit, Nagoya City University<br />

Graduate School of Medical Sciences, Nagoya, Japan<br />

Background: The measurement of anti-hepatitis B core antigen<br />

(anti-HBc) is a significant method to predict the risk of hepatitis<br />

B virus (HBV) reactivation in patients with immunosuppressive<br />

drug therapies. In Japan, several commercially assays for anti-<br />

HBc are available. The aim of this study is to compare the<br />

anti-HBc measurement methods and to evaluate those clinical<br />

utilities. Methods: Three kinds of anti-HBc assays were evaluated<br />

with 77 clinical samples which showed low anti-HBc titers<br />

around the cutoff values (from 0.1 to 2.0 C.O.I). The anti-HBc<br />

assays used in this study were two chemiluminescent enzyme<br />

immunoassay (HISCL anti-HBc [Sysmex] and Lumipulse anti-<br />

HBc N [Fujirebio, Inc.]) and one chemiluminescent immunoassay<br />

(Architect HBc II [Abbott Japan]). This study protocol was<br />

approved by the appropriate institutional ethics review committees.<br />

Results: In the direct comparison between HISCL anti-<br />

HBc and Architect HBc II, the concordance rate was 79.2%<br />

(61/77). Between HISCL anti-HBc and Lumipulse anti-HBc N,<br />

the concordance rate was 80.5% (62/77). Between Lumipulse<br />

anti-HBc N and Architect HBc II, the concordance rate was<br />

96.1% (74/77). In the 77 clinical samples, there were 17<br />

discrepant samples (TABLE). Of 9 discrepant samples negative<br />

only by HISCLE, 6 samples were positive for anti-HBs without<br />

the history of HB vaccination, indicating that they have been<br />

exposed to HBV. Of 8 discrepant samples positive by HISCLE,<br />

7 samples were positive by absorption test for HISCLE anti-<br />

HBc, implying that the results by Lumipulse and Architect were<br />

false negative. One of the discrepant samples positive only for<br />

HISCLE anti-HBc was HBsAg seropositive, and HBV-DNA was<br />

also detected (chronic hepatitis B carrier). Conclusions: In this<br />

study, the results of anti-HBc show discrepancies between each<br />

of the anti-HBc assays. In cases with near the cutoff values of<br />

anti-HBc, it would need to be retested with other assays or<br />

molecular methods. Future <strong>studies</strong> should standardize the anti-<br />

HBc titer measurement system to prevent HBV reactivation.<br />

Discrepancies (17 samples)<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Takako Inoue, Tomoya Iwase,<br />

Satomi Kani, Kumiko Ohne, Takaaki Goto, Yukio Wakimoto<br />

1619<br />

Does Cesarean section and formula feeding reduce<br />

vertical transmission in chronic hepatitis B pregnant<br />

mothers?<br />

Rasham Mittal, Amandeep K. Sahota; Kaiser Permanente Los<br />

Angeles Medical Center, Los Angeles, CA<br />

Background Perinatal acquisition of hepatitis B virus (HBV) carries<br />

85% to 95% risk of chronic infection. While few <strong>studies</strong><br />

have supported cesarean section (C-section) and bottle-feed-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 999A<br />

ing, others have challenged it. Limited data is available from<br />

United States. Study Aim: Effect of C-section and formula feeding<br />

on vertical transmission in a cohort of CHB pregnant mothers<br />

from United States. Methods: Retrospective study (January<br />

2005- June 2010) at a large integrated health care system in<br />

Southern California. CHB pregnant mothers identified using<br />

regional perinatal database. Inclusion criteria: age > 18 years,<br />

positive serum HBsAg (hepatitis B surface antigen). Exclusion<br />

criteria: negative/ false positive HBsAg; termination of pregnancy<br />

or lost to follow up. Data Collection; demographics,<br />

first live-birth pregnancy reviewed, mode of delivery, breast<br />

feeding (defined as breastfeeding for ≥2 weeks after birth),<br />

maternal hepatitis B e antigen (HBeAg), and HBsAg testing of<br />

infants at 9-18 months of age. Lost to follow up if child not seen<br />

at 9-18 months of age. Fisher’s exact test used to determine<br />

statistical significance. Results: A total of 472 infants were born<br />

to 462 CHB mothers (mean maternal age 32.5 years, range<br />

18-45; 78.8% Asian-Pacific Islanders). 308 (66.7%) delivered<br />

vaginally and 154 (33.3%) through C-section. All C-sections<br />

were performed for obstetrical indications, with labor arrest<br />

being most common. 384 (83.1%) mothers breast-fed after<br />

delivery. Mode of feeding could not be established for 13<br />

(2.8%) mothers. Hepatitis Be antigen (HBeAg) status of mothers<br />

outlined in table 1. Only 1 infant (delivered vaginally and<br />

breast-fed) born to HBeAg positive mother (unknown antenatal<br />

HBV viral load) had reactive HBsAg. No statistically significant<br />

difference found for risk of vertical transmission between<br />

mothers who breastfed vs. formula feeding (1/230 vs. 0/35;<br />

p>.05); normal vaginal vs. cesarean-section delivery (1/185<br />

vs. 0/85; p>.05). Limitation of data: 104 (26.2%) infants did<br />

not have HBsAg tested at 9-18 months follow up. Missing antenatal<br />

HBeAg and HBV viral load data for mothers. Conclusion:<br />

C-section and formula feeding did not reduce vertical transmission<br />

risk in cohort of CHB pregnant mothers from United States.<br />

Table 1 Maternal practices: Infant feeding, delivery mode<br />

HBV-related chronic liver diseases needs to be further studied,<br />

especially regarding the role of hepcidin. Methods: In the<br />

present study, we investigated the association between serum<br />

hepcidin, iron metabolism and the clinical indictors in the<br />

patients with HBV-related chronic liver diseases in a case-control<br />

study(2013-2015). A total of 318 subjects were included<br />

in the study, including 78 cases with HBV-related chronic hepatitis,<br />

85 cases with HBV-related liver cirrhosis, 77 cases with<br />

HBV-related hepatocellular carcinomar(HCC), and 78 healthy<br />

individuals as controls. Results: Compared with the healthy controls,<br />

all groups of cases had significantly higher levels of serum<br />

ferritin, and much lower levels of serum hepcidin. Compared<br />

with HBV-related chronic hepatitis, the levels of HBV-DNA loading,<br />

serum iron, total iron binding force(TIBC) and serum transferrin<br />

were more lower in the patients with cirrhosis and HCC,<br />

and the serum hepcidin levels were obviously higher. The level<br />

of serum ferritin was increased with the severity of liver fibrosis<br />

and inflammation activity. The multiple linear regression analysis<br />

showed that serum hepcidin levels had obvious correlation<br />

with HBV-DNA loading levels (β=-0.277, t=-4.438, P


1000A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

.Only results obtained with at least ten successful acquisitions,<br />

success rate of at least 60% and IQR/LSM ratio less than 30%<br />

were considered reliable. Metavir liver fibrosis stages were<br />

assessed by the same pathologist experienced blinded to the<br />

results of LSM. Statistical analysis was made according to SAS.<br />

RESULTS: Statistical analysis was conducted in 109 patients,<br />

26 patients were excluded for unreliable LB or LSM. Sex ratio<br />

M/F was 1.65, mean age was 38.6 years. The mean biopsy<br />

length was 26.4mm. The distribution of liver fibrosis according<br />

to the METAVIR score were: F0 (n = 17) (15.5%),F1: n = 51<br />

(46.7% ) F2: n = 20 (18.3%), F3: n = 15 (13.7%) and F4: n<br />

= 6 (5.5%). The mean delay between LSM and LB was 6 days,<br />

the mean acquisition success rate was 95% and the median<br />

IQR / LSM ratio was 14.6%. The area under receiver –operating<br />

characteristic curves were 0.87 (95%confidence intervals<br />

0.80-0.94) for F≥2, 0.90 (0.82-0.98) for F≥ 3 and 0.93<br />

(0.83-1)for F4.Optimal LSM cut-off values were 7.1 for ≥F2,8.6<br />

for≥F3 and 12 for F4 by maximizing of sensitivity and specificity.In<br />

univariate analysis, LSM was significantly correlated with<br />

Metavir fibrosis stage,activity grade,gender, and biopsy length<br />

.In multivariate analysis, LSM was significantly correlated only<br />

with Metavir fibrosis stage,activity grade, and biopsy length.<br />

CONCLUSION: LSM is an accurate method for assessment liver<br />

fibrosis in patients with chronic hepatitis B e negative with high<br />

performance diagnostic for cirrhosis.<br />

Disclosures:<br />

Salah Sahraoui - Employment: bms, roche<br />

Nabil Debzi - Speaking and Teaching: roche, msd, bms, Janssen, Abbvie<br />

The following authors have nothing to disclose: Nawal Guessab, Nawel Afredj,<br />

Rafik Kerbouche, Ibtissem Ouled Cheikh, Sonia Ait Younes, Samir Gourari,<br />

Yazid Chikhi, Salima Cheraitia, Mohamed Merniz, François Montestruc, Saadi<br />

Berkane<br />

1622<br />

Evaluation of HBV infection status in spouses of<br />

HBsAg-positive individuals<br />

Ebru Dik, Selma Tosun, Seher A. Coskuner, Alpay Ari, Meltem<br />

Avci; Infectious Diseases, Bozyaka Education and Research Hospital,<br />

Izmir, Turkey<br />

Objective: The transmission of infection through sexual intercourse<br />

is one of the most common routes of transmission for<br />

hepatitis B virus. Therefore the spouses of HBsAg-positive<br />

individuals must be screened for HBsAg positivity. The aim<br />

of the present study was to evaluate HBV infection status in<br />

spouses of HBsAg-positive individuals. Method: Using face-toface<br />

interview technique, a questionnaire was administered to<br />

HBsAg-positive cases who were followed in Viral Hepatitis Outpatient<br />

Clinics at zmir Bozyaka Education and Research Hospital<br />

in order to evaluate HBV infection status in their spouses,<br />

and the results were recorded if tests were available. Results:<br />

The study included a total of 949 HBsAg-positive cases who<br />

were aged 17 to 76 years. The spouses of 409 (43%) cases<br />

were never tested for HBV infection. When the results of the<br />

tested individuals were evaluated, the rate of HBsAg positivity<br />

was 6% (n=32), rate of antiHBcIgG+AntiHBs positivity was<br />

18%(n:96), rate of antiHBs positivity was 12%, rate of antiHBcIgG<br />

positivity alone was 1%, and rate of negative results in all<br />

HBV parameters was 62%. Conclusion: The data suggest that<br />

not all spouses of the individuals were tested for HBV infection.<br />

It is a striking finding that HBV infection status was unknown<br />

in 43% of the spouses. The rate of HBsAg positivity was 6%<br />

and the rate of immunization was very low (12%). AntiHBcIg-<br />

G+AntiHBs positivity was 18%. These results suggest that HBV<br />

transmission to the spouses from HBsAg-positive individuals is<br />

common, but the awareness of the individuals was very low.<br />

Low rate of testing and high rate of HBsAg positivity among<br />

the spouses suggest that screening of spouses for HBV infection<br />

must not be neglected.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ebru Dik, Selma Tosun, Seher A.<br />

Coskuner, Alpay Ari, Meltem Avci<br />

1623<br />

Is booster beneficial after universal HBV vaccination?<br />

Selma Tosun 1 , Erhun Kasirga 2 ; 1 Infectious Diseases, Bozyaka Education<br />

and Research Hospital, Izmir, Turkey; 2 Pediatric Gastroenterology,<br />

Celal Bayar University The Faculty of Medicine, Manisa,<br />

Turkey<br />

Objective: The infants are administered hepatitis B vaccination<br />

after birth at 0, 1, and 6 months. HBV vaccine-induced<br />

antibody titers are maintained in the blood for a long time<br />

(possibly lifelong) in immunocompetent individuals. However,<br />

there are some <strong>studies</strong> suggesting a considerable decrease<br />

in antiHBs titers below protective level 7 to 8 years after HBV<br />

vaccine administered immediately after birth. The <strong>studies</strong> also<br />

indicate an excellent antiHBs response with a booster dose<br />

of HBV vaccine. The aim of the present study was to evaluate<br />

antiHBs response in children with or without a booster vaccine<br />

dose after universal HBV vaccination. Method: The study<br />

included primary school children who received HBV vaccine<br />

after birth. The study subjects were divided into 2 groups and<br />

of the students in Group 1, 157 received three booster doses of<br />

HBV vaccine 7 to 9 years after national vaccination program.<br />

Totally 142 students in Group 2 did not receive any vaccine<br />

after the initial HBV vaccines administered after birth. AntiHBs<br />

titers were studied in blood samples, and the students with a<br />

antiHBs titer below 10 IU/mL were also tested for HBsAg and<br />

antiHBcIgG . Results: The study included a total of 299 student<br />

aged 7 to 12 years. The rate of positive antiHBs titers in students<br />

that received three booster doses of vaccine in Group 1<br />

was statistically significantly higher compared to the students<br />

in the second group (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1001A<br />

1624<br />

Assessment Of Treatment Response By Elastography<br />

During The Tenofovir Treatment In Nucleos(t)ide-Naïve<br />

Chronic Hepatitis B Patients<br />

Taeheon Lee, Byung Ik Kim, Yong Kyun Cho, Woo Kyu Jeon, Hong<br />

Joo Kim; Kangbuk Samsung Hospital, Seoul, Korea (the Republic<br />

of)<br />

Aim: Several longitudinal <strong>studies</strong> have demonstrated that elastrograpy<br />

could be a useful tool for monitoring of liver fibrosis<br />

during the treatment of chronic hepatitis B. However, not<br />

enough <strong>studies</strong> assessing regression of liver fibrosis by elastography<br />

are available to date. Therefore, the purpose of this study<br />

was to evaluate the liver stiffness before and during the treatment<br />

and assess the usefulness of this non-invasive modality.<br />

Methods: Fourty-eight treatment-naïve patients started treatment<br />

with tenofovir. Liver stiffness measurement by a shear wave<br />

elastography and a conventional ultrasonography was performed<br />

at baseline, after median 6 and 12 months. Cut-off levels<br />

to Metavir classification for fibrosis stage F0-1, F2, F3 and<br />

F4 were ≤5.6 kPa, ≤7.44 kPa, ≤8.71 kPa, and ≥10.87 kPa,<br />

respectively. Result: Median liver stiffness decreased from 6.68<br />

kPa(SD, ±3.90), 6.27(SD, ±3.88) and 5.69 kPa(SD, ±5.01) in<br />

patients treated with tenofovir at baseline, after median 6 and<br />

12 months. There were no significant difference in the degree<br />

of regression of liver stiffness after median 6 (P =0.29) and 12<br />

months (P =0.14). Twenty-two(45.9%) subjects were classified<br />

as F0-1, 3 (6.3%) as F2, 20 (41.7%) as F3, and 3 (6.3%) as<br />

F4. After median 6 and 12 months later, liver stiffness value of<br />

patients recorded as F0-1 was decreased mean 0.60 kPa(SD,<br />

±2.38) and 0.05 kPa(SD, ±2.17), F2 as 0.81 kPa(SD, ±1.69)<br />

and 0.99 kPa(SD, ±3.01), F3 as 0.81 kPa(SD, ±3.75) and<br />

1.02 kPa(SD, ±5.67)a and F4 as 13.23 kPa(SD, ±7.59) and<br />

4.50 kPa (P for trend =0.01 and 0.04 after 6months and<br />

12 months) Conclusion: Liver stiffness was not significantly<br />

improved during the tenofovir treatment in patients with chronic<br />

hepatitis B during 6 and 12 months. But these result showed a<br />

trend that in a patient with advanced fibrosis on baseline stiffness<br />

stage was more improvedcompared to that of mild fibrosis<br />

during 6 and 12 months. Long term prospective <strong>studies</strong> are<br />

required to investigate the usefulness of elastography for relationship<br />

between elastography stiffness and treatment response<br />

in chronic hepatitis B patients treated with tenofovir.<br />

Disclosures:<br />

The following authors have nothing to disclose: Taeheon Lee, Byung Ik Kim, Yong<br />

Kyun Cho, Woo Kyu Jeon, Hong Joo Kim<br />

1625<br />

Inadequate Treatment of Chronic Hepatitis B in Pregnant<br />

Women at High Risk for Vertical Transmission<br />

Diana Y. Wu 1,2 , muhammad I. Shafi 3,2 , Mohammed M. Aboelsoud<br />

3,2 , Colin Woodard 4,2 ; 1 Gastroenterology, Women and<br />

Infants Hospital, Providence, RI; 2 Brown University, Alpert Medical<br />

School, Providence, RI; 3 Internal Medicine, Memorial Hospital,<br />

Pawtucket, RI; 4 Internal Medicine, Kent Hospital, Warwick, RI<br />

Aim: To describe the clinical presentation and management of<br />

a cohort of pregnant women with hepatitis B in a single center<br />

specializing in obstetrics care. Methods: Pregnant women<br />

with positive hepatitis B surface antigen who were seen at<br />

Women and Infants Hospital, Providence, Rhode Island from<br />

2009-2014 were included. Medical records were reviewed for<br />

demographic and clinical information from initial presentation<br />

for positive hepatitis B surface antigen until time of delivery.<br />

Results: 112 women were included, median age 28 yrs. Of the<br />

76 women who had their birth countries recorded in their medical<br />

record, only seven (9%) were born in the United States.<br />

Median follow up was 4.8 months. 19 (17%) women had ALT<br />

levels more than twice the upper limit of normal for women (19<br />

U/L) during their pregnancy. 20 (18%) women were hepatitis<br />

B e-antigen positive. 15 women (13%) had HBV DNA levels<br />

> 200,000 IU/ml in the third trimester of pregnancy, which<br />

qualified them for initiation of treatment for chronic hepatitis B<br />

to prevent vertical transmission to their fetus. Five of these pregnant<br />

women were initiated on Lamivudine, three on Tenofovir,<br />

and one on Adefovir in their third trimesters. However, five of<br />

the eligible 15 (33%) women were not initiated on treatment<br />

during pregnancy to prevent vertical transmission, and a sixth<br />

woman declined therapy. Two of the 112 women in the study<br />

were already on therapy for chronic hepatitis B before pregnancy,<br />

one on Lamivudine and one on Adefovir, which were<br />

continued throughout pregnancy. One woman was on Entecavir<br />

before pregnancy and was not switched to a different oral<br />

agent during pregnancy. Another woman on Entecavir before<br />

pregnancy was switched to Telbivudine during pregnancy.<br />

Conclusions: Although there is universal screening of pregnant<br />

women in prenatal care in the US for chronic hepatitis B, there<br />

remains inadequate initiation of treatment in the third trimester<br />

for patients with high viral load, who are at high risk of vertical<br />

transmission. Furthermore, this study shows that there should<br />

be improved knowledge of safe oral antiviral therapies for<br />

hepatitis B in pregnancy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Diana Y. Wu, muhammad I.<br />

Shafi, Mohammed M. Aboelsoud, Colin Woodard<br />

1626<br />

NK cell function is restored with sequential NUC therapy<br />

and correlates with treatment response in Chronic Hepatitis<br />

B<br />

Upkar S. Gill 1 , Dimitra Peppa 2 , Lorenzo Micco 2 , Harsimran D.<br />

Singh 2 , Ivana Carey 3 , Graham R. Foster 1 , Mala K. Maini 2 , Patrick<br />

T. Kennedy 1 ; 1 Hepatology Unit, Centre for Immunobiology,<br />

Blizard Institute, Barts and The London, School of Medicine & Dentistry,<br />

QMUL, London, United Kingdom; 2 Division of Infection &<br />

Immunity, UCL, London, United Kingdom; 3 Institute of Liver Studies,<br />

Kings College London, London, United Kingdom<br />

INTRODUCTION: New treatment strategies to achieve HBsAg<br />

loss in Chronic Hepatitis B (CHB) are required. Sequential/<br />

combined therapeutic approaches comprising Pegylated-Interferon<br />

(Peg-IFNα) and nucleot(s)ide analogues (NUC) are being<br />

given greater consideration. We previously demonstrated<br />

boosting of NK cells in eAg- patients treated with Peg-IFNα<br />

(Micco et al, J. Hepatol, 2013). Here we studied differential<br />

NK cell responses in patients receiving sequential NUC therapy,<br />

after Peg-IFNα exposure, to elucidate a possible treatment<br />

advantage over current therapy strategies. PATIENTS &<br />

METHODS: PBMC from 43 eAg+ patients were analysed. 21<br />

patients underwent a 48-week course of Peg-IFNα and were<br />

studied longitudinally; 12/21 progressed to sequential NUC<br />

therapy and were followed until viral suppression. 12 patients<br />

receiving de novo NUC therapy, without Peg-IFNα exposure<br />

were compared for analysis. In addition 10 patients exposed<br />

to Peg-IFNα, and not progressing to sequential NUC therapy,<br />

were also studied for comparison. Phenotypic and functional<br />

analysis of NK cell subsets was performed by multicolour<br />

flow-cytometry and findings correlated with on-therapy HBsAg<br />

changes. RESULTS: There was cumulative expansion of activated<br />

and functional CD56 bright NK cells throughout 48-weeks<br />

of Peg-IFNα (p=0.0001), maintained at higher than pre-treatment<br />

levels for at least 9-months after switching to sequential


1002A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

NUCs. Proliferating NK cells with antiviral capacity increased<br />

significantly, and HBsAg significantly decreased, when NUC<br />

treatment was preceded by Peg-IFNα compared to de-novo<br />

NUCs. Interestingly, the frequency of antiviral NK cells, in<br />

those patients exposed to Peg-IFNα, without further therapy,<br />

reduced to baseline, 9-12 months post treatment cessation.<br />

In addition, increased IFNγ production and cytotoxicity, but<br />

reduced TRAIL expression, were noted in those patients (9/12)<br />

with HBsAg decline on sequential NUCs. CONSLUSIONS:<br />

The potent and cumulative expansion of activated NK cells<br />

with antiviral potential induced by Peg-IFNα is maintained on<br />

sequential NUCs. Patients receiving sequential NUCs achieved<br />

a greater decline in HBsAg than those treated with de novo<br />

NUCs, particularly those with TRAIL low IFNγ/CD107 hi NK cells.<br />

TRAIL expression in this setting may be pathogenic, in line with<br />

our previous data showing that TRAIL+ NK cells can delete<br />

antiviral T cells (Peppa et al, JEM 2013), suggesting the TRAIL<br />

pathway as a potential future therapeutic target. These findings<br />

provide mechanistic insights to support further consideration of<br />

sequential Peg-IFNα/NUC therapy to enhance HBsAg reduction,<br />

and achieve sustained treatment responses.<br />

Disclosures:<br />

Ivana Carey - Grant/Research Support: Gilead, Roche; Speaking and Teaching:<br />

BMS<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

Patrick T. Kennedy - Grant/Research Support: Roche, Gilead; Speaking and<br />

Teaching: BMS, Roche, Gilead<br />

The following authors have nothing to disclose: Upkar S. Gill, Dimitra Peppa,<br />

Lorenzo Micco, Harsimran D. Singh, Mala K. Maini<br />

1627<br />

Analysis of the effect on HBV life cycle by HBV genome<br />

editing using two different CRISPR/Cas9 systems<br />

Hiromi Abe, Keiichi Masaki, Tetsushi Sakuma, Masataka Tsuge,<br />

Nobuhiko Hiraga, Michio Imamura, C. Nelson Hayes, Hiroshi<br />

Aikata, Takashi Yamamoto, Kazuaki Chayama; Hiroshima University,<br />

Hiroshima, Japan<br />

Background and aim: Interferon and nucleoside analogue therapy<br />

effectively controls HBV replication in chronically infected<br />

patients. However, it is still difficult to eradicate HBV completely<br />

because covalently closed circular DNA (cccDNA) stably<br />

remains in the nucleus of hepatocytes. Recently, a new<br />

genome editing system, Cas9-nickase, which has mutagenic<br />

Cas9 activity, has been developed to target specific regions<br />

of double stranded DNA sequences with less off target effects<br />

compared with the original CRISPR/cas9. Using this approach<br />

we developed an anti-HBV system that targets three different<br />

portion of the HBV genome using only one plasmid. The aim<br />

of this study is to investigate the efficacy of these two CRISPR/<br />

Cas9 systems. Method: Two CRISPR/cas9 plasmids, nuclease<br />

and nickase types, were constructed to target three regions<br />

of the HBV genome: HBs, polymerase, and core. The original<br />

CRISPR-nuclease digests the HBV genome at these three<br />

regions using three target guide RNAs whereas the nickase<br />

type enzyme nicked both strands of the target site DNA using<br />

three pairs of guide RNAs transcribed from one plasmid.<br />

HepG2 cells grown in 6 well-plates in DMEM containing 10%<br />

FBS were co-transfected with 1.4×HBV genome and one of<br />

the two CRISPR-Cas9 encoding plasmids in a 1:2 ratio. Three<br />

days after co-transfection, cells and culture medium were harvested<br />

to evaluate the efficacy of the genome editing. Core<br />

associated HBV DNA was measured by qPCR after immunoprecipitation<br />

with anti-HBc antibody. The CRISPR/Cas9 plasmid<br />

was introduced into Baculovirus infection system. 14 days after<br />

infection, we measured HBsAg, eAg, and HBV DNA in the<br />

supernatant and intracellular HBV DNA and core-associated<br />

HBV DNA. Results: More than 78-95% of double stranded<br />

DNA of HBs, core and polymerase cleavage by both nuclease<br />

and nickase activity of CRISPR/Cas9 were confirmed by reporter-based<br />

assay. When we co-transfected 1.4×HBV genome<br />

and each CRISPR/Cas9-encoding plasmid, both HBsAg and<br />

eAg levels declined significantly in comparison with control<br />

plasmids in both types of CRISPR/Cas9 systems. Nuclease type<br />

Cas9 more efficiently reduced both HBsAg and eAg levels than<br />

nuclease type to less than 10% of control. Cas9-nickase also<br />

reduced HBsAg and eAg to approximately 10% of control cells<br />

(P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1003A<br />

HBV core-related antigen level. In the human hepatocytes,<br />

HDI decreased HBs antigen level by 70%, HBV DNA level<br />

by 90% and 3.5kb HBV RNA level by 50% in the cell. HDI<br />

increased HBs antigen level in the HepG2 cells transfected<br />

with the entire HBV genome and also increased 3.5 kb HBV<br />

RNA level by 3~4-folds. HDI increased HBV DNA associated<br />

with acetylated histone H4 as well as acetylation of histone H4<br />

in the HepG2.2.15 cells. Furthermore, histone acetyltransferase<br />

inhibitor garcinol partially inhibited the increase in 3.5kb<br />

HBV RNA level by HDI indicating the possible role of histone<br />

acetylation on the regulation of HBV replication. Conclusion:<br />

HDI increases HBV replication but decreases HBs antigen production<br />

in hepatoma cell line with HBV DNA integration. In<br />

non-tumor hepatocytes infected with HBV, HDI inhibits HBV<br />

production as well as HBV replication. Epigenetic mechanism<br />

may be involved in the regulation of HBV replication at multisteps<br />

in human hepatocytes.<br />

Disclosures:<br />

Tatsuo Kanda - Grant/Research Support: MSD K.K.; Speaking and Teaching:<br />

AJINOMOTO PHARMACEUTICALS CO., LTD, CHUGAI PHARMACEUTICAL<br />

CO., LTD, GlaxoSmithKlein, BMS, Jansen, Daiichisankyo; Stock Shareholder:<br />

Mitsubishi Tanabe Pharma<br />

Osamu Yokosuka - Grant/Research Support: Chugai, Taiho, Bristol Myers,<br />

Takeda<br />

The following authors have nothing to disclose: Shingo Nakamoto, Shuang Wu,<br />

Yuki Haga, Reina Sasaki, Masato Nakamura, Hiroshi Shirasawa<br />

1629<br />

CRISPR/Cas9 “double”-nickase mediated inactivation of<br />

hepatitis B virus replication<br />

Madina Karimova 2 , Niklas Beschorner 2 , Werner Dammermann 1 ,<br />

Jan Chemnitz 2 , Daniela S. Indenbirken 2 , Adam Grundhoff 2,3 , Stefan<br />

Lueth 1 , Frank Buchholz 4 , Schulze zur Wiesch Julian 1,3 , Joachim<br />

Hauber 2,3 ; 1 1. Department of Medicine, University Medical Center<br />

Hamburg-Eppendorf, Hamburg, Germany; 2 Heinrich Pette Institute<br />

– Leibniz Institute for Experimental Virology, Hamburg, Germany;<br />

3 German Center for Infection Research (DZIF), partner site Hamburg,<br />

Hamburg, Germany; 4 Department of Medical Systems Biology,<br />

University Hospital and Medical Faculty Carl Gustav Carus,<br />

TU Dresden, Dresden, Germany<br />

Background: Current antiviral therapies cannot cure hepatitis B<br />

virus (HBV) infection, since successful HBV eradication would<br />

require the inactivation of the viral genome, which primarily<br />

persists in host cells as episomal covalently closed circular<br />

DNA (cccDNA) and, to a lesser extent, as chromosomally integrated<br />

sequences. However, novel designer-enzymes, such as<br />

the CRISPR/Cas9 RNA-guided nuclease system, provide the<br />

technology for the development of advanced therapy strategies<br />

that directly attack the HBV genome. Methods: We report here<br />

the identification of cross-genotype conserved HBV sequences<br />

in the HbS and HbX region of the HBV genome that are specifically<br />

and effectively inactivated by a Cas9 double-nickase<br />

approach. Pairs of appropriately spaced Cas9 nickase<br />

mutants introduced two single-strand breaks on the opposite<br />

DNA strands that were subject to NHEJ repair in order to avoid<br />

off-target mutations by improving specificity by up to 1,500-fold<br />

relative to the wild-type Cas9 enzyme. Results: We show that<br />

this approach equally inactivated episomal cccDNA as well as<br />

chromosomally integrated HBV target sites in reporter cell lines<br />

as well as in chronically infected hepatoma cell lines. Analysis<br />

of Cas9n activity on ORF S and X target sites by next generation<br />

sequencing revealed efficient editing of cccDNA molecules<br />

targeted by either gRNA in HBV-infected HepG2.2.15,<br />

HepG2-H1.3 and HepG2-NTCP cells. The efficiency of X-specific<br />

sgRNAs was particularly high, with approximately 90%<br />

of all amplicons reads showing clearly discernible indel signatures.<br />

Conclusion: Our data support the feasibility of using<br />

the CRISPR/Cas9 nickase system for novel therapy strategies<br />

aiming to provide a cure for HBV infection.<br />

(A) Depiction of the HBV genome with location of amplicons<br />

spanning the S- and X-specific sgRNA target regions. The two<br />

amplicons are shown as dark gray boxes labeled S- and X-amplicon,<br />

respectively. Arrows shown at the top indicate the genomic<br />

location of gRNA target sequences.<br />

Disclosures:<br />

Stefan Lueth - Advisory Committees or Review Panels: BMS, Gilead, MSD, Janssen<br />

The following authors have nothing to disclose: Madina Karimova, Niklas Beschorner,<br />

Werner Dammermann, Jan Chemnitz, Daniela S. Indenbirken, Adam<br />

Grundhoff, Frank Buchholz, Schulze zur Wiesch Julian, Joachim Hauber<br />

1630<br />

GORASP2 regulates HCV and HBV through RNA and/or<br />

DNA replication and virion trafficking<br />

Dachuan Cai 1,2 , Jian Hong 1 , Xiao Liu 1 , Sae Hwan Lee 1 , Dahlene<br />

Fusco 1 , Esperance A. Schaefer 1 , Cynthia Brisac 1 , Anna Lidofsky 1 ,<br />

Wenyu Lin 1 , Raymond T. Chung 1 ; 1 Gastrointestinal Unit, Massachusetts<br />

General Hospital, Harvard Medical School, Boston, MA;<br />

2 Department for infectious diseases, The second affiliated hospital<br />

of Chongqing medical university, Chongqing, China<br />

Background/Aims: Hepatitis B virus (HBV) and HCV constitute<br />

major causes of blood transmitted hepatitis. Chronic HCV<br />

and HBV infection induce liver cirrhosis and hepatocellular<br />

carcinoma (HCC). However, the mechanisms by which these<br />

infections alter host immunity are not well understood. In a<br />

previous RNA-Seq analysis, we identified Golgi reassembly<br />

stacking protein 2 (GORASP2) as a host factor that participates<br />

in IFN-a regulation of HCV replication (Lin et al, J Hepatology<br />

2015, 62:1024). GORASP2 is involved in establishing the<br />

Golgi apparatus structure. We hypothesized that GORASP2<br />

protein regulates HCV and HBV through HCV RNA replication,<br />

HBV DNA and pregenomic RNA replication, and HCV/<br />

HBV virion trafficking. Methods: We performed siRNA knockdown<br />

or overexpression of GORASP2 in JFH1 HCV-infected<br />

Huh7.5.1 cells or HepG2.2.15 HBV replicon cells with or<br />

without IFN treatment. Selected gene mRNA variants and their<br />

proteins, together with HBV replication in cells and HBV virions<br />

in supernatant, were monitored by qRT-PCR and Western blot.<br />

HBV replication was assessed by measuring HBV X and S protein<br />

levels, HBV DNA levels and pregenomic RNA in cells, and<br />

supernatant virus DNA. Results: We found that GORASP2 has<br />

two isoforms based on RNA-Seq exon splicing bioinformatics<br />

analysis. GORASP2 V2 includes exon #2 while V1 does not.<br />

We found that siRNA knockdown of GORASP2 increased HCV<br />

RNA and protein levels in JFH1-infected Huh7.5.1 cells and<br />

HBV levels in DNA and pregenomic RNA in HepG2.2.15 cells.<br />

GORASP2 mRNA and protein expression were not affected by<br />

IFN treatment, indicating that GORASP2 expression is independent<br />

of direct IFN regulation. We found that GORASP2<br />

V1 overexpression reduced HCV RNA and protein levels in<br />

JFH1-infected cells and HBV DNA and pregenomic RNA levels<br />

in HepG2.2.15 cells. In contrast, overexpression of GORASP2<br />

V2 had no significant effect on HCV RNA or protein levels<br />

in JFH1 cells. Golgi phosphoprotein 3 (GOLPH3) has been


1004A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

reported to regulate HCV virion secretion through interaction<br />

with phosphatidylinositol 4-phosphate. We confirmed that<br />

GOLPH3 knockdown reduced HCV virion secretion in JFH1 cell<br />

culture supernatants and HBV DNA and pregenomic RNA levels<br />

in HepG2.2.15 cells. Moreover, we found that GORASP2<br />

knockdown increased both intracellular HCV RNA replication<br />

and virion trafficking and release to the supernatant. Conclusions:<br />

We conclude that GORASP2 protein regulates HCV and<br />

HBV replication through RNA replication and virion trafficking.<br />

Further exploration of the mechanism by which GORASP2 or<br />

GOLPH3 alters viral nucleic acid replication and virion trafficking<br />

are warranted.<br />

Disclosures:<br />

Dahlene Fusco - Grant/Research Support: Gilead<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Dachuan Cai, Jian Hong, Xiao<br />

Liu, Sae Hwan Lee, Esperance A. Schaefer, Cynthia Brisac, Anna Lidofsky,<br />

Wenyu Lin<br />

compared to those without sustained HBeAg seroconversion<br />

were shown in table. Multivariate analysis found pre-treatment<br />

HBVDNA < 7 logIU/ml (OR=5.7, 95%CI=1.1-28.6, p=0.03)<br />

and rs12794714 CYP2R1 TT genotype (OR=4.1, 95%CI=1.1-<br />

15.5, p=0.04) strongly predicted HBeAg seroconversion at 24<br />

weeks post treatment. HBVDNA before treatment and at week<br />

12 was significant lower in rs12794714 CYP2R1 TT genotype<br />

than non-TT genotype. There was no association between<br />

the remaining 12 SNPs and treatment response, including sustained<br />

HBeAg seroconversion, combined response (HBeAg<br />

loss and HBVDNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1005A<br />

of 1,653 Japanese samples representing 892 HBV patients<br />

and 761 healthy controls; we used the data from these 2,605<br />

individuals to assess the effects of HLA genotype on CHB infection.<br />

PATIENTS/METHODS: Of the 2,605 Japanese genomic<br />

DNA samples used in this study, 1,653 samples were newly<br />

obtained from healthy volunteers (n=761) or HBV patients<br />

(n=892 each patient was categorized as either inactive carrier<br />

(IC), chronic hepatitis (CH), or acute exacerbation of CHB, liver<br />

cirrhosis and hepatocellular carcinoma); the 892 patients were<br />

treated at 14 multi-center hospitals throughout Japan; the other<br />

952 patient samples were those used in our previous study. For<br />

each newly collected sample (n=1,653), HLA-DPB1 genotyping<br />

was performed by PCR-Luminex method. The associations of<br />

HLA-DPB1 alleles or genotypes with CHB infection or disease<br />

progression were assessed via a χ2 test. RESULTS: Significant<br />

associations between CHB infection and five DPB1 alleles (two<br />

susceptibility alleles, DPB1 * 05:01 and * 09:01, and three protective<br />

alleles, DPB1 * 02:01, * 04:01, and * 04:02) were confirmed<br />

in a population comprising 2,605 Japanese individuals.<br />

Odds ratios for CHB were higher for those with both DPB1<br />

susceptibility alleles than for those with only one susceptibility<br />

allele; therefore, effects of susceptibility alleles were additive<br />

for risk of CHB infection. Similarly, protective alleles showed an<br />

additive effect on protection from CHB infection. Moreover, heterozygotes<br />

of any protective allele showed stronger association<br />

with CHB than did homozygotes, suggesting that heterozygotes<br />

may bind a greater variety of hepatitis B-derived peptides,<br />

and thus present these peptides more efficiently to T cell receptors<br />

than homozygotes. Notably, compound-heterozygote of<br />

DPB1*04:02 (protective) and DPB1*05:01 (susceptibility) was<br />

significantly associated with protection against CHB infection,<br />

which indicated that one protective HLA-DPB1 molecule can<br />

provide dominant protection. CONCLUSIONS: Identification of<br />

the HLA-DPB1 genotypes associated with susceptibility to and<br />

protection from CHB infection is essential for future analysis of<br />

the mechanisms responsible for immune recognition of hepatitis<br />

B virus antigens by HLA-DPB1 molecules.<br />

Disclosures:<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

Namiki Izumi - Speaking and Teaching: MSD Co., Chugai Co., Daiichi Sankyo<br />

Co., Bayer Co., Bristol Meyers Co.<br />

Naoya Sakamoto - Grant/Research Support: Gilead Sciences, TRSS; Speaking<br />

and Teaching: Bristol Myers Squibb, Janssen Pharm, Chugai co ltd<br />

The following authors have nothing to disclose: Nao Nishida, Jun Ohashi, Masaya<br />

Sugiyama, Takayo Tsuchiura, Ken Yamamoto, Keisuke Hino, Masao Honda,<br />

Hiroshi Yatsuhashi, Kazuhiko Koike, Osamu Yokosuka, Eiji Tanaka, Akinobu<br />

Taketomi, Masayuki Kurosaki, Yuichiro Eguchi, Takehiko Sasazuki, Katsushi<br />

Tokunaga, Masashi Mizokami, Tatsuya Kanto<br />

influence the outcome of the infection. Furthermore, SNPs in<br />

HLA-C have been implicated in hepatitis B chronicity. However,<br />

whether there is a role for these variations in prediction of<br />

treatment outcome in chronic hepatitis B (CHB) patients is currently<br />

unknown. Methods: We included 86 CHB patients (41<br />

HBeAg-positive; 45 HBeAg-negative) from a previously conducted<br />

study who completed 48 weeks of peginterferon alfa-2a<br />

and adefovir combination therapy followed by a treatment-free<br />

follow-up (week 72). Patient DNA was isolated from PBMCs.<br />

Twelve SNPs in the HLA-C gene (±500 base pairs) were genotyped<br />

with the Illumina Human Omni1-Quad BeadChip. Genotyping<br />

of KIRs was performed by PCR. Combined response at<br />

week 72 (HBeAg negativity, HBV-DNA ≤2,000 IU/mL, and<br />

ALT normalization) was achieved in 14/41 HBeAg-positive<br />

and 17/45 HBeAg-negative patients. Results: One SNP in<br />

HLA-C (rs2308557; A allele) was significantly associated with<br />

combined response in HBeAg-positive patients after correction<br />

for multiple testing (p=0.003), but not in HBeAg-negative<br />

patients. This SNP is in high linkage with the SNP defining the<br />

presence of an Asn or Lys at position 80 of the α 1<br />

domain of<br />

HLA-C, important for binding specific KIRs (HLA-C group C1 or<br />

C2, respectively). The distribution of C1C1, C1C2, C2C2 genotypes<br />

in all patients was 36 (42%), 36 (42%), and 17 (20%),<br />

respectively. HBeAg-positive patients with the combination<br />

KIR2DL1-C2 had higher rates of combined response (13/24<br />

vs 1/17, p=0.001), and higher baseline ALT levels (median<br />

ALT 136 vs 50 U/L, p=0.002) compared to patients with<br />

only C1 alleles. In HBeAg negative patients, the combination<br />

KIR2DS2-C1 was more frequent in combined responders versus<br />

non-responders (11/17 versus 9/28 respectively, p=0.037) In<br />

multivariable analysis, presence of HLA-C2/KIR2DL1 predicted<br />

response independent of HBV genotype A (p=0.012) and ALT<br />

levels at baseline (p=0.005). Furthermore, the presence of HLA-<br />

C1/KIR2DS2 predicted response in HBeAg negative patients<br />

independent of HBV DNA (p=0.045) and HBsAg levels at<br />

baseline (p=0.040). Conclusion: KIR/HLA-C combination was<br />

strongly associated with combined response in both HBeAg<br />

positive and negative CHB patients treated with combination<br />

therapy. These findings support an important role for the interaction<br />

of NK cell receptors and their HLA-C ligands in determining<br />

host immune activity and response to interferon-based<br />

therapy in chronic hepatitis B.<br />

Disclosures:<br />

Bart Takkenberg - Speaking and Teaching: Roche, Gilead, BMS<br />

Hendrik W. Reesink - Consulting: Abbvie, Alnylam, BMS, Gilead, Janssen-Cilag,<br />

Merck, PRA-International, Regulus, Replicor, Roche, R-Pharm; Grant/Research<br />

Support: AbbVie, BMS, Boehringer Ingelheim, Gilead, Janssen-Cilag, Merck,<br />

PRA-International, Regulus, Replicor, Roche<br />

The following authors have nothing to disclose: Femke Stelma, Louis Jansen,<br />

Marjan J. Sinnige, Karel A. van Dort, Ester M. van Leeuwen, Neeltje A. Kootstra<br />

1633<br />

The Combination of KIR and HLA-C Is Associated with<br />

Response to Peginterferon and Adefovir Combination<br />

Therapy in Chronic Hepatitis B Patients.<br />

Femke Stelma 1,2 , Louis Jansen 1,2 , Marjan J. Sinnige 2 , Karel A. van<br />

Dort 2 , Bart Takkenberg 1 , Ester M. van Leeuwen 2 , Neeltje A. Kootstra<br />

2 , Hendrik W. Reesink 1 ; 1 Gastroenterology and Hepatology,<br />

Academic Medical Center, Amsterdam, Netherlands; 2 Experimental<br />

Immunology, Academic Medical Center, Amsterdam, Netherlands<br />

Introduction: Killer cell immunoglobulin-like receptors (KIR)<br />

expressed on NK cells interact with human leukocyte antigen<br />

class C (HLA-C) ligands. From acute hepatitis C <strong>studies</strong>, it is<br />

known that genetic variations in KIR/HLA-C combinations can<br />

1634<br />

NFATc3 exerts anti-viral and anti-tumor dual function<br />

in hepatitis B virus-related hepatocellular carcinoma by<br />

activating interferon pathway<br />

Xiangmei Chen, Jin Cheng, Zhenzhen Zeng, Fengmin Lu; Department<br />

of Microbiology & Infectious Disease Center, Peking Univercity<br />

Health Science Center, Beijing, China<br />

Purpose: This study aimed to explore the aberrant status of<br />

innate immune response-related genes, NFATc3 and ADAR1,<br />

and their potential function in HBV-related hepatocellular carcinoma(HCC).<br />

Methods: Real-time PCR was used to determine<br />

the mRNA level of NFATc3 and ADAR1 in HCC samples. The<br />

transcriptional regulation of IFN-related genes by NFATc3 or<br />

ADAR1 was detected by luciferase reporter assay and realtime<br />

PCR. The anti-HBV function of NFATc3 was evaluated


1006A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

in vitro andin vivo. CCK-8 assay, wound healing, and transwell<br />

assay were used to determine the effect of NFATc3<br />

on HCC oncogenic properties. Results: Our previous aCGH<br />

results showed that NFATc3 gene was frequently deleted in<br />

HCC samples(7/24), while ADAR1 gene was largely amplified(10/24).<br />

NFATc3 was significantly downregulated in HCC<br />

tissues compared to paired non-tumor tissues(P=0.002, N=75),<br />

and the decrease of NFATc3 in tumor tissues was significantly<br />

correlated with poor prognosis(P=0.021). ADAR1 was significantly<br />

upregulated in HCC tissues(P=0.007, N=66). NFATc3<br />

overexpression induced IFN-α, β and γ transcriptional activity<br />

and upregulated the mRNA level of IFN stimulated genes<br />

in HCC cells, while ADAR1 suppressed the transcriptional<br />

activity of IFN-β promoter. After mutating of NFAT binding<br />

site within the IFN-β gene promoter, NFATc3 failed to induce<br />

its transcriptional activity, demonstrating that NFATc3 exerted<br />

its function through directly binding with IFN gene promoter.<br />

In Hep3B cells transiently transfected with 1.2x HBV replicon<br />

or HepAD38 cells stably expressing HBV, NFATc3 overexpression<br />

significantly decreased HBsAg, HBeAg, HBV pregenomic<br />

RNA and HBV total DNA level, and adding IFN receptor<br />

blocker in the culture medium abolished this inhibitory effect.<br />

While in immune deficient HCC cell line Huh-7, ectopic expression<br />

of NFATc3 couldn’t activate IFN pathway or suppress HBV<br />

replication. The above observations demonstrated that NFATc3<br />

exerted its anti-HBV function by upregulating IFN pathway. In<br />

mouse model established by hydrodynamic tail vein injection of<br />

1.2x HBV replicon, co-injection of NFATc3 expression plasmid<br />

significantly decreased HBsAg and HBeAg level in peripheral<br />

blood of the mice. More importantly, for HCC patients with<br />

lower NFATc3 expression, the HBV DNA level in their peripheral<br />

blood was significantly higher than patients with higher<br />

NFATc3(P=0.040). Lastly, restoration of NFATc3 significantly<br />

reduced the proliferation and migration ability of HCC cells.<br />

Conclusion: Our results suggest that NFATc3 exerts anti-viral<br />

and anti-tumor dual function by activating IFN pathway, which<br />

provided a novel option of clinical treatment for HBV-related<br />

HCC by targeting NFATc3.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xiangmei Chen, Jin Cheng,<br />

Zhenzhen Zeng, Fengmin Lu<br />

1635<br />

Generation of human peripheral blood mononuclear<br />

cells-engrafted mice for the study of immunological<br />

response against hepatitis B virus<br />

Satoshi Aono 1 , Tomohide Tatsumi 1 , Seiichi Tawara 1 , Yoshiki Onishi<br />

1 , Akira Nishio 1 , Shigeki Suemura 1 , Hayato Hikita 1 , Ryotaro<br />

Sakamori 1 , Naoki Hiramatsu 1 , Hiroshi Suemizu 2 , Takeshi Takahashi<br />

2 , Tetsuo Takehara 1 ; 1 Osaka University, Suita, Japan; 2 Central<br />

Institute for Experimental Animals, Kawasaki, Japan<br />

Background&Aims: Immunologically humanized mouse model<br />

is required for the assessment of immunological responses<br />

against hepatitis B virus (HBV) . Human peripheral blood mononuclear<br />

cells (PBMC) –engrafted mouse model using immunodificient<br />

NOG mice resulted in severe graft-versus-host disease.<br />

Recently, we generated NOG-Iaβ/β2m double KO mice which<br />

were NOG mice deficient in both MHC class I and II (DKO-<br />

NOG mice) . In this study, we established a new humanized<br />

mouse model capable of evaluating immune response against<br />

HBV derived proteins. Methods and Results: We used NOG<br />

(NODShi.Cg-Prkdc scid Il12rg tm1sug ) mice and DKO-NOG mice.<br />

Human PBMC were inoculated into these mice via tail vein.<br />

After injection of human immune cells, both infiltration into the<br />

liver of human immune cells and hepatocyte damage were<br />

observed in NOG mice, but not in DKO-NOG mice. Serum<br />

ALT levels were severely elevated in NOG mice, but not in<br />

DKO-NOG mice (510±299 IU/l vs, 14.5±4.12 IU/L at day<br />

15, 939±444 IU/l vs, 36.4±19.7 IU/L at day 29, respectively)<br />

. Seven of 8 NOG mice died within 2 months after injection of<br />

human PBMC whereas all DKO-NOG mice survived more than<br />

70 days. Liver mononuclear cells (MNCs) isolated from these<br />

mice were subjected to flow cytometry. On day 28 after injection<br />

of human PBMC, replacement rates of human immune cells<br />

in the liver increased up to 85% in DKO-NOG mice. On day<br />

15, the expressions of PD-1 and Tim-3 on T cells from DKO-<br />

NOG mice were significantly lower than those from NOG<br />

mice. On day 15, the frequencies of B cells and dendritic cells<br />

(DCs) in DKO-NOG mice were significantly higher than those<br />

in NOG mice. Next, we evaluated the production of anti-HBs<br />

antibody (anti-HBs) in sera of human PBMC-engrafted DKO-<br />

NOG mice. Inoculations of recombinant hepatitis B vaccine<br />

resulted in the production of anti-HBs in 2 of four vaccinated<br />

mice and also in the increase of CD138, a plasma cell marker,<br />

positive B cells in splenocytes. Finally, We evaluated the induction<br />

of HBc-derived peptide-specific cytotoxic T lymphocytes<br />

(CTLs) by using specific tetramer in human PBMC-engrafted<br />

DKO-NOG mice. Vaccination of HBc-derived peptide-pulsed<br />

DCs resulted in significant increase of HBc-derived peptide-specific<br />

CTLs in vaccinated mice. Moreover, HBV expression in the<br />

liver by hydrodynamic injection resulted in significant increase<br />

of HBc-derived peptide-specific CTLs. Conclusion: The present<br />

study demonstrates that immunological responses against HBV<br />

could be induced in the human PBMC-engrafted DKO-NOG<br />

mice. This system would provide us with a new promising<br />

model evaluating immunological responses against HBV in<br />

human chimeric livers.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Satoshi Aono, Tomohide Tatsumi,<br />

Seiichi Tawara, Yoshiki Onishi, Akira Nishio, Shigeki Suemura, Ryotaro Sakamori,<br />

Naoki Hiramatsu, Hiroshi Suemizu, Takeshi Takahashi<br />

1636<br />

NK cells play important roles in the pathogenesis of<br />

HBV infection via alternation of NKp46 high NKG2A high<br />

subset<br />

Teppei Yoshioka, Yoshinobu Yokoyama, Akira Nishio, Yoshiki<br />

Onishi, Kaori Mukai, Satoshi Aono, Minoru Shigekawa, Hayato<br />

Hikita, Ryotaro Sakamori, Naoki Hiramatsu, Takuya Miyagi,<br />

Tomohide Tatsumi, Tetsuo Takehara; Gastroenterology and Hepatology,<br />

Osaka university graduate school of medicine, Suita, Japan<br />

Background/Aim: Natural Killer (NK) cells attract attention in<br />

the control of viral infection. However, the role of NK cells<br />

against HBV infection remains controversial. The aim of this<br />

study was to investigate the role of NK cells in the pathogenesis<br />

of chronic hepatitis B (CHB), focusing on an activating receptor<br />

NKp46 and an inhibitory receptor NKG2A. Method: Untreated<br />

CHB patients (n = 53) and healthy subjects (HS; n = 35) were<br />

enrolled. Peripheral blood lymphocytes were obtained and<br />

the frequencies of NKp46 and NKG2A were analyzed by<br />

flow cytometry. To evaluate the function in each NK subsets,<br />

we examined the cytotoxicity by RT-PCR and flow cytometry.<br />

NK cells were co-cultured with HepG2.2.15 (2215), which<br />

express HBV related protein, or HepG2 (G2), and the role of<br />

NKp46 was evaluated by neutralizing antibody of NKp46<br />

(9E2). Finally, alternation of the frequencies of NK cell subset


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1007A<br />

in patients treated with nucleos(t)ide analogue (NA; n = 17) or<br />

interferon-α (IFNα; n = 9) were analyzed by flow cytometry.<br />

Result: CHB patients were divided into inactive CHB group<br />

(I-CHB, n = 27; patients showed both serum ALT less than 40<br />

IU/ml and HBV-DNA less than 4 LC /ml) and active CHB group<br />

(A-CHB, n = 26; remaining patients). The expression of NKp46<br />

was higher in A-CHB than in HS or I-CHB. A unique subset<br />

which strongly expressed both NKp46 and NKG2A was identified<br />

(NKp46 high NKG2A high subset). The frequencies of this<br />

subset in NK cells were 4.8 ± 2.9%, 4.9 ± 3.6% and 10.3 ±<br />

6.3% in HS, I-CHB and A-CHB, respectively. The frequencies of<br />

this subset positively correlated with serum ALT level (P


1008A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

using Chi-square test and logistic regression analysis. Results:<br />

In the CHB group, 28.5% were inactive carriers, 67.7% were<br />

HBe-antigen negative, 17.2% had cirrhosis, 6.5% had hepatocellular<br />

carcinoma and 4.4% were liver transplanted. 85.6%<br />

had already received treatment for CHB. The mean age and<br />

sex distribution differed significantly between the CHB and the<br />

SC groups (45.38±15.4 vs. 61.35±14.5; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1009A<br />

HBV-specific CD8+ T cells was almost completely diminished in<br />

the absence of MHC class I expression on non-hematopoietic<br />

cells, resulting in more than 2,000 fold decrease in the number<br />

of HBV-specific CD8 T cells compared with their counterparts<br />

in MHC class I positive control animals. HBV-specific CD8+<br />

T cells did not express granzyme B, either, even though they<br />

clearly recognized antigen and expressed activation marker<br />

CD69. Conclusion: These results suggest that cross-presentation<br />

by hematopoietic cells is required for the optimum expansion<br />

of HBV-specific CD8 T cells, but endogenous antigen presentation<br />

by non-hematopoietic cells, presumably hepatocytes, is far<br />

more essential in the expansion and cytolytic differentiation of<br />

HBV-specific CD8+ T cells.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Masanori Isogawa, Yasuhiro<br />

Murata, Keigo Kawashima<br />

1641<br />

Initial sites of hepadnavirus-host genome integration<br />

after de novo infection of HepaRG cells and woodchuck<br />

livers<br />

Ranjit Chauhan, Norma D. Churchill, Thomas I. Michalak; Molecular<br />

Virology and Hepatology Research Group, Memorial University,<br />

St. John’s, NF, Canada<br />

Backgound: HBV is highly oncogenic virus and its integration<br />

into human genome precedes development of hepatocellular<br />

carcinoma (HCC). It is expected that the sites and the spread<br />

of HBV integrations across human genome might be relevant to<br />

the establishment and progression of HCC. The initial sites of<br />

HBV-host DNA junctions remain unknown. Aim: To identify the<br />

earliest sites of hepadnavirus integration into host’s genome by<br />

exploring de novo infection of human hepatocyte-like HepaRG<br />

cells with authentic HBV and woodchucks infected with woodchuck<br />

hepatitis virus (WHV). Methods: HepaRG cells infected<br />

with HBV were investigated in 15 min, 30 min, 1 h, 3 h, 24<br />

h, 72 h, 7 days (d), 10 d, 2 weeks (wk), 4 wk and 7 wk<br />

post-infection. Liver biopsies were obtained from woodchucks<br />

infected with WHV at 1 h or 3 h and at 6 wk post-infection (JVI<br />

2008;82:8579). HepaRG cells were analyzed for HBV X gene<br />

integrations, while liver biopsies for WHV X and preS sequence<br />

integrations applying inverse-PCR followed by amplicon cloning<br />

and sequencing. Results: Intracellular HBV DNA and its replication<br />

intermediates became detectable in HepaRG cells from<br />

1 h and remained present up to 7 wk post-exposure. HBV DNA<br />

integrations became identifiable in 1 h post-infection. Junctions<br />

of HBV X gene in retrotransposon sequences, such as LINE1<br />

(long-interspersed nuclear element 1) and LINE2, became<br />

prominent from 3 days post-infection and they represented<br />

approximately half the virus-host junctions detected at that time.<br />

Insertions of HBV X integrant into other genes, including human<br />

satellite HSAT-II sequence, neurotrimin on chromosome (Ch)<br />

11q25, ZNF-782 on Ch-9q22.3, myosin 3B on Ch-2q31.1,<br />

and ten-eleven translocation-1 gene (TET-1) on Ch-10q21.3<br />

were also detected. In woodchucks, integration of WHV preS<br />

genomic sequence into MAML2 gene on Ch-11q21 (based on<br />

88% homology with human) was identified in 1 h post-infection<br />

and WHV X gene integrants into unidentified so far woodchuck<br />

genes were evident in 1 h and 3 h post-infection. The<br />

HBV core promoter/enhancer-II region followed by enhancer I<br />

region and its WHV equivalent were predisposed to form the<br />

earliest junctions with host DNA. In addition, multiple virus-virus<br />

DNA joints were identified from 1 h post-infection onward in<br />

both HBV-infected cells and WHV-infected woodchuck livers.<br />

Conclusions: HBV DNA integrates very early into the host’s<br />

genome and may utilize retrotransposon elements and translocation<br />

genes from the early stages of infection for spread and<br />

induction of pro-oncogenic perturbations throughout the host’s<br />

genome. The predisposition to very early integration was also<br />

evident in natural WHV infection in woodchucks.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ranjit Chauhan, Norma D. Churchill,<br />

Thomas I. Michalak<br />

1642<br />

An impact of cellular core-fucosylation in hepatitis B<br />

virus infection into hepatoma cell lines<br />

Eiji Miyoshi 2 , Shinji Takamatsu 1 , Mayuka Shimomura 1 , Yoshihiro<br />

Kamada 1,3 , Haruka Maeda 1 , Tomoaki Subajima 2 , Hayato Hikita 3 ,<br />

Iijima Mayumi 4 , Yuta Okamato 5 , Ryo Misaki 5 , Kazuhito Fujiyama 5 ,<br />

Tetsuo Takehara 3 , Keiji Ueda 6 , Shunichi Kuroda 4 ; 1 Department of<br />

Molecular Biochemistry & Clinical Investigation, Osaka University<br />

Graduate School of Medicine, Suita, Japan; 2 Department of<br />

Molecular Biochemistry and Clinical Investigation, Osaka University<br />

Graduated School of Medicine, Suita, Japan; 3 Department of<br />

Gastroenterology and Hepatology, Osaka University Graduated<br />

School of Medicine, Suita, Japan; 4 Department of Bioengineering<br />

Sciences, Graduate School of Bioagricultural Sciences, Nagoya<br />

University, Nagoya, Japan; 5 Applied Microbiology Laboratory,<br />

International Center for Biotechnology, Osaka University, Suita,<br />

Japan; 6 Department of Microbiology, Osaka University Graduated<br />

School of Medicine, Suita, Japan<br />

Background & Aims: It is a well known fact that oligosaccharides<br />

on cell surface proteins such as growth factor receptors<br />

as well as virus/bacteria entry receptors regulate their functions<br />

dynamically. These oligosaccharide structures are regulated<br />

by a set of expression of several kinds of glycosyltransferase.<br />

In the present study, we investigated the involvement of glycosylation<br />

in hepatitis B virus (HBV) entry onto hepatoma cells.<br />

Methods: We used several kinds of hepatoma cell lines such<br />

as Huh7, Hep3B, Huh6 and HB611 cells. HB611 cells are<br />

established by transfection of tandem repeats of HBV genome<br />

DNA into Huh6 cells and secrete HBV-related antigens into<br />

the conditioned medium. Bio nanocapsules (BNC) are hollow<br />

particles consisting of HBV surface antigen (HBs) large proteins<br />

embedded in a unilamellar liposome, and used as pseudo-HBV<br />

virus. Incorporation of BNC in oligosaccharide remodeling<br />

hepatoma cells was evaluated by flow cytometer and con-focal<br />

microscopy. Oligosaccharide structure was analyzed with<br />

lectin blot, lectin microarray and mass spectrometry. To identify<br />

association molecules with sodium taurocholate cotransporting<br />

polypeptide (NTCP), which is one of major candidate for HBV<br />

receptors, immune-precipitation with NTCP antibody followed<br />

by LC-MS analyses was performed. Results: Transfection of<br />

glycosyltransferase genes into several kinds of hepatoma cells<br />

changed the level of BNC uptake. Among them, HB611 cells<br />

showed dramatic increases in BNC incorporation, compared<br />

to parental cell, Huh6 cells. As a result from glycomic analyses,<br />

increases in cellular fucosylation and sialylation were observed<br />

in HB611 cells. Knock down of fucosyltransferase 8 (FUT8)<br />

gene as well as sialidase treatment decreased BNC entry into<br />

HB611 cells. In contrast, overexpression of FUT8 in Huh6 cells<br />

increased BNC incorporation. Addition of core-fucose specific<br />

lectin, Pholiota squarrosa lectin (PhoSL) in the conditioned<br />

medium inhibited the entry of BNC into HB611 cells. While<br />

expression of sodium taurocholate cotransporting polypeptide<br />

(NTCP) was lower in HB611 cells than Huh6 cells, levels of<br />

co-precipitated molecules with NTCP were dependent on levels<br />

of core fucosylation, suggesting that core-fucosylation regu-


1010A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

lates BNC entry into hepatoma cells. As a result from confocal<br />

microscopy observation, endocytosis of NTCP seemed to be<br />

up-regulated in HB611 cells in terms of core fucosylation. Conclusions:<br />

Core-fucosylation and sialylation have pivotal roles in<br />

HBV infection into hepatoma cells. Especially, downregulation<br />

of core-fucosylation with Fut8 knockdown or treatment with<br />

core-fucose specific lectin might be a novel target for HBV<br />

therapy.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Eiji Miyoshi, Shinji Takamatsu,<br />

Mayuka Shimomura, Yoshihiro Kamada, Haruka Maeda, Tomoaki Subajima,<br />

Iijima Mayumi, Yuta Okamato, Ryo Misaki, Kazuhito Fujiyama, Keiji Ueda,<br />

Shunichi Kuroda<br />

1643<br />

Indoleamine-2, 3-dioxygenase as the intrinsic anti-HBV<br />

effector that leads to durable immune responses in<br />

patients with acute hepatitis B: A comparative analysis<br />

with the sequence in hepatic flare<br />

Sachiyo Yoshio 1 , Masaya Sugiyama 1 , Hirotaka Shoji 1 , Yohei<br />

Mano 1 , Yoshihiko Aoki 1 , Toru Okamoto 2 , Yoshiharu Matsuura 2 ,<br />

Masashi Mizokami 1 , Tatsuya Kanto 1 ; 1 The Research Center for<br />

hepatitis and immunology, National Center for global health and<br />

medicine, Chiba, Japan; 2 Molecular Virology, Research Institute<br />

for Microbial Diseases, Osaka University, Suita, Japan<br />

BACKGROUD&AIMS: Primary HBV infection to immune competent<br />

adults is mostly a self-limited disease, resulting in the HBV<br />

clearance. However, HBV is hard-to-eradicate once it attains<br />

persistent infection. The factors that distinguish these conditions<br />

should be crucial for the control of HBV. Indoleamine-2, 3-dioxygenase<br />

(IDO) is induced in hepatocytes by IFN-γ and catalyzes<br />

tryptophan to form kynurenine. IDO exerts dual function<br />

in infectious diseases; acting as a suppressor of intracellular<br />

pathogens and as an immune regulator. We aimed to explore<br />

the factors, including IDO, capable of suppressing HBV, by<br />

comprehensive immunological analysis between patients with<br />

acute hepatitis with HBV clearance and those at hepatic flare<br />

with persistent infection. METHODS: We enrolled HBV-positive<br />

patients with acute hepatitis (AH, N=26), chronic hepatitis<br />

(CH, 12), hepatic flare (HF, 14) and healthy volunteers (HV,<br />

10) in this study. We longitudinally compared serum levels of<br />

40 cytokines and chemokine by a multiplexed assay. Systemic<br />

IDO activity was determined by the ratio of kynurenine to tryptophan<br />

measured by HPLC. In order to examine the mechanism<br />

of IDO induction and its impact on HBV replication, we set up<br />

the co-culture consisting of HBV transfected Huh7 (HBV+Huh7),<br />

freshly-isolated human NK cells and plasmacytoid dendritic<br />

cells (pDCs). The IDO-knock out (IDO-KO) Huh7 was established<br />

by CRISPR-Cas9 system. RESULTS: The AH group showed<br />

a robust increase of IDO activity prior to the peak of ALT elevation,<br />

the levels of which were higher than those in the CH,<br />

HF or HV group. The IDO activity subsided overtime in accordance<br />

with the decline of ALT and HBVDNA. In the AH group,<br />

the levels of CXCL9, CXCL10 and CXCL11 were increased in<br />

the acute phase, followed by the sustained increase of IFN-γ,<br />

TNF-α, CCL19 and CCL22 until the disappearance of HBsAg.<br />

In contrast, such sequential transition of the immune factors,<br />

beginning from the robust IDO activation, was not observed<br />

in the HF group. In the co-culture, the coexistence of NK and<br />

pDCs synergistically increased the production of IFN-γ and<br />

IFN-α thereby enhancing IDO activity in HBV+Huh7, resulting<br />

in HBV suppression. In IDO-KO HBV+Huh7, the suppressive<br />

effect of NK and pDCs on HBV replication was abrogated<br />

(P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1011A<br />

MX1 N-terminal domain to specifically identify its effects on<br />

HBV and HCV replication is warranted.<br />

Disclosures:<br />

Dahlene Fusco - Grant/Research Support: Gilead<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

The following authors have nothing to disclose: Wenyu Lin, Dachuan Cai, Jian<br />

Hong, Xiao Liu, Hong Zhao, Chuanlong Zhu, Esperance A. Schaefer, Cynthia<br />

Brisac, Sae Hwan Lee, Anna Lidofsky<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Masataka Tsuge, Nobuhiko<br />

Hiraga, Hiromi Abe, Daiki Miki, Michio Imamura, Hidenori Ochi, C. Nelson<br />

Hayes<br />

1645<br />

Difference of intracellular responses by HBV genotype A<br />

and C infection in humanized mouse model<br />

Masataka Tsuge 1 , Nobuhiko Hiraga 1 , Hiromi Abe 1 , Daiki Miki 2 ,<br />

Michio Imamura 1 , Hidenori Ochi 1 , C. Nelson Hayes 1 , Kazuaki<br />

Chayama 1 ; 1 Department of Gastroenterology and Metabolism,<br />

Hiroshima university, Hiroshima, Japan; 2 Institute of Physical and<br />

Chemical Research (RIKEN), Hiroshima, Japan<br />

Background & Aims: It is well known that the characteristics of<br />

chronic hepatitis B, including the clinical course and responsiveness<br />

to anti-viral treatments, differ among hepatitis B virus<br />

(HBV) genotypes. However, the cause of these differences has<br />

not been clarified. We have previously established a chronic<br />

HBV infection model using human hepatocyte transplanted<br />

uPA-SCID mice lacking human immune cells. We previously<br />

reported the effects of HBV infection in human hepatocytes<br />

using cDNA microarray. In the present study, we analyzed<br />

mRNA expression profiles using next generation sequencing in<br />

this mouse model and compared the responses between HBV<br />

genotypes A and C. Methods: Fifteen chimeric mice were prepared<br />

and divided into the following three groups: uninfected<br />

control mice, HBV genotype A infection, and HBV genotype C<br />

infection. Ten mice were inoculated with HBV genotype A or C,<br />

and human hepatocytes in the mouse livers were collected after<br />

mouse serum HBV DNA titers had reached a plateau of around<br />

8 Log copies/ml. To analyze changes in gene expression in<br />

human hepatocytes, we performed next generation sequencing<br />

and compared gene expression profiles. Results: HBV genotype<br />

A and C infection resulted in significant changes in mRNA<br />

expression levels in HBV-infected hepatocytes in 780 and 208<br />

genes, respectively (P


1012A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

B, where the liver damage is T-cell mediated, our data suggest<br />

that a unique type of humoral immunity against HBcAg plays a<br />

primary role in the pathogenesis of HBV-ALF.<br />

Disclosures:<br />

The following authors have nothing to disclose: Zhaochun Chen, Huaying Zhao,<br />

Peter Schuck, Ronald E. Engle, Fausto Zamboni, Giacomo Diaz, David E. Kleiner,<br />

Robert H. Purcell, Patrizia Farci<br />

1647<br />

The expansion of CD11b - CD27 - NK-cell subset contributes<br />

to HBV persistence via TGF-β1<br />

Qiongfang Zhang 1,2 , Wenwei Yin 1,2 , Jian-ying Shao 1,2 , Qian<br />

Liu 1,2 , Peng Hu 1,2 , Huaidong Hu 1,2 , Hong Ren 1,2 , Dazhi Zhang 1,2 ;<br />

1 The institute of viral hepatitis,Department of infectiouse diseases,<br />

Chongqing Medical University, Chongqing, China; 2 The Second<br />

Affiliated Hospital of Chongqing Medical University, Chong qing,<br />

China<br />

Background /purposeThe mechanism underlying persistent<br />

hepatitis B virus (HBV) infection remains unclear. Natural killer<br />

(NK) cells, as a major component of the antiviral immune<br />

response, has been reported to alter their distribution of NK-cell<br />

subsets. We reported a novel mechanism associated with viral<br />

persistence from NK cell-mediated antiviral immunity in patients<br />

in immune tolerant (IT) phase. Method We investigated the<br />

role of NK subsets to persistent HBV infection in 109 chronic<br />

hepatitis B (CHB) patients and 38 healthy controls (HC) by flow<br />

cytometry. Serum cytokine concentrations were analyzed using<br />

a Cytometric Bead Array (CBA) Inflammation Kit and a standard<br />

sandwich enzyme-linked immunosorbent assay (ELISA)<br />

Kit. Result We found a substantial CD11b - CD27 - (DN) NK-cell<br />

population harboured in patients in IT phase compared with<br />

patients in immune active (IA) phase, patients in immune inactive<br />

(IN) phase and HC. And there existed a positive correlation<br />

between the expanding DN NK-cell subtype and HBV-DNA<br />

load. This DN NK subset displayed an immature and inactive<br />

phenotype and showed poor degranulation and IFN-γproduction.<br />

Higher concentrations of transforming growth factor beta1<br />

(TGF-β1) were found in sera from IT-phase patients, which positively<br />

correlated with HBV-DNA load. In vitro, TGF-β1 up-regulated<br />

the presence of DN NK-cell population. Moreoverthe,<br />

percentage of DN NK-cell subset in fresh NK cells preincubated<br />

with sera from IT patients was markedly higher compared with<br />

IA phase patients, IN phase patients and HC. Anti-TGF-β1 antibodies<br />

could partially restore the altered NK-cell subsets. Conclusion<br />

Our findings demonstrated that the IT microenvironment<br />

may render circulating NK cells into an inefficient subtype by<br />

up-regulating the levels of TGF-β1 in serum, which was correlated<br />

with persistent HBV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Qiongfang Zhang, Wenwei Yin,<br />

Jian-ying Shao, Qian Liu, Peng Hu, Huaidong Hu, Hong Ren, Dazhi Zhang<br />

1648<br />

Comprehensive analysis of the Rab family to screen<br />

membrane traffic pathways utilized by hepatitis B virus<br />

Jun Inoue, Yasuteru Kondo, Teruyuki Umetsu, Takuya Nakamura,<br />

Takayuki Kogure, Yu Nakagome, Yuta Wakui, Tatsuki Morosawa,<br />

Yasuyuki Fujisaka, Satoshi Takai, Tooru Shimosegawa; Division<br />

of Gastroenterology, Tohoku University Graduate School of Medicine,<br />

Sendai, Japan<br />

Background/aim: The trafficking pathways of hepatitis B virus<br />

(HBV) components in the host cells have remained largely<br />

unknown. Previously, we showed that the small GTPase Rab7<br />

regulates HBV secretion from multivesicular bodies (MVBs)<br />

by facilitating viral degradation by lysosomes (Inoue at al.,<br />

J Cell Sci 2015). It is thought that the HBV core particles are<br />

enveloped in MVBs, but the transport pathway of the envelope<br />

HBs proteins to MVBs has not been determined. In this study,<br />

we performed a comprehensive screening of Rab proteins to<br />

identify important pathways in the assembly/release of HBV.<br />

Methods: Using HepG2.2.15 cells stably expressing HBV, 61<br />

Rab proteins were depleted separately with siRNA-mediated<br />

knockdown. With an siRNA concentration that maintained cell<br />

viability of more than 80%, HBV DNA and HBsAg in the culture<br />

supernatant were quantified 3 days after siRNA transfection.<br />

Results: The effects of Rab knockdown were divided into 7<br />

groups as follows: 1) decreased HBV DNA, 2) increased HBV<br />

DNA, 3) decreased HBsAg, 4) increased HBsAg, 5) decreased<br />

both HBV DNA and HBsAg, 6) increased both HBV DNA and<br />

HBsAg, 7) increased HBV DNA but decreased HBsAg. Group<br />

5 Rab proteins, which were considered to play roles in the<br />

assembly/release of both HBsAg subviral particles (SVPs)<br />

and HBV particles included Rab8, Rab13, and Rab38. These<br />

were reported to be associated with the vesicle transport from<br />

Golgi. Interestingly, the depletion of Rab5B, which is one of<br />

three isotypes of Rab5 and is known to be associated with<br />

early endosomes, increased both HBV DNA and HBsAg. The<br />

Rab5B depletion increased HBV DNA significantly more than<br />

the depletion of other Rab5 isotypes and the effect was rescued<br />

by the expression of GFP-Rab5B. Sucrose density gradient centrifugation<br />

revealed that the density of increased HBV particles<br />

after the Rab5B depletion was lowered from 1.24 g/ml to<br />

1.17-1.19 g/ml. These data indicate that Rab5B depletion<br />

may increase the release of immature HBV particles whose<br />

envelope components are altered. The significance of Golgi-related<br />

Rab proteins and Rab5B in early endosomes for the<br />

formation of mature HBV particles suggests the presence of a<br />

novel HBV life cycle pathway: HBs proteins are transported to<br />

the cellular membrane once via Golgi, and are endocytosed to<br />

early endosomes and late endosomes/MVBs, where the HBV<br />

core particles are thought to be enveloped. Conclusion: The<br />

findings of the Rab family screening suggested the importance<br />

of both the Golgi-related pathway and endocytotic pathway in<br />

the formation of mature HBV particles. This is a novel concept<br />

in the HBV life cycle and these steps could become targets of<br />

antiviral therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jun Inoue, Yasuteru Kondo,<br />

Teruyuki Umetsu, Takuya Nakamura, Takayuki Kogure, Yu Nakagome, Yuta<br />

Wakui, Tatsuki Morosawa, Yasuyuki Fujisaka, Satoshi Takai, Tooru Shimosegawa<br />

1649<br />

Genome-wide scan of miRNA polymorphisms revealed<br />

a suggestive association of miR-3143 with chronic hepatitis<br />

B virus infection<br />

Daiki Miki 1,2 , Hidenori Ochi 1,2 , C. Nelson Hayes 1,2 , Hiromi Abe 1,2 ,<br />

Sakura Akamatsu 1,2 , Atsushi Ono 1,2 , Takashi Nakahara 1,2 , Yizhou<br />

Zhang 1,2 , Keiichi Masaki 1,2 , Hatsue Fujino 1,2 , Eisuke Miyaki 1,2 ,<br />

Hiromi Kan 1,2 , Takuro Uchida 1,2 , Nobuhiko Hiraga 1,2 , Masataka<br />

Tsuge 1,2 , Tomokazu Kawaoka 1,2 , Michio Imamura 1,2 , Yoshiiku<br />

Kawakami 1,2 , Hiroshi Aikata 1,2 , Kazuaki Chayama 1,2 ; 1 Laboratory<br />

for Digestive Diseases, RIKEN Center for Integrative Medical<br />

Sciences, Hiroshima, Japan; 2 Department of Gastroenterology and<br />

Metabolism, Hiroshima University Hostital, Hiroshima, Japan<br />

Background & Aims: Recent genome-wide association <strong>studies</strong><br />

(GWAS) have revealed several host genetic factors associated<br />

with various liver diseases, but the majority of genetic factors<br />

remain unclear. MicroRNAs (miRNAs) are small non-coding<br />

RNAs that control gene expression post-transcriptionally and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1013A<br />

may explain some of the missing heritability. Single nucleotide<br />

polymorphisms in mature or precursor miRNAs (miRNA-SNPs)<br />

could alter expression levels or binding-affinity to their target<br />

genes, resulting in changes of expression levels of many target<br />

genes. In this study, we conducted a large scale case-control<br />

study of chronic hepatitis B virus (HBV) infection focusing<br />

on miRNA-SNPs. Methods: We extracted autosomal SNPs<br />

in mature or precursor miRNA sequences using dbSNP137<br />

and miRBase19 and excluded indel variations and miR-SNPs<br />

with minor allele frequencies (AFs) < 1% in the Asian population,<br />

as reported by the 1000 Genomes database. These<br />

miR-SNPs were then genotyped in 2,564 chronic HBV patients<br />

and 1,128 healthy controls by multiplex-PCR-based Invader<br />

assay. Results: Based on a database search, we selected 315<br />

autosomal miR-SNPs with minor AFs > 1%, of which 93 were<br />

in mature miRNAs and the remaining 222 were in precursor<br />

miRNAs. We successfully genotyped 205 miR-SNPs and analyzed<br />

them using an additive model for genotype-phenotype<br />

association. We detected only one SNP that reached significance<br />

after adjustment for multiple testing (P = 7.7 × 10 -6<br />

< 0.05/205 = 2.4 × 10 -4 ). This SNP, rs17737028, located<br />

in the precursor of miR-3143, showed a significantly higher<br />

minor AF in HBV patients than in healthy controls (0.048 and<br />

0.027, respectively) with odds ratio 1.81. Because this miR-<br />

SNP was located on chromosome 6 and was not far from<br />

the MHC region containing HLA-DPs previously reported by<br />

GWAS as the locus most associated with chronic HBV infection,<br />

we decided to examine whether our result were affected<br />

by HLA-DP genotype. After adjustment for rs3077 in HLA-DPA1<br />

and rs9277535 in HLA-DPB1, we still found significant association<br />

between rs17737028 in pre-miR-3143 and chronic<br />

HBV infection, although the significance level was substantially<br />

reduced (P = 0.027, OR = 1.42) despite weak linkage<br />

between this miR-SNP and the 2 HLA-DP SNPs. Conclusions:<br />

We performed a genome-wide scan of miRNA polymorphisms<br />

and found that rs17737028 in pre-miR-3143 was significantly<br />

associated with chronic HBV infection, although the association<br />

might be affected considerably by HLA-DP genotype. Further<br />

validation <strong>studies</strong> and functional analyses of miR-3143 as well<br />

as its target genes are needed.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Daiki Miki, Hidenori Ochi, C.<br />

Nelson Hayes, Hiromi Abe, Sakura Akamatsu, Atsushi Ono, Takashi Nakahara,<br />

Yizhou Zhang, Keiichi Masaki, Hatsue Fujino, Eisuke Miyaki, Hiromi Kan, Takuro<br />

Uchida, Nobuhiko Hiraga, Masataka Tsuge, Tomokazu Kawaoka, Michio<br />

Imamura, Yoshiiku Kawakami, Hiroshi Aikata<br />

1650<br />

Hepatitis delta virus-induced T-cell recruiting chemokines<br />

contribute to inflammatory liver injury in chronic<br />

hepatitis delta<br />

Eui-Cheol Shin 1,2 , Nydiaris Hernandez-Santos 1 , Christopher Koh 1 ,<br />

Mary DeMino 1 , Theo Heller 1 , Barbara Rehermann 1 ; 1 Liver Diseases<br />

Branch, NIDDK, National Institutes of Health, Bethesda, MD;<br />

2 Korea Advanced Institute of Science and Technology, Daejeon,<br />

Korea (the Republic of)<br />

Hepatitis delta virus (HDV) is a small defective RNA virus that<br />

depends on hepatitis B virus (HBV) surface antigen to envelope<br />

its genome. Superinfection of chronic hepatitis B with<br />

HDV severely enhances inflammatory liver injury and progression<br />

to liver cirrhosis. The reason for the enhanced disease<br />

pathogenesis remains unknown. Here, we compared HDV<br />

and HBV-specific T cell and chemokine responses of 16 HDV/<br />

HBV co-infected patients to those of 25 HBV-infected patients.<br />

HDV/HBV co-infected patients had significantly higher ALT<br />

values than HBV-infected patients (medium 92 vs. 36 IU/L,<br />

p=0.001). Using IFN-gamma Elispot assays with overlapping<br />

HDV peptides we detected strong HDV-specific CD4 and CD8<br />

T cell responses in the peripheral blood of a patient with acute<br />

HDV/HBV infection and a patient who spontaneously cleared<br />

chronic HDV infection. However, HDV-specific T cell responses<br />

were barely detectable in chronic HDV/HBV co-infection, and<br />

were not unmasked by depletion of regulatory T cells or generation<br />

of T cell lines. Likewise, HBcore and HBsurface-specific<br />

CD4 and CD8 T cell responses were weak in HDV/HBV co-infected<br />

patients and did not differ from those of HBV-infected<br />

patients. Next, we quantitated serum chemokines known to<br />

recruit T cells to inflammatory sites. The serum levels of the<br />

two CXCR3-ligands CXCL9 and CXCL10, but not the chemokines<br />

CCL2, CCL3, CCL4 and CCL5 were significantly higher<br />

in HDV/HBV co-infected patients than in HBV-mono-infected<br />

patients. Serum CXCL9 and CXL10 levels correlated with ALT<br />

levels and histological activity score. Furthermore, intrahepatic<br />

CXCL9 and CXCL10 expression was greater in HDV/<br />

HBV co-infected patients than in HBV-mono-infected patients.<br />

To investigate whether HDV specifically induced these T cell-recruiting<br />

chemokines, we infected Huh-7 cells with recombinant<br />

HDV and demonstrated the production of CXCL9 and CXCL10<br />

at both RNA and protein level in a type I IFN-dependent manner.<br />

In summary, differences in the disease severity of HDV/<br />

HBV co-infection and HBV-infection were not attributed to the<br />

strength of the HDV and HBV-specific T cell response, but to<br />

HDV’s capacity to specifically induce T-cell recruiting chemokines<br />

such as CXCL9 and CXCL10.<br />

Disclosures:<br />

The following authors have nothing to disclose: Eui-Cheol Shin, Nydiaris Hernandez-Santos,<br />

Christopher Koh, Mary DeMino, Theo Heller, Barbara Rehermann<br />

1651<br />

Deletion of APOBEC3B is Associated With Hepatitis B<br />

eAg Status in Patients with Chronic HBV Infection<br />

Alex J. Thompson 1 , Ondrej Podlaha 2 , Zhaoshi Jiang 2 , Xin Guo 2 ,<br />

Luting Zhuo 2 , Dongliang Ge 2 , Anuj Gaggar 2 , Mani Subramanian<br />

2 , David B. Goldstein 3 , Henry Lik-Yuen Chan 4 , Edward J.<br />

Gane 5 , Harry L. Janssen 6 ; 1 St. Vincent’s Hospital and the University<br />

of Melbourne, Melbourne, VIC, Australia; 2 Gilead Sciences,<br />

Inc., Foster CIty, CA; 3 Institute for Genomic Medicine, Columbia<br />

University, New York, NY; 4 The Chinese University of Hong Kong,<br />

Hong Kong, China; 5 Auckland City Hospital, Auckland, New Zealand;<br />

6 University of Toronto, Toronto General Hospital, Toronto,<br />

ON, Canada<br />

Background: The prevalence of HBV infection varies across<br />

regions, with a higher proportion of HBeAg positive patients<br />

among Asian populations. APOBEC3 genes play a major<br />

role in intracellular defense against viral infection including<br />

HBV. Here we investigated the association between APO-<br />

BEC3B (A3B) deletions and HBeAg status among Asian and<br />

Caucasian patients. Methods: Asian (n=70) and Caucasian<br />

(n= 167) patients with available genomic samples from TDF<br />

HBV study (GS-US-174-0102/0103) were genotyped using<br />

Illumina Omini5M+Exome BeadChip platform. A3B deletions<br />

were inferred from SNP data using the PennCNV software.<br />

Allele frequency spectrum and linkage equilibrium methods<br />

were implemented on public human population data (1000<br />

Genomes Project) to test for signatures of recent positive selec-


1014A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tion of A3B deletion. Results: A3B deletions were detected in<br />

30% of Asian patients and 5% of Caucasian patients. We<br />

identified a significant association between A3B deletion and<br />

positive HBeAg status in Asian HBV patients (p = 0.002).<br />

Patients with this deletion were more likely to be HBeAg positive<br />

and remained HBeAg positive at an older age (Figure). A<br />

similar trend was observed among Caucasians patients though<br />

was not significant, likely due to a low A3B allele deletion frequency<br />

(p = 0.2). By analyzing the SNP data from the 1000<br />

Genomes Project, we observed a significant reduction of variation<br />

around A3B deletion region, suggesting a recent positive<br />

selection on A3B deletion allele in Asian, Indian, and Caucasian<br />

populations. Comparing A3B sequences across primates<br />

revealed a long-term positive selection signature, particularly in<br />

the human lineage. Conclusion: Structural variation of the APO-<br />

BEC3B gene is associated with HBeAg status among Asian<br />

patients. The high frequency of A3B deletion allele in Asian<br />

populations may be partially explained by positive selection<br />

forces unrelated to HBV. Further confirmation in independent<br />

patient cohorts is required.<br />

Disclosures:<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

Zhaoshi Jiang - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Scieces, Inc.<br />

Dongliang Ge - Employment: Gilead Sciences, Inc<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

David B. Goldstein - Advisory Committees or Review Panels: Sever Adverse<br />

Events Consortium; Grant/Research Support: Gilead, Astra Zeneca, Biogen,<br />

Johnson and Johnson, Labcorp, Merck, Fundazione Telethon, NIH, CURE, UCB,<br />

IL28B and ITPA finds<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Ondrej Podlaha, Xin Guo, Luting<br />

Zhuo<br />

1652<br />

Effect of a novel synthetic FXR agonist EYP001 on hepatitis<br />

B virus replication in HepaRG cell line and primary<br />

human hepatocytes<br />

Pauline Radreau 2 , Marine Porcherot 1 , Jacky Vonderscher 2 , Vincent<br />

Lotteau 1 , Patrice André 1 ; 1 CIRI U1111, INSERM, Lyon, France;<br />

2 EnyoPharma, Lyon, France<br />

HBV replication is tightly linked to bile acids (BA) metabolism.<br />

1) BA nuclear receptor FXR binds to two response elements in<br />

the HBV core promoter region and its activation by ligands<br />

regulates the HBV core promoter activity 2) the Na-taurocholate<br />

cotransporting polypeptide (NTCP) responsible of BA<br />

uptake by hepatocytes was identified as a functional receptor<br />

for HBV and 3) reciprocally, HBV infection induces changes<br />

of the expression of several genes involved in the BA synthesis<br />

pathway in the liver of infected patients. We investigated the<br />

effect of FXR activity modulation on HBV replication by a novel<br />

synthetic non-steroidal molecule EYP001 and compare its activity<br />

to the BA derived 6-ethyl-chenodeoxycholic acid (6-ECDCA)<br />

and the synthetic FXR agonist GW4064. First, we tested the<br />

effect of EYP001 1) on the expression of genes under the control<br />

of FXR in HepaRG cells and in primary human hepatocytes<br />

(PHH) and 2) on activation of the membrane associated BA<br />

receptor GpBAR1. Second, we tested the effect of these FXR<br />

modulators on HBV replication in HepaRG and in PHH, two<br />

cell culture models that support complete HBV replication cycle.<br />

Cells were infected with HBV produced by the HepG2.2.15<br />

line and treated from day 2 to 12 post infection with FXR modulators.<br />

EYP001 up-regulates the expression of SHP mRNAs<br />

and down-regulates the expression of the APOA1 mRNA and<br />

thus behaves as a FXR agonist. In addition treatment for 10<br />

days also reduces the expression of FXR mRNA, likely by an<br />

SHP mediated negative feedback. On the opposite, EYP001<br />

does not activate GpBAR1 and thus appears as a specific<br />

FXR agonist. Treatment of HBV infected dHepaRG and PHH<br />

with EYP001 as well as with 6-ECDCA or GW4064 strongly<br />

inhibited, to similar extent, the secretion of HBV DNA, HBsAg,<br />

HBeAg and of HBcAg synthesis in a dose dependent manner<br />

(70 to 80 % inhibition at 1 or 10 microMol) as well as the viral<br />

pregenomic RNA synthesis, cccDNA copies number and cellular<br />

total HBV DNA. Cyclosporine A, an NTCP ligand and HBV<br />

entry inhibitor, did not modify the effect of agonists suggesting<br />

that the effect did not depend on entry inhibition. Treatment<br />

consistently reversed the HBV induced modification of FXR and<br />

related genes expression. In conclusion, this novel FXR agonist<br />

led to a sustained repression of HBV replication in the HepaRG<br />

and PHH cell culture systems. This effect was likely mediated by<br />

the modulation of FXR activation that could perturb the complex<br />

FXR network of transcription factors, which is highly targeted<br />

and controlled by HBx. These data stress out the importance to<br />

exploit drug regulation of metabolism pathways in controlling<br />

HBV replication.<br />

Disclosures:<br />

Pauline Radreau - Employment: ENYO Pharma<br />

Vincent Lotteau - Consulting: enyopharma; Stock Shareholder: enyopharma<br />

Patrice André - Stock Shareholder: Enyo Pharma<br />

The following authors have nothing to disclose: Marine Porcherot, Jacky Vonderscher


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1015A<br />

1653<br />

HBcrAg during nucleos(t)ide analog therapy are related<br />

to intra-hepatic HBV replication and development of<br />

hepatocellular carcinoma<br />

Masao Honda, Takayoshi Shirasaki, Takeshi Terashima, Kazunori<br />

Kawaguchi, Mikiko Nakamura, Naoki Oishi, Tetsuro Shimakami,<br />

Hikari Okada, Kuniaki Arai, Taro Yamashita, Yoshio Sakai, Tatsuya<br />

Yamashita, Eishiro Mizukoshi, Shuichi Kaneko; Gastroenterology,<br />

Kanazawa University Graduate School of Medical Science,<br />

Kanazawa, Japan<br />

Background & Aim Although nucleos(t)ide analog (NA) therapy<br />

effectively reduces hepatitis B virus (HBV)-DNA in serum and<br />

improves liver histology in patients with chronic hepatitis B (CH-<br />

B), it does not completely reduce the incidence of hepatocellular<br />

carcinoma (HCC). The objective of this study is to elucidate clinical<br />

and virological features associated with the development<br />

of HCC during NA therapy. Material & Methods We enrolled<br />

109 patients with CH-B who started to receive NA therapy from<br />

2001 to 2011 at our hospital. All patients received NA therapy<br />

for more than 2 years, and a liver function test and virological<br />

markers before treatment and at the end of the follow-up period<br />

were compared. HBsAg was measured quantitatively using an<br />

Architect HBsAg-QT assay (Abbot). The amount of HBV-DNA in<br />

serum was measured with COBAS AmpliPrep-COBAS TaqMan<br />

HBV Test. HBcrAg was measured by CLEIA using a Lumipulse<br />

HBcrAg assay (Fujirebio, Inc., Tokyo, Japan). HBV-DNA, HBV-<br />

RNA and cccDNA in thirteen liver tissue samples taken before<br />

treatment and 23 liver tissue samples taken during treatment<br />

were analyzed. Gene expression profiling was performed by<br />

using an Affymetrix GeneChip. Results During NA therapy for<br />

6.5 ± 2.8 years, 36 patients (33%) developed HCC. Multivariate<br />

cox regression analysis showed age (>56 years, HR<br />

= 3.1) and fibrosis stage (F3–4, HR = 3.4) before treatment<br />

and the presence of hepatitis core-related antigen (HBcrAg;<br />

HR = 3.53) during treatment were significantly associated with<br />

the development of HCC. There was no significant difference<br />

in the amount of covalently closed circular DNA in HBcrAg(+)<br />

patients (n = 16) and HBcrAg(-) patients (n = 12); however,<br />

the amount of HBV-DNA and RNAs (pregenome, preS/S, and<br />

HBx) were significantly higher in HBcrAg(+) patients than in<br />

HBcrAg(-) patients, suggesting the presence of active HBV replication<br />

in HBcrAg(+) liver. Hepatic gene expression profiling<br />

before and during treatment (n = 13) showed a significant<br />

improvement of inflammation signaling during treatment; however,<br />

in HBcrAg(+) liver, viral replication-related signaling such<br />

as pro-apoptosis and oxidative phosphorylation were up-regulated<br />

along with cancer-related signaling (n = 11 vs. 12).<br />

Interestingly, transcription factors that activate HBV promoter<br />

activity, such as HNF4α and PPARα, were up-regulated in<br />

HBcrAg(+) liver. Metformin efficiently repressed HBV-RNA and<br />

HBcrAg levels by repressing the expression of HNF4α and<br />

PPARα in primary human hepatocytes. Conclusions: Modulating<br />

HBV transcription factors by metformin in combination with<br />

NA therapy would potentiate anti-HBV activity and reduce the<br />

incidence of HCC in HBcrAg (+) patients.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Masao Honda, Takayoshi Shirasaki,<br />

Takeshi Terashima, Kazunori Kawaguchi, Mikiko Nakamura, Naoki<br />

Oishi, Tetsuro Shimakami, Kuniaki Arai, Taro Yamashita, Yoshio Sakai, Tatsuya<br />

Yamashita, Eishiro Mizukoshi<br />

1654<br />

Free episomal and integrated HBV DNA In HBsAg-negative<br />

patients with intrahepatic cholangiocarcinoma<br />

Teresa Pollicino 1 , Cristina Musolino 2 , Gianluca Tripodi 2 , Marika<br />

Lanza 2 , Giuseppina Raffa 2 , Carlo Saitta 2 , Salvatore Benfatto 1 ,<br />

Concetta Beninati 1 , Giuseppe Navarra 3 , Pietro Invernizzi 4 , Domenico<br />

Alvaro 5 , Giovanni Raimondo 2 ; 1 Pediatric, Gynecological,<br />

Microbiological and Biomedical Sciences, University of Messina,<br />

Messina, Italy; 2 Clinical and Experimental Medicine, University<br />

Hospital of Messina, Messina, Italy; 3 Human Pathology, University<br />

Hospital of Messina, Messina, Italy; 4 Medicine, Humanitas<br />

Clinical and Research Center, Milan, Italy; 5 University of Rome<br />

“Sapienza”, Roma, Italy<br />

Backgroud: Intrahepatic cholangiocarcinoma (ICC) is a fatal<br />

primary liver cancer with very poor prognosis. Genome-wide<br />

<strong>studies</strong> have made major advances in understanding the molecular<br />

basis of this disease, although most aspects remain unclear.<br />

Accumulating evidence indicates that chronic HBV infection is<br />

associated with an increased risk of ICC development and<br />

suggests an etiological role of HBV in the development of this<br />

tumor. Aims of the study were to investigate the prevalence of<br />

occult HBV infection (OBI) in cases with ICC and to characterize<br />

the molecular status of HBV in OBI-positive ICC specimens.<br />

Methods: Frozen paired tumor and non-tumor tissue specimens<br />

from 40 HBsAg-negative patients with ICC, who underwent<br />

surgical resection were tested for OBI by 4 different HBV-specific<br />

nested PCR. To reveal HBV cccDNA, DNA extracts were<br />

digested with a plasmid-safe ATP-dependent DNase and amplified<br />

by nested PCR with cccDNA-specific primers. Finally, for<br />

the detection of HBV DNA integrations the Alu-PCR technique<br />

was coupled to deep-sequence analysis. Results: HBV genomic<br />

sequences were detected in tumor and/or non-tumor specimens<br />

from 28 of the 40 (70%) ICC patients analysed. In particular,<br />

20/40 (50%) tumors and 13/23 (56.5%) non-tumor tissues<br />

were HBV DNA positive. HBV cccDNA was detected in tissue<br />

specimens from 10/28 OBI-positive patients (36%) (both in<br />

tumor and non-tumor specimens in 3 patients; only in tumor<br />

tissues in 4 patients; only in non-tumor tissues in 3 patients).<br />

HBV integrants were detected in 3 of 10 cases examined so<br />

far, and included portions of the HBx gene sequence (including<br />

the Basic Core Promoter/Enhancer II) in 2 cases and part of the<br />

core gene sequence in one case. The analysis of the integration<br />

sites revealed that the HBx sequences were located 3,374<br />

nucleotides upstream the sequence encoding the cat eye syndrome<br />

critical region protein 5 isoform and within the coding<br />

sequence of the thromboxane A synthase 1, respectively, and<br />

that the core gene sequence was located within the cystinosin<br />

isoform 1 precursor coding sequence. Conclusion: Occult HBV<br />

infection is highly prevalent in patients with ICC. Both free viral<br />

genomes and integrated HBV DNA can be detected in these<br />

cases. These results suggest an involvement of HBV in the carcinogenic<br />

process leading to ICC development even in cases<br />

with occult infection.<br />

Disclosures:<br />

Teresa Pollicino - Speaking and Teaching: Gilead, Roche, Janssen, BMS, MSD,<br />

Bayer, Abbvie<br />

Giovanni Raimondo - Speaking and Teaching: BMS, Gilead, Roche, Merck,<br />

Janssen, Bayer, MSD<br />

The following authors have nothing to disclose: Cristina Musolino, Gianluca<br />

Tripodi, Marika Lanza, Giuseppina Raffa, Carlo Saitta, Salvatore Benfatto, Concetta<br />

Beninati, Giuseppe Navarra, Pietro Invernizzi, Domenico Alvaro


1016A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1655<br />

A Novel Entecavir Resistance Mutation RtA186T Causes<br />

Viral Breakthrough in Chronic Hepatitis B Patients<br />

Sanae Hayashi 1 , Shuko Murakami 1 , Katsumi Omagari 1 , Takeshi<br />

Matsui 2 , Etsuko Iio 1 , Masanori Isogawa 1 , Tsunamasa Watanabe<br />

1 , Yoshiyasu Karino 3 , Yuki Takamatsu 4 , Satoru Kohgo 5 , Kenji<br />

Maeda 5 , Hiroaki Mitsuya 4,5 , Yasuhito Tanaka 1 ; 1 Department of<br />

Virology & Liver unit, Nagoya City University Graduate School of<br />

Medical Sciences, Nagoya, Japan; 2 Center for Gastroenterology,<br />

Teine Keijinkai Hospital, Sapporo, Japan; 3 Department of Gastroenterology,<br />

Sapporo Kosei General Hospital, Hokkaido, Japan;<br />

4 Experimental Retrovirology Section, HIV and AIDS Malignancy<br />

Branch, National Cancer Institute, National Institutes of Health,<br />

Bethesda, MD; 5 National Center for Global Health and Medicine,<br />

Tokyo, Japan<br />

Background & Aim: Entecavir (ETV) is approved for the first-line<br />

treatment of chronic hepatitis B virus (HBV) infections due to a<br />

high genetic barrier. The aim of this study is to characterize<br />

two novel reverse transcriptase (RT) mutations associated with<br />

viral breakthrough (VBT) in an ETV-refractory patient. Methods:<br />

Serum HBV from an ETV refractory patient was sequenced<br />

before ETV treatment and after VBT. The clones with candidate<br />

ETV resistance (ETVr) mutations (rtI163T and rtA186T)<br />

were analyzed for replication efficacy and susceptibility to<br />

ETV, Adefovir (ADV), Tenofovir Disoproxil Fumarate (TDF), and<br />

novel nucleoside analogues (NAs) (4’-C-cyano-2-amino-2’-deoxyadenosine<br />

(CAdA), 4’-C-cyano-2’-deoxyguanosine (CdG))<br />

with similar efficacy to ETV in vitro. Moreover, chimeric mice<br />

with human hepatocytes were inoculated with the patient’s<br />

serum at VBT, and monitored for viral mutation pattern by<br />

next-generation sequencing. To confirm the prevalence of the<br />

novel mutations, 21 ETV-refractory patients were also examined.<br />

Results: The novel mutations, rtI163V and rtA186T, were<br />

detected together with LAMr at VBT, but not before ETV treatment.<br />

RtA186T significantly reduced viral replication efficacy,<br />

while rtI163V had no impact on it. RtA186T plus LAMr reduced<br />

susceptibility to ETV as strongly as previously reported ETVr<br />

mutations rtS202G plus LAMr, resulting in more than 111.1-<br />

fold resistance to ETV compared with the wild-type clone. In<br />

contrast, rtI163V plus LAMr resulted in a 20.4-fold resistance<br />

to ETV, suggesting that rt186T was mainly responsible for the<br />

VBT. The novel ETVr mutations did not confer cross-resistance<br />

to ADV, TDF, as well as novel NAs, CAdA and CdG. The viral<br />

mutation pattern in the chimeric mice indicated dominant proliferation<br />

of a clone containing rtI163V and rtA186T mutations<br />

plus LAMr under ETV treatment. In silico docking simulation<br />

indicated that rtA186T reduced the RT binding ability to ETV.<br />

In clinical practice, rtA186T, but not rtI163V, was detected<br />

in another ETV-refractory patient with VBT, suggesting that<br />

rtA186T could confer ETV resistance independently of rtI163V.<br />

Conclusion: A novel ETV resistance mutation rtA186T causes<br />

VBT, and should be closely monitored when chronic hepatitis B<br />

patients exhibit VBT during prolonged ETV treatment.<br />

Disclosures:<br />

Yoshiyasu Karino - Speaking and Teaching: BMS KK<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Sanae Hayashi, Shuko Murakami,<br />

Katsumi Omagari, Takeshi Matsui, Etsuko Iio, Masanori Isogawa, Tsunamasa<br />

Watanabe, Yuki Takamatsu, Satoru Kohgo, Kenji Maeda, Hiroaki Mitsuya<br />

1656<br />

Hepatitis B Virus Surface Protein-induced hPIAS1 Transcription<br />

Requires TAL1, E47, MYOG, NFI, andMAPK<br />

Signal Pathways<br />

Hongyan Wang, Di Wu, Xiaofeng Wang, Guang Chen, Yuanya<br />

Zhang, Weiming Yan, Meifang Han, Qin Ning; Department and<br />

institute of Infectious Disease,, Tongji Hospital, Tongji Medical College,<br />

Huazhong University of Science and Technology, Wuhan,<br />

China<br />

The protein inhibitor of activated STAT1 (PIAS1) is ubiquitous<br />

in mammals and fulfills key functions in cell proliferation and<br />

inflammation responses. Despite the importance of PIAS1 in<br />

regulating virus-induced or IFN-stimulated chronic hepatitis, little<br />

is known about the molecular mechanisms of activation in<br />

response to hepatic virus infection. Here, we report that the<br />

hepatitis B virus surface protein (HBs) enhancedhPIAS1 promoter<br />

activity in luciferase reporter assays. A strong regulatory<br />

region from −712 to −507 (relative to the transcriptional start<br />

site) was responsible forhPIAS1 gene transcription in response<br />

to HBs. TAL1, E47, myogenin (MYOG), and NFI were highly<br />

expressed and localized to the nucleus in response to HBs.<br />

Binding of TAL1, E47, MYOG, and NFI to a cis-regulatory<br />

element in the hPIAS1 promoter was confirmed by electrophoretic<br />

mobility shift assays and chromatin immunoprecipitation.<br />

siRNA knockdown of TAL1, E47, MYOG, and NFI<br />

expression inhibited hPIAS1transcription in response to HBs.<br />

Increased ERK and p38MAPK phosphorylation was observed<br />

in HBs-transfected HepG2 cells, and treatment with the ERK<br />

inhibitor PD098059 or p38MAPK inhibitor SB203580 abolished<br />

increase in nuclear TAL1, E47, MYOG, and NFI levels<br />

and hPIAS1 induction. Peripheral blood mononuclear cells from<br />

patients withhigh HBsAg levels (also withhigh levels of HBV<br />

DNA ) displayed increased ERK and p38MAPK phosphorylation<br />

and high levels of TAL1, E47, MYOG, and NFI, compared<br />

to those with low HBsAg levels or healthy controls. Increased<br />

hPIAS1 expression was detected in liver biopsies from CHB<br />

patients with high HBV replication compared to those of undetectable<br />

for HBV or healthy controls.These findings suggest that<br />

the HBs protein enhances hPIAS1transcription through activities<br />

of TAL1, E47, MYOG, and NFI that are dependent on MAPK<br />

signaling pathways.<br />

Disclosures:<br />

Qin Ning - Advisory Committees or Review Panels: ROCHE, NOVARTIS, BMS,<br />

MSD, GSK; Consulting: ROCHE, NOVARTIS, BMS, MSD, GSK; Grant/Research<br />

Support: ROCHE, NOVARTIS, BMS; Speaking and Teaching: ROCHE, NOVAR-<br />

TIS, BMS, MSD, GSK<br />

The following authors have nothing to disclose: Hongyan Wang, Di Wu,<br />

Xiaofeng Wang, Guang Chen, Yuanya Zhang, Weiming Yan, Meifang Han<br />

1657<br />

The activating receptor NKP46 is essential for the development<br />

of chronic hepatitis B<br />

Wanyu Li, Xiuzhu Gao, Ruqi Mei, Jinglan Jin, Junqi Niu; the first<br />

hospital of jilin university, Changchun, China<br />

BACKGROUND & AIMS: Our pervious study indicated that<br />

NKp46 expression regulated NK cell cytolytic function. NKp46<br />

may moderate NK cell activity during HBV replication suppression<br />

and HBV-associated liver damage in peripheral blood.<br />

Here we studied the functions of NKP46 during hepatitis B<br />

virus (HBV) infection in vivo and in the liver of patients. METH-<br />

ODS: Intrahepatic NKP46 was investigated in chronic hepatitis<br />

B patients(n=25) and healthy controls(n=15) by immunohistochemistry.We<br />

examined the effects of blocking antibodies<br />

against NKP46(50ug) in immunocompetent mice that express<br />

HBV from a pAAV/HBV1.2 plasmid(10ug) and are positive


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1017A<br />

for serum hepatitis B surface antigen (a mouse model of HBV<br />

infection)(n=10).Goat IgG isotype control antibody was administered<br />

to control groups.Serum HBsAg titers and HBVDNA<br />

levels were tested at baseline and post-injection. RESULTS:<br />

Intrahepatic NKP46 expression is up-regulated in patients with<br />

CHB compared with healthy control(p=0.032).After injecting<br />

HBV plasmid,mice that remained HBsAg-positive were used to<br />

test whether anti-NKP46 mAb treatment affected HBsAg and<br />

HBVDNA levels during HBV infection.Serum different value<br />

of HBsAg levels(1508.57±1654.94vs 5597.50±4729.13<br />

IU/ml;p


1018A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

entially expressed Genes & miRNAs between patient groups,<br />

key pathways & biological processes regulated were identified<br />

by using Volcano plot, miRTarbase & GOElite tool. Differentially<br />

expressed genes & miRNA along with pathways &<br />

biological categories regulated, were subjected to Integrome<br />

analysis to identify disease baseline & stage specific signatures.<br />

Results: In AVH-B compared to controls,1178 genes,42<br />

miRNAs were up-regulated & 1348 genes,9 miRNAs were<br />

down regulated. In CHB-PNALT compared to controls,167<br />

genes,1 miRNA were-up regulated & 307 genes,11 miRNAs<br />

were down-regulated. In CHB-RALT compared to controls,82<br />

genes,4 miRNAs were up-regulated & 58 genes,17 miRNAs<br />

were down-regulated. In CHB-PNALT compared to AVH-B,500<br />

genes,9 miRNAs were up regulated & 1307 genes,12 miR-<br />

NAs were down regulated. In CHB-RALTcompared to AVH-B<br />

1132 genes,26 miRNAs were up regulated &1694 genes,21<br />

miRNAs were down regulated. In CHB-RALT compared to CHB-<br />

PNALT, 183 genes, 3 miRNAs were up regulated & 70 genes,<br />

3 miRNAs were down regulated.We identified 267 genes &<br />

32 differentially expressed miRNAs,13 key pathways, biological<br />

categories & gene families in one or more of the infection<br />

stagesthat could play major role either in clearance or persistence<br />

of HBV infection (Fig.1). Conclusion: Integrome analysis<br />

revealed induction of unique cluster of miRNAs & repression<br />

of their target genes that differentiates acute & chronic infection<br />

stage, chronic infection with & without hepatic injury in DCs in<br />

HBV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Avishek K. Singh, Robert Geffers,<br />

Sheetalnath B. Rooge, Aditi Varshney, Madavan Vasudevan, Ankit Bhardwaj,<br />

Manoj Kumar, Pawan Malhotra, Sunil K. Mukherjee, Raj K. Bhatnagar, Nirupma<br />

Trehanpati, Shiv K. Sarin<br />

1660<br />

Functional restoration of CD56 bright NK cells via IL-15<br />

and NKG2D correlates with antiviral treatment efficacy<br />

in chronic hepatitis B patients<br />

Tao Chen 1 , Aichao Shi 1 , Xiaoping Zhang 1 , Lin Zhu 1 , Zeguang<br />

Wu 1 , Weiming Yan 1 , Xiaojing Wang 1 , Hong Ma 2 , Jidong Jia 2 ,<br />

Wei Guo 1 , Qin Ning 1 ; 1 Department and Institute of Infectious<br />

Disease, Tongji Hospital, Tongji Medical College, Huazhong<br />

University of Science and Technology, Wuhan, China; 2 Beijing<br />

Friendship Hospital, Beijing, China<br />

Hepatitis B virus is intrinsically immunogenic, and the immune<br />

status of the host has a prognostic impact on patients with<br />

chronic infection and influences the effects of antiviral treatment.<br />

This current study aimed to investigate the dynamic<br />

changes of NK cells post antiviral treatment and its potential<br />

influence on treatment efficacy. This study involved 52 hepatitis<br />

B e antigen (HBeAg) - positive chronic hepatitis B (CHB)<br />

patients who received telbivudine (Ldt) for 52 weeks. Blood<br />

samples were collected at baseline and week 12, 24, 36 and<br />

48 of treatment, and tested for HBV DNA, HBsAg, HBeAg, ALT,<br />

AST, and additional immunological markers. Compared with<br />

baseline, the percentages and absolute number of peripheral<br />

CD3 - CD56 + NK cells were significantly higher from week 36<br />

to week 48, especially CD3 - CD56 bright NK cells. The expression<br />

(percentage and MFI) of activating receptors NKG2D<br />

and NKP46 was enhanced, while inhibitory receptor NKG2A<br />

decreased. NKG2D expression was significantly enhanced on<br />

peripheral NK cells in patients with HBeAg seroconversion,<br />

particularly in CD3 - CD56 bright NK cell. The serum interleukin<br />

15 (IL-15) level significantly elevated during Ldt treatment,<br />

especially in patients with HBeAg seroconversion. Co-culture<br />

of Ldt with purified peripheral NK cells from treatment naïve<br />

HBeAg positive CHB patients showed significantly enhanced<br />

expression of NKG2D and IL-15. These findings indicate that<br />

antiviral treatment exerted by Ldt in CHB patients may play as<br />

an “adjuvant” role capable of inducing or restoring NK cell<br />

function via upregulated NKG2D and IL-15 expression, and<br />

this in turn correlated with HBeAg seroconversion.<br />

Disclosures:<br />

Jidong Jia - Consulting: BMS, GSK, MSD, Novartis, Roche<br />

Qin Ning - Advisory Committees or Review Panels: ROCHE, NOVARTIS, BMS,<br />

MSD, GSK; Consulting: ROCHE, NOVARTIS, BMS, MSD, GSK; Grant/Research<br />

Support: ROCHE, NOVARTIS, BMS; Speaking and Teaching: ROCHE, NOVAR-<br />

TIS, BMS, MSD, GSK<br />

The following authors have nothing to disclose: Tao Chen, Aichao Shi, Xiaoping<br />

Zhang, Lin Zhu, Zeguang Wu, Weiming Yan, Xiaojing Wang, Hong Ma, Wei<br />

Guo<br />

1661<br />

Knockdown of NTCP reduces susceptibility to HBV infection<br />

in humanized-liver mice<br />

Tasuku Nakabori, Hayato Hikita, Satoshi Aono, Yoshinobu<br />

Yokoyama, Kazuhiro Murai, Takuo Yamai, Yugo Kai, Yuki Makino,<br />

Yasutoshi Nozaki, Yoshinobu Saito, Satoshi Tanaka, Ryotaro<br />

Sakamori, Tomohide Tatsumi, Tetsuo Takehara; Osaka University<br />

Graduate School of Medicine, Suita, Japan<br />

Background and Aim: Hepatitis B Virus (HBV) infects very limited<br />

species such as humans and chimpanzees, which hamper<br />

HBV research. Humanized-liver mice have potential to resolve<br />

these difficulties. Recently, Na+-taurocholate cotransporting<br />

polypeptide (NTCP) was reported as a candidate of HBV entry<br />

receptor. However, the effect of its inhibition on HBV infection<br />

remains unclear. In this study, we clarified it using humanized-liver<br />

mice. Methods: TK-NOG mice were administrated<br />

ganciclovir followed by transplantation of human primary<br />

hepatocytes. Humanized-liver TK-NOG mice (humanized mice)<br />

were inoculated with HBV obtained from a chronic hepatitis B<br />

patient. The institutional ethics committee approved the study.<br />

For in vitro study, primary hepatocytes isolated from humanized<br />

mice were inoculated with HBV derived from the culture supernatant<br />

of 2.2.15 cells. Results: Serum HBV-DNA were detected<br />

in humanized mice from 2 weeks after HBV inoculation. After<br />

8 to 10 weeks post-inoculation, HBV-DNA levels reached a<br />

plateau (6-8 log copies/ mL) and almost all of human hepatocytes<br />

were positive for HBc antigen by immunohistochemistry.<br />

Administration of siRNA against human NTCP by tail vein successfully<br />

reduced its expression level in human hepatocytes,<br />

evidenced by immunohistochemistry and real time RT-PCR.<br />

After 2 weeks of HBV inoculation following administration of<br />

NTCP siRNA, the levels of serum HBV-DNA, serum HBs antigen<br />

and intracellular covalently closed circular DNA (cccDNA) were<br />

significantly reduced. In sharp contrast, siRNA-mediated NTCP<br />

knockdown at 12 weeks after HBV inoculation did not affect<br />

them. Primary hepatocytes from humanized mice were maintained<br />

in culture over 2 months. After inoculation with HBV for<br />

24 hours, HBV-DNA, HBs antigen and HBe antigen levels in<br />

culture medium, gradually increased and kept detectable over<br />

a month. cccDNA was also detected. Immunohistochemistry


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1019A<br />

using anti-HBc antibody confirmed their infection with HBV.<br />

siRNA-mediated knockdown of NTCP efficiently downregulated<br />

its expression level. Consistent with in vivo data, transfection of<br />

NTCP siRNA before inoculation with HBV significantly reduced<br />

HBs antigen, HBe antigen levels in the culture medium, and<br />

intracellular cccDNA, while its transfection after 20 days of<br />

HBV inoculation did not affect them. Conclusion: Humanized<br />

mice and their isolated primary hepatocytes are susceptible to<br />

HBV infection. Our results suggest that NTCP inhibition would<br />

be useful for decrease of susceptibility to HBV infection but that<br />

NTCP inhibition alone might not have an effect on persistent<br />

infection with HBV.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Tasuku Nakabori, Satoshi Aono,<br />

Yoshinobu Yokoyama, Kazuhiro Murai, Takuo Yamai, Yugo Kai, Yuki Makino,<br />

Yasutoshi Nozaki, Yoshinobu Saito, Satoshi Tanaka, Ryotaro Sakamori, Tomohide<br />

Tatsumi<br />

1662<br />

The Association of vitamin D metabolic pathway<br />

related genes polymorphisms with virologic response in<br />

HBeAg-negative patients treated with pegylated interferon:<br />

Multicenter study<br />

Kessarin Thanapirom 1 , Sirinporn Suksawatamnuay 1 , Wattana<br />

Sukeepaisarnjaroen 2 , Tawesak Tanwandee 3 , Satawat Thongsawat<br />

4 , Teerha Piratvisuth 5 , Rattana Boonsirichan 6 , Chalermrat<br />

Bunchorntavakul 7 , Chaowalit Pattanasirigool 8 , Bubpha Pornthisarn<br />

9 , Supot Tantipanichtheerakul 10 , Ekawee Sripariwuth 11 , Woramon<br />

Jeamsripong 12 , Theeranun Sanpajit 13 , Piyawat Komolmit 1 ;<br />

1 Medicine, Chulalongkorn University, Bangkok, Thailand; 2 Medicine,<br />

Khon Kaen University, Khon Kaen, Thailand; 3 Medicine,<br />

Mahidol University, Bangkok, Thailand; 4 Medicine, Chiang Mai<br />

University, Chiang Mai, Thailand; 5 Medicine, Prince of Songkla<br />

University, Songkla, Thailand; 6 Medicine, Vajira Hospital, Bangkok,<br />

Thailand; 7 Medicine, Rajavithi Hospital, Bangkok, Thailand;<br />

8 Medicine, Police General hospital, Bangkok, Thailand; 9 Medicine,<br />

Thammasat University Hospital, Pathum thani, Thailand;<br />

10 Medicine, Bhumibol Adulyadej Hospital, Bangkok, Thailand;<br />

11 Medicine, Naresuan University, Phitsanulok, Thailand; 12 Medicine,<br />

Buddhachinaraj Hospital, Phitsanulok, Thailand; 13 Medicine,<br />

Phramongkutklao Hospital, Bangkok, Thailand<br />

Background/aims: In chronic hepatitis B (CHB)-infected<br />

patients, low serum vitamin D level are associated with high<br />

level of hepatitis B virus (HBV) replication. From GWAS data,<br />

the number of single nucleotide polymorphisms (SNPs) within<br />

the 7-dehydrocholesterol reductase (DHCR7), 1-a-hydroxylase<br />

(CYP27B1), Cytochrome P450, family 2, subfamily R, polypeptide<br />

1 (CYP2R1), vitamin D binding protein (GC) and vitamin<br />

D receptor (VDR) of the vitamin D synthetic pathway is related<br />

with vitamin D level and function. This study aimed to determine<br />

the association between the SNPs of vitamin D cascade<br />

and response to peginterferon (PegIFN) therapy in patients<br />

with HBeAg-negative CHB infection. Methods: 115 patients<br />

with HBeAg-negative CHB infection treated for 48 weeks with<br />

PegIFN-alfa 2a from 13 hospitals were prospectively enrolled.<br />

Thirteen SNPs across the vitamin D cascade related genes,<br />

including DHCR7 (rs12785878), CYP27B1 (rs10877012),<br />

CYP2R1 (rs2060793, rs12794714), GC (rs4588, rs7041,<br />

rs222020, rs2282679) and VDR (FokI, BsmI, Tru9I, ApaI,<br />

TaqI) were genotyped. The virologic response was defined<br />

as HBV-DNA < 2,000 IU/ml. Results: Majority of patients<br />

(81.7%) had HBV genotype C and 5.3% (n=5) had liver cirrhosis.<br />

At 24 weeks after therapy, 55.7% (n=54) achieved<br />

sustained virologic response (SVR) and 6.3% (n=6) cleared<br />

HBsAg. Seventy-two patients (79.1%) gained end-of-therapy<br />

virologic response (ETVR). The non-CC allele of FokI (85.3%)<br />

was significant associated with higher ETVR than CC allele of<br />

FokI (60.9%)(p


1020A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

PlasmaBcells(4.1±3.8%Vs12.1±6.45%, p=0.03)&humoral<br />

immunity were also reduced in Gr A(Fig.B,C).RNA sequencing<br />

showed decreased expression (>2 fold) of TFH and B cell<br />

related genes;CD4, BCL6, ICOS, SLAMF1,CXCR5 and BAF-<br />

F,BATF,CD40, BAFFR in GrA than B(Fig1A).This was validated<br />

by qRT-PCR showing down-regulation of TFH related genes<br />

(Fig1D).ROC showed significance of CD4,TFH and B cells<br />

with cut-off values for prediction of transmission; with highest<br />

specificity&sensitivity with CD4+ICOS+ & CD4+CXCR5+cells<br />

as predictors for a transmitting mother(Table).Conclusions:Humoral<br />

Immunity is essential for viral neutralization.Decreased<br />

CD4 helper&TFH cells lead to impaired plasma B cell function&facilitate<br />

HBV transmission to the newborn. Plasma B cell<br />

can be used as a marker to predict HBV Transmission.<br />

beclin-1 (6 fold), SIRT-1 (4 fold), c-myc (4 fold), and PTEN with<br />

HA treatment (100 μg/ml). The expression of p-akt, p-erk-1/2<br />

were also inhibited while the expression of the caspase-3 and<br />

β-catenin were found to be up-regulated. The autophagosome<br />

formation was inhibited significantly (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1021A<br />

1666<br />

Presence of naïve-like CD8+ T cells specific for subdominant<br />

hepatitis B virus epitopes in chronically infected<br />

patients<br />

Anita Schuch 1,2 , Muthamia Kiraithe 1,2 , Julia Lang 1 , Christoph Neumann-Haefelin<br />

1 , Robert Thimme 1 ; 1 Department of Medicine II, University<br />

Hospital Freiburg, Freiburg, Germany; 2 Faculty of Biology,<br />

University of Freiburg, Freiburg, Germany<br />

In chronically HBV-infected patients, virus-specific CD8+ T<br />

cells are rarely detectable ex vivo by conventional tetramer<br />

stainings. We aimed to analyze whether his lack of HBV-specific<br />

CD8+ T-cell responses may be explained by insufficient<br />

priming or terminal deletion. To phenotypically characterize<br />

HBV-specific CD8+ T cells that are not detectable by conventional<br />

methods we used a combination of tetramer-associated<br />

magnetic bead enrichment and multiparametric flow cytometry<br />

and additionally performed viral sequence analyses. We analyzed<br />

CD8+ T-cell populations specific for 4 well-described<br />

HLA-A*02:01 restricted HBV epitopes (Core 18-27<br />

, Env 183-191<br />

,<br />

Env 335-343<br />

and Pol 455-463<br />

) in a heterogeneous cohort of 34<br />

chronically HBV-infected patients. We were able to detect 108<br />

of 136 (79%) HBV-specific CD8+ T-cell populations. Of note,<br />

each patient had at least one detectable CD8+ T-cell population,<br />

suggesting that these cells are not completely deleted<br />

in chronic infection. Enriched CD8+ T cells specific for the<br />

immunodominant Core 18-27<br />

epitope predominantly showed an<br />

antigen-experienced effector-memory phenotype (CD45RA - ,<br />

CCR7 - , CD27 + , CD11a + ) which was also detectable in patients<br />

with resolved HBV infection. In contrast, CD8+ T cells specific<br />

for Env 183-191<br />

, Env 335-343<br />

and Pol 455-463<br />

often displayed a naïvelike<br />

phenotype (15%, 44% and 19% respectively) defined by<br />

high expression of CD45RA, CCR7 and CD27. Absence of<br />

CD11a and CD95 expression additionally indicated a naïve<br />

status of these enriched HBV-specific CD8+ T cells. Importantly,<br />

the presence of wild type sequences in patients with naïve-like<br />

HBV-specific CD8+ T cells indicated that the presence of naïvelike<br />

cells is not due to infection with a variant HBV containing<br />

mismatched epitopes. It is also important to note that the presence<br />

of HBV-specific CD8+ T cells with naïve-like phenotype<br />

did not correlate with any clinical parameter such as viral load,<br />

transaminase levels or HBeAg status, and was independent<br />

of antiviral treatment. Taken together, our results indicate that<br />

HBV-specific CD8+ T cells are not completely deleted in chronically<br />

infected patients, but are rather maintained at a very low<br />

frequency. Furthermore, a substantial fraction of CD8+ T cells<br />

specific for subdominant HBV epitopes display a naïve-like<br />

phenotype despite ongoing viral replication. This suggests that<br />

insufficient priming and/or antigen ignorance of HBV-specific<br />

CD8+ T cells may also contribute to CD8+ T-cell failure in<br />

chronic HBV infection and that therapeutic vaccination may be<br />

a feasible approach in this patient cohort.<br />

Disclosures:<br />

Christoph Neumann-Haefelin - Advisory Committees or Review Panels: AbbVie;<br />

Speaking and Teaching: AbbVie, Gilead, Bristol-Myers Squibb<br />

The following authors have nothing to disclose: Anita Schuch, Muthamia Kiraithe,<br />

Julia Lang, Robert Thimme<br />

1667<br />

Precore and Basal Core Promoter mutations in HBeAg<br />

negative chronic hepatitis B virus infection in Cape<br />

Town, South Africa<br />

Mark W. Sonderup 1 , Heidi Smuts 2 , Ruud A. Roos 3,1 , Helen<br />

Wainwright 4 , Neliswa A. Gogela 1 , Mashiko Setshedi 1 , Henry N.<br />

Hairwadzi 1 , C. W. Spearman 1 ; 1 Department of Medicine and<br />

Division of Hepatology, University of Cape Town, Cape Town,<br />

South Africa; 2 Division of Medical Virology, University of Cape<br />

Town and NHLS, Cape Town, South Africa; 3 Vrije Universiteit,<br />

Amsterdam, Netherlands; 4 Department of Anatomical Pathology,<br />

University of Cape Town, Cape Town, South Africa<br />

Chronic HBeAg negative HBV represents the immune control<br />

or escape phase of infection. Precore (PC) and basal core<br />

promoter (BCP) mutations are associated with progressive liver<br />

disease or HCC although this is poorly studied in sub-Saharan<br />

Africa. Our aim was to characterize mutations in HBeAg negative<br />

patients. Methods: Prospectively, over 2 years, HBeAg negative<br />

patients were analysed by amplifying viral PC and BCP<br />

regions using nested primers. Amplicons were direct Sanger<br />

sequenced in the forward & reverse directions and compared<br />

to reference sequences. Relevant demographic, clinical, virological<br />

and histological data were captured. Results: 124<br />

patients, median age 35.1 yrs (IQR 29-43.2), 60.5% (n=75)<br />

male, 3.2% (n=4) HIV co-infected, were analysed. Ethnically,<br />

52% were mixed ancestry, 28% Black African, 16% White<br />

and 4% Asian. Genotype distribution - A 51%, B 2.3%, C<br />

1.1%, D 42% and E 3.6%. Mutations were detected in 98.4%<br />

(n=122); 89% (n=109) with PC; 53.3% (n=65) BCP and 46%<br />

(n=56) both. Mutation frequencies are listed. Median ALT (U/L)<br />

was significantly higher when both PC/BCP than PC only mutations<br />

were present, [44 (IQR 41-47); 27 (19-52)] p=0.0007,<br />

but not with BCP only mutations [32 IQR (22-65)]; p=0.069.<br />

With dual A1762T G1764A BCP mutations, median ALT and<br />

HBV DNA (log 10<br />

IU/ml) was significantly higher compared to<br />

the T1753C only BCP mutation, [52 (IQR 25-134); 24 (IQR21-<br />

49)]; p=0.019 and [5.2 (IQR 4-7.6); 3.8 (3-5.3)]; p=0.02.<br />

In those biopsied (n=31), median Ishak necro-inflammatory<br />

stage and fibrosis grade scores were 5 (IQR 3-7) & 2 (IQR 1-3)<br />

respectively with no differences observed in PC, BCP or PC/<br />

BCP groups. Six patients, 4.8%, developed HCC with 5 (83%)<br />

having both PC/BCP mutations, 3 (50%) the dual A1762T<br />

G1764A BCP mutation and 4 (66%) the G1896A PC mutation.<br />

Conclusion: In HBeAg negative, mostly A or D genotype CHB<br />

patients, PC were present more frequently than BCP mutations;<br />

almost half having both. Dual PC/BCP predicted for higher ALT<br />

and HBV DNA; a similar trend observed for the dual A1762T<br />

G1764A BCP mutation. Mutation analysis may have a role in<br />

guiding the need for antiviral therapy.


1022A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Mark W. Sonderup, Heidi Smuts,<br />

Ruud A. Roos, Helen Wainwright, Neliswa A. Gogela, Mashiko Setshedi, Henry<br />

N. Hairwadzi, C. W. Spearman<br />

1668<br />

Differential Serum Cytokine Profiles in Patients with<br />

Hepatitis B Virus (HBV), Hepatitis C Virus (HCV), and<br />

Non-Viral Non-Autoimmune Liver Disease, with or without<br />

Hepatocellular Carcinoma (HCC)<br />

Vincent L. Chen 2 , Philip Vutien 3,5 , Biao Li 4 , Ondrej Podlaha 4 ,<br />

Ellen T. Chang 5 , Zhaoshi Jiang 4 , Dongliang Ge 4 , Anuj Gaggar 4 ,<br />

Mindie H. Nguyen 1 ; 1 Division of Gastroenterology and Hepatology,<br />

Stanford University Medical Center, Palo Alto, CA; 2 Department<br />

of Medicine, Stanford University Medical Center, Stanford,<br />

CA; 3 Rush University Medical Center, Chicago, IL; 4 Gilead Sciences,<br />

Inc., Foster City, CA; 5 Department of Health Research and<br />

Policy (Epidemiology), Stanford University School of Medicine,<br />

Stanford, CA<br />

Aims: HBV and HCV are the most common causes of HCC.<br />

Differences in immune responses to HBV, HCV, non-viral disease,<br />

and HCC remain incompletely defined. Our goal was<br />

to compare the cytokine profiles of patients with HBV, HCV, or<br />

non-viral liver disease, with or without HCC. Methods: Serum<br />

samples were obtained from 381 patients seen at a US medical<br />

center from 2000-2007: 78 patients with HCC (33 HBV,<br />

45 HCV) and 303 without HCC (119 HBV, 131 HCV, 53<br />

non-viral). Microplex analysis (Luminex 200 IS) was used to<br />

measure serum levels of 51 common cytokines. Random forest<br />

machine learning was used to generate receiver operator<br />

characteristic (ROC) curves and determine individual cytokine<br />

importance, using Z scores of mean fluorescence intensity for<br />

individual cytokines. Results: Among non-HCC patients, 60%<br />

were male with mean age 52 years. Most HBV patients were<br />

Asian (89%), while only 18% of HCV and 4% of non-viral<br />

patients were Asian. Among HCC patients, 85% were male<br />

with mean age 62 and similar ethnic distribution. Cytokine concentrations<br />

distinguished non-HCC HBV and HCV patients, with<br />

area under the ROC curve (AUC) of 0.91 (Fig. 1A). The cytokines<br />

with greatest predictive power were soluble FasL, M-CSF,<br />

TNF-β, and IP-10. Cytokine patterns also differed between non-<br />

HCC HBV and non-viral patients (Fig. 1A; AUC 0.84), and the<br />

most important cytokines were leptin, resistin, IL-8, and M-CSF.<br />

Cytokine profiles were less distinct between HCV vs. non-viral<br />

liver disease (AUC 0.72). Of note, the cytokine pattern of HCC<br />

patients was different from that of non-HCC patients in HBV<br />

(Fig. 1B; AUC 0.77) but not HCV cohort (Fig. 1B; AUC 0.48).<br />

Finally, cytokine profiles differed between non-HCC patients<br />

with alanine aminotransferase ≥ 2x vs. < 2x upper limit of normal<br />

(AUC 0.75) but not between Asian and non-Asian patients<br />

(AUC 0.55 for HBV and 0.51 for HCV patients). Conclusions:<br />

Serum cytokine profile distinguishes non-HCC patients with<br />

HBV from those with HCV or non-viral disease, and between<br />

HBV patients with and without HCC. Future work will expand<br />

the cohort, perform additional subgroup analysis, and identify<br />

cytokine profiles that correlate with disease progression.<br />

Disclosures:<br />

Biao Li - Employment: Gilead Sciences, Inc.<br />

Ellen T. Chang - Employment: Exponent, Inc. (a for-profit science and engineering<br />

consulting company that works on behalf of numerous health care clients, among<br />

others)<br />

Zhaoshi Jiang - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Scieces, Inc.<br />

Dongliang Ge - Employment: Gilead Sciences, Inc<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Vincent L. Chen, Philip Vutien,<br />

Ondrej Podlaha<br />

1669<br />

Possible involvement of hepatic steatosis in the elevation<br />

of aminotransferases in hepatitis B virus-infected<br />

patients who are HBeAg-negative and have low DNA<br />

levels<br />

Hirayuki Enomoto, Nobuhiro Aizawa, Yoshiyuki Sakai, Naoto<br />

Ikeda, Kazunori Yoh, Ryo Takata, Chikage Nakano, Kunihiro Hasegawa,<br />

Kyohei Kishino, Yoshihiro Shimono, Akio Ishii, Tomoyuki<br />

Takashima, Takashi Nishimura, Hiroki Nishikawa, Yoshinori<br />

Iwata, Hiroko Iijima, Shuhei Nishiguchi; Division of Hepatobiliary<br />

and Pancreatic Medicine, Department of Internal Medicine, Hyogo<br />

College of Medicine, Nishinomiya, Japan<br />

Background: The clinical significance of hepatic steatosis in<br />

hepatitis B virus (HBV)-infected patients who are HBeAg-negative<br />

and have low HBV-DNA levels (referred to as asymptomatic<br />

carriers) has not been sufficiently clarified. Methods:<br />

Among a total of 112 treatment-naïve HBV-infected patients<br />

with a liver biopsy, we assessed the histological and laboratory<br />

findings to evaluate whether hepatic steatosis and its<br />

related metabolic disorders were associated with an elevation<br />

in ALT levels in HBeAg-negative patients (N=70). The histological<br />

degrees of hepatic steatosis were classified as Grade 0<br />

(33%-66%), and Grade<br />

3 (>66%) as described by Kleiner et al. Clinical variables with<br />

suggested associations with hepatic steatosis, such as the levels<br />

of homeostatic model assessment insulin resistance (HOMA-IR)<br />

and serum ferritin were measured. We also calculated the<br />

NAFIC score, which is a simple scoring system determined by<br />

the levels of three variables (serum ferritin concentration, fasting<br />

insulin, and type IV collagen 7S) that has been proposed<br />

for predicting nonalcoholic steatohepatitis (NASH) in Japanese<br />

patients. The study conformed to the ethical guidelines<br />

of the Declaration of Helsinki, and written informed consent<br />

was obtained from all patients on admission. This study was<br />

approved by the ethics committee of the institutional review<br />

board at our institution. Results: In HBeAg-negative patients<br />

with high HBV-DNA levels (HBV-DNA ≥ 2000 IU/ml), the elevation<br />

of the ALT levels was related to the HBV-DNA levels<br />

independent of the histological severity of hepatic steatosis


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1023A<br />

and the levels of hepatic steatosis-related metabolic variables.<br />

However, in HBeAg-negative patients with low DNA levels<br />

(HBV-DNA < 2000 IU/ml), the ratio of patients with moderate<br />

to severe hepatic steatosis (Grades 2 or 3) and the levels of<br />

hepatic steatosis-related variables were significantly higher in<br />

patients with elevated ALT levels than those without (steatosis:<br />

53.8% vs. 5.0%, p


1024A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

untreated cells. Finally, the HBV enhancer-1 activity was examined<br />

by luciferase assay using luciferase vector inserted HBV<br />

enhancer-1 region. GGA treatment resulted in approimately<br />

50% reduction of HBV enhancer-1 activity in Huh7 cells compared<br />

to untreated cells. (Conclusion) GGA reduces HBV<br />

related protein and mRNA in Huh7 hepatoma cells by suppressing<br />

HBV enhancer-1 activity. GGA may increase anti-HBV<br />

effect in combination with other therapies.<br />

Disclosures:<br />

The following authors have nothing to disclose: Masafumi Haraguchi, Satoshi<br />

Miuma, Yuko Akazawa, Hidetaka Shibata, Takuya Honda, Hisamitsu Miyaaki,<br />

Naota Taura, Tatsuki Ichikawa, Kazuhiko Nakao<br />

1672<br />

Dynamic Analysis Of Cellular Factors Modulated By<br />

Hepatitis B Virus Protein Expression And Replication<br />

Kenichi Morikawa 1 , Tomoe Shimazaki 1 , Takaaki Izumi 1 , Machiko<br />

Umemura 1 , Masato Nakai 1 , Goki Suda 1 , Darius Moradpour 2 ,<br />

Naoya Sakamoto 1 ; 1 Department of Gastroenterology and Hapatology,<br />

Hokkaido University, Sapporo, Japan; 2 University of Lausanne,<br />

Lausanne, Switzerland<br />

Background: Many viruses manipulate host factors for their<br />

purposes. The hepatitis B virus (HBV) is believed to regulate<br />

host factors for viral replication, persistence and pathogenesis.<br />

However, the mechanisms by which HBV establishes and<br />

maintains chronic infection are poorly understood. The aim of<br />

this study was to perform a dynamic analysis of cellular factors<br />

modulated by HBV for a better understanding of the life cycle<br />

and pathogenesis of HBV. Methods: Huh7 cells inducibly replicating<br />

HBV were analyzed by stable isotopic labeling using<br />

amino acids in cell culture (SILAC) coupled with mass spectrometry.<br />

In these cells, the core and the polymerase are inducibly<br />

expressed upon tetracycline withdrawal from the culture<br />

medium while pre-S1, pre-S2, and S as well as X are constitutively<br />

expressed from endogenous HBV promoters. Metabolites<br />

were analyzed by mass spectrometry in the same cell lines.<br />

Oxidative stress was assessed by the GSH/GSSG ratio. Mitochondrial<br />

superoxide was measured by MitoSOX using flow<br />

cytometry. Morphological changes of the endoplasmic reticulum<br />

and mitochondria, as well as autophagosome induction<br />

were visualized by negative stain transmission electron microscopy<br />

(TEM). Results: SILAC coupled with mass spectrometry<br />

identified > 3800 proteins, and mass spectrometry measured<br />

112 metabolites. We observed an increase in the abundance<br />

of mitochondrial proteins and a decrease of ribosomal proteins<br />

upon replication of HBV. There was no significant difference<br />

in mitochondrial morphology between cells replicating<br />

vs. not replicating HBV. More cells containing lysosomes and<br />

autophagosomes were observed in the presence of replicating<br />

HBV. HBV replication was associated with the production of<br />

mitochondrial superoxide. However, only a slight increase of<br />

oxidative stress was noted. Conclusions: The observed changes<br />

suggest a slow down or stop of cell cycle and proliferation for<br />

viral fitness. The defense mechanism against oxidative stress<br />

induced by HBV might be maintained by autophagy. Analyses<br />

of cellular dynamics should yield new insights into the HBV life<br />

cycle and the pathogenesis of hepatitis B and may reveal novel<br />

angles for therapeutic intervention.<br />

Disclosures:<br />

Naoya Sakamoto - Grant/Research Support: Gilead Sciences, TRSS; Speaking<br />

and Teaching: Bristol Myers Squibb, Janssen Pharm, Chugai co ltd<br />

The following authors have nothing to disclose: Kenichi Morikawa, Tomoe Shimazaki,<br />

Takaaki Izumi, Machiko Umemura, Masato Nakai, Goki Suda, Darius<br />

Moradpour<br />

1673<br />

IFN-λ3 induction by nucleotide, not nucleoside, analogs<br />

and its clinical significance in HBV patients<br />

Kazumoto Murata 1 , Akihiro Matsumoto 2 , Taisuke Inoue 3 , Minoru<br />

Sakamoto 3 , Masaya Sugiyama 1 , Nobuyuki Enomoto 3 , Eiji<br />

Tanaka 2 , Masashi Mizokami 1 ; 1 The Research Center for Hepatitis<br />

and Immunology, National Center for Global Health and Medicine,<br />

Ichikawa, Japan; 2 Department of Medicine, Shinshu University<br />

School of Medicine, Matsumoto, Japan; 3 First Department of<br />

Medicine, University of Yamanashi, Chuo, Japan<br />

BACKGROUND & AIM: Clinical significances of IL-28B polymorphisms<br />

in patients with hepatitis B virus (HBV) infection are<br />

controversial while its pivotal roles were established in chronic<br />

hepatitis C. Therefore, we sought to clarify the physiological<br />

roles of its protein, IFN-λ3 levels in HBV patients. METHODS &<br />

RESULTS: We measured serum IFN-λ3 levels together with IFNλ1<br />

or IFN-λ2 in asymptomatic carriers (n=83), chronic hepatitis<br />

(n=113), and liver cirrhosis (n=58). No patients were treated<br />

with IFN when serum was taken. IFN-λ3 was measured by<br />

our newly developed chemiluminescence enzyme immunoassay<br />

(CLEIA) kit, which demonstrated little or no cross reactivity<br />

between IFN-λ2 and IFN-λ3. IFN-λ1 or IFN-λ2 was measured<br />

by commercial ELISA kit. No significant differences in serum<br />

IFN-λ3 levels were observed in disease progression, IL-28B<br />

polymorphisms or HBeAg positivity. Unexpectedly, however,<br />

we found higher serum IFN-λ3 levels in patients treated with<br />

adefovir pivoxil (ADV) or tenofovir disoproxil fumarate (TDF),<br />

compared with lamivudine (LAM) or entecavir (ETV) (35.4 ±<br />

27.6 pg/ml vs 3.0 ± 4.9 pg/ml, p < 0.0001) although no differences<br />

in their clinical background were observed. IFN-λ1 or<br />

IFN-λ2 did not reveal these associations. Serial measurements<br />

of serum IFN-λ3 in each patient demonstrated its elevation after<br />

initiation of ADV or TDF and decline after cessation of it, suggesting<br />

ADV or TDF itself induced IFN-λ3. These were further<br />

confirmed by in vitro experiments with colon cancer cell lines<br />

that ADV or TDF, not LAM or ETV, induced IFN-λ3 in a dose<br />

dependent manner with additive effects of IFN. In addition,<br />

immunohistochemistry of IFN-λ3 showed clear staining in the<br />

cytoplasm. No nucleos(t)ide analogs (NUC) induced IFN-λ3 in<br />

other cell lines originated from blood, liver, lung, and skin. In<br />

another setting, 83 HBV patients were prospectively switched<br />

to PEG-IFN treatment for 48 weeks after cessation of NUC<br />

(sequential therapy). Low HBsAg (< 3.0 Log IU/ml) and high<br />

serum IFN-λ3 (>19.0 pg/ml) before PEG-IFN were identified<br />

as significant factors associated with HBsAg reduction at the<br />

end of PEG-IFN by logistic regression analysis. CONCLUSIONS:<br />

This is a first report that nucleotide analogs (ADV or TDF), not<br />

nucleoside analogs (LAM or ETV), have an additional effect<br />

to induce IFN-λ3 as well as HBV polymerase inhibition. It is<br />

possible that IFN-λ3 induced in the gastrointestinal tracts by<br />

orally administered ADV or TDF reaches to the liver via portal<br />

tract and has immuno-modulatory effects against HBV. These<br />

results strongly indicate that prospective <strong>studies</strong> with additional<br />

PEG-IFN on ADV or TDF after achievement of 2 factors above<br />

would be warranted.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kazumoto Murata, Akihiro Matsumoto,<br />

Taisuke Inoue, Minoru Sakamoto, Masaya Sugiyama, Nobuyuki Enomoto,<br />

Eiji Tanaka, Masashi Mizokami


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1025A<br />

1674<br />

Long-term follow-up study of hepatitis delta virus quasispecies<br />

complexity<br />

Maria Homs 1,2 , Maria Buti 1,3 , Alicia Ruiz 4 , Pilar Reimundo 4 , Josep<br />

Gregori 5 , Maria Blasi 1,2 , Rosario Casillas 5 , David Tabernero 1,2 ,<br />

Marta Vila 5 , Leonardo Nieto 2 , Josep Quer 1,5 , Mar Riveiro-Barciela<br />

3 , Rafael Esteban 1,3 , Francisco Rodriguez-Frias 1,2 ; 1 Centro<br />

de Investigacion Biomedica en Red de Enfermedades Hepaticas y<br />

Digestivas, CIBERehd, Barcelona, Spain; 2 Microbiology, Hospital<br />

Vall d’Hebron, Barcelona, Spain; 3 Hepatology, Hospital Vall d’Hebron,<br />

Barcelona, Spain; 4 Biochemistry, Hospital Vall d’Hebron,<br />

Barcelona, Spain; 5 Liver Diseases, Research Institute of Vall d’Hebron,<br />

Barcelona, Spain<br />

Background The hepatitis delta virus (HDV) genome encodes<br />

two antigens essential for viral cycle, small and large (S- and<br />

L-HDAg). Genomes with a stop codon in position 196 encode<br />

S-HDAg, and with a tryptophan encode L-HDAg. Immune<br />

responses likely play a key role in controlling HDV infection;<br />

but changes occurring in the quasispecies (QS) complexity and<br />

in the percentage of genomes encoding S-HDAg and L-HDAg<br />

in the circulating HDV QS are unknown. Aim Evaluate changes<br />

in the complexity of HDV QS and in the percentage of genomes<br />

encoding S-HDAg and L-HDAg during long-term chronic HDV<br />

infection. Patient and methods Twelve consecutive serum samples<br />

from a patient with chronic HDV followed for 12 years<br />

were included. The patient was HDV genotype 1, HBeAg-negative,<br />

and HBV-DNA levels were


1026A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1676<br />

Noninvasive markers of liver fibrosis remain stable in<br />

the majority of hepatitis B virus and human immunodeficiency<br />

virus co-infected patients undergoing tenofovir-containing<br />

antiretroviral therapy<br />

Anders Boyd 1 , Karine Lacombe 1,2 , Pierre-Marie Girard 1,2 , Caroline<br />

Lascoux-combe 3 , Patrick Miailhes 4 , Hayette Rougier 2 , Lawrence<br />

Serfaty 2 ; 1 Inserm UMR_S1136, Paris, France; 2 Hôpital<br />

Saint-Antoine, Paris, France; 3 Hôpital Saint-Louis, Paris, France;<br />

4 Hôpital Croix-Rousse, Lyon, France<br />

Long-term tenofovir (TDF) use has been associated with significant<br />

regression of liver fibrosis during hepatitis B virus (HBV)<br />

monoinfection. However, there are few data in TDF-treated<br />

patients infected with both HBV and the human immunodeficiency<br />

virus (HIV), in whom HIV-associated immunosuppression<br />

could affect liver repair. In this prospective study, 168 HIV-HBV<br />

co-infected patients enrolled in the French HIV-HBV cohort were<br />

included if they initiated TDF-containing antiretroviral therapy<br />

and had liver fibrosis measurements at baseline and ≥1 during<br />

treatment. Patients with hepatitis C of D virus were not included.<br />

Fibrosis was assessed at baseline and every 6-12 months and<br />

≥F3 fibrosis determined from two noninvasive measures: biochemical<br />

score (≥0.59, FibroTest®) and/or transient elastography<br />

(≥7.6kPa, FibroScan). Determinants for fibrosis regression<br />

or progression were evaluated using Cox proportional hazards<br />

regression. Patients were predominately male (86.3%) with<br />

a median age of 41 years (IQR=36-48). At baseline, HBV-<br />

DNA was detectable in 134 (80.2%) patients with a median<br />

HBV-DNA viral load of 5.02 IU/mL (IQR=2.95-7.15) and 104<br />

(61.9%) had HBeAg-positive serology. After a median follow-up<br />

of 59.3 months (IQR=36.1-92.7), 140 (83.3%) patients<br />

were able to maintain or achieve undetectable HBV-DNA viral<br />

loads. Among patients with F3-F4 baseline liver fibrosis, 24/53<br />

(45.3%) remained at the same level of fibrosis during treatment<br />

and 29/53 (54.7%) regressed to F0-F1-F2 fibrosis after<br />

a median 16.9 months (IQR=8.6-32.8). Of the latter group,<br />

10/29 (34.5%) maintained F0-F1-F2 fibrosis until the end of<br />

follow-up. Liver fibrosis regression was significantly associated<br />

with CD4+ cell count ≥500/mm 3 (HR=2.09, 95%CI=1.13-<br />

3.88) in multivariable analysis. Among patients with F0-F1-F2<br />

baseline liver fibrosis, 70/115 (60.9%) remained at the same<br />

level of fibrosis and 45/115 (39.1%) progressed to F3-F4<br />

fibrosis after a median 29.0 months (IQR=18.0-47.4). Of the<br />

latter group, 28/45 (59.6%) maintained F3-F4 fibrosis. Liver<br />

fibrosis progression was significantly associated with higher<br />

age (HR/year=1.10, 95%CI=1.06-1.14), alanine aminotransferase<br />

>2X ULN (HR=2.10-9.28), and tended to be associated<br />

with an AIDS-defining event (HR=1.79, 0.98-3.27) in<br />

multivariable analysis. In conclusion, liver fibrosis levels are<br />

stable for the majority of HIV-HBV co-infected patients undergoing<br />

TDF-containing ART, despite extensive control of HBV<br />

replication. Immunosuppression plays a strong role in reducing<br />

fibrosis levels, supporting the need for earlier ART-initiation in<br />

this patient population.<br />

Disclosures:<br />

Lawrence Serfaty - Advisory Committees or Review Panels: MSD, Janssen, Roche,<br />

Gilead, BMS, Abbvie; Speaking and Teaching: Aptalis<br />

The following authors have nothing to disclose: Anders Boyd, Karine Lacombe,<br />

Pierre-Marie Girard, Caroline Lascoux-combe, Patrick Miailhes, Hayette Rougier<br />

1677<br />

Factors associated with decreases in estimated glomular<br />

filtration rates for patients co-infected with hepatitis B<br />

virus and human immunodeficiency virus undergoing<br />

tenofovir-containing antiretroviral therapy<br />

Anders Boyd 1 , Karine Lacombe 2,1 , Patrick Miailhes 3 , Caroline<br />

Lascoux-combe 4 , Hayette Rougier 2 , Julie Bottero 2,1 , Pierre-Marie<br />

Girard 2,1 ; 1 Inserm UMR_S1136, Paris, France; 2 Hôpital Saint-Antoine,<br />

Paris, France; 3 Hospices Civils de Lyon, Lyon, France; 4 Hôpital<br />

Saint-Louis, Paris, France<br />

Renal toxicity remains one of the more common tolerance<br />

issues encountered during prolonged treatment with tenofovir<br />

(TDF). Although determinants for renal impairment during TDF<br />

have been well established among patients infected with hepatitis<br />

B virus (HBV) or human immunodeficiency virus (HIV), little<br />

is known how these specific factors influence renal function in<br />

HIV-HBV co-infected patients. In this prospective study, 175<br />

HIV-HBV co-infected patients enrolled in the French HIV-HBV<br />

cohort were included if they initiated TDF-containing antiretroviral<br />

therapy (ART) and had creatinine levels available at<br />

baseline and ≥1 during treatment. Patients with hepatitis C<br />

or D virus were not included. Estimated glomular filtration<br />

rates (eGRF) were calculated using the CKD-EPI equation<br />

from creatinine values obtained at baseline and every 6-12<br />

months. Mixed-effect linear regression models were used to<br />

evaluate differences in eGRF changes from baseline (ΔeGFR)<br />

between certain risk-factors. Determinants towards impaired<br />

renal function (90 (p=0.002), males (p=0.04), an AIDS-defining illness<br />

(p=0.03), shorter duration of ART at baseline (p=0.01), concomitant<br />

therapy with a protease inhibitor (p=0.005), and<br />

baseline liver cirrhosis (p=0.03). Among patients with normal<br />

baseline renal function, 51/114 (44.7%) maintained eGFR<br />

levels >90 mL/min/1.73m 2 , while 63/114 (55.3%) developed<br />

either mild (60-89 mL/min/1.73m 2 , n=59) or moderate<br />

(30-59 mL/min/1.73m 2 , n=4) renal impairment after 15.0<br />

months (IQR=7.1-35.8). In multivariable analysis, higher age<br />

(HR/year=1.05, 95%CI=1.01-1.10) and AIDS-defining illness<br />

(HR=1.68, 95%CI=1.05-2.69) were associated with developing<br />

renal impairment, whereas undetectable HBV-DNA was<br />

significantly protective (HR=0.41, 95%CI=0.23-0.76). In conclusion,<br />

liver fibrosis and AIDS-defining illnesses negatively<br />

and independently affect renal function in HIV-HBV co-infected<br />

patients undergoing long-term TDF. Controlled HBV-infection<br />

prevents further progression towards renal impairment, stressing<br />

the need for effective therapy in this population.<br />

Disclosures:<br />

The following authors have nothing to disclose: Anders Boyd, Karine Lacombe,<br />

Patrick Miailhes, Caroline Lascoux-combe, Hayette Rougier, Julie Bottero,<br />

Pierre-Marie Girard


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1027A<br />

1678<br />

No Detectable Resistance to Tenofovir Disoproxil Fumarate<br />

(TDF) when Given Alone or in Combination with<br />

Emtricitabine (FTC) in Chronic Hepatitis B (CHB) Patients<br />

with Documented Resistance to Lamivudine (LAM): Final<br />

5 Year Results<br />

Thomas Berg 2 , Edward J. Gane 3 , Maciej S. Jablkowski 4 , Petr<br />

Urbanek 5 , Amoreena C. Corsa 1 , Yang Liu 1 , Kyungpil Kim 1 , John<br />

F. Flaherty 1 , Cihan Yurdaydin 6 , Scott Fung 7 , Kathryn M. Kitrinos 1 ;<br />

1 Gilead Sciences, Foster City, CA; 2 University Hospital Leipzig,<br />

Leipzig, Germany; 3 New Zealand Liver Transplant Unit, Auckland,<br />

New Zealand; 4 Medical University of Lodz, Lodz, Poland; 5 Central<br />

Military Hospital Prague, Prague, Czech Republic; 6 Ankara University<br />

Medical School, Ankara, Turkey; 7 Toronto General Hospital,<br />

Toronto, ON, Canada<br />

Aim: To compare amino acid changes within HBV polymerase/reverse<br />

transcriptase (pol/RT) after up to 5 years (240<br />

weeks) of treatment with TDF or the combination of FTC and<br />

TDF. Methods: Patients enrolled in the study were receiving<br />

LAM, harbored LAM resistance mutations in pol/RT (rtM204V/<br />

I±rtL180M) as determined by INNO-LiPA Multi-DR v2/v3, and<br />

were randomized 1:1 to receive TDF or FTC/TDF for up to<br />

240 weeks. A subset of enrolled patients received prior CHB<br />

treatment in addition to LAM: 13 received entecavir (ETV) and<br />

61 received adefovir (ADV). In addition, a subset of enrolled<br />

patients harbored ETV resistance mutations at baseline (TDF =<br />

14, FTC/TDF = 11). Virologic breakthrough (VB) was defined<br />

as confirmed HBV DNA >1 log 10<br />

increase from nadir or HBV<br />

DNA ≥69 IU/mL after


1028A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Yuji Ishida, Chihiro Yamasaki,<br />

Ami Yanagi, Yasumi Yoshizane, Kazuyuki Fujikawa, Koichi Watashi, Hiromi<br />

Abe, Takaji Wakita, Nelson Hayes, Chise Tateno<br />

1680<br />

Modeling Hepatits B Virus infection in stem cell derived<br />

hepatocytes reveals role of perturbations in NTCP genotype<br />

and expression in HBV permissiveness<br />

Angela Frankel, Robert E. Schwartz; Weill Cornell Medical College,<br />

New York, NY<br />

Worldwide, hepatitis B virus (HBV) infection is the most common<br />

viral hepatitis having infected over two billion people and<br />

chronically infecting more than 400 million, putting them at<br />

increased risk to develop cirrhosis and hepatocellular carcinoma.<br />

HBV research has been hampered by the virus’s narrow<br />

host range and cellular tropism for hepatocytes, which has led<br />

to a paucity of robust and reliable infectious systems for HBV<br />

study. Current model systems to study HBV include hepatoma<br />

cell lines which do not faithfully recapitulate adult hepatocyte<br />

phenotype or function as they have undergone a variety of<br />

genetic and metabolic changes. As a consequence, major components<br />

of the viral entry process and viral life cycle - including<br />

the establishment and persistence of a nuclear cccDNA pool<br />

- and many aspects of virus-host interactions have been poorly<br />

understood. We have recently shown that micropatterned<br />

co-cultures of primary human hepatocytes with stromal cells<br />

and inducible pluripotent stem cells differentiated into hepatocyte-like<br />

cells (iHeps) are permissive to HBV infection. iHeps<br />

become permissive only during late stages of differentiation,<br />

following the activation of the HBV transcription machinery,<br />

and the appearance of the recently discovered HBV receptor<br />

sodium-taurocholate co-transporter polypeptide (NTCP). Mutations<br />

of NTCP have been shown to severely impair bile acid<br />

uptake but the impact that particular variants play in HBV infection<br />

is largely unknown. In particular the NTCP Ser267Phe variant<br />

has recently been shown to be associated with resistance<br />

to the development of chronic HBV in a recently completed<br />

genome-wide association study. We hypothesized that mutations<br />

in NTCP may be directly responsible for this association.<br />

A gene editing approach using CrispR (and guide RNA’s targeting<br />

the NTCP locus) along with homology directed repair<br />

enabled the production of induced pluripotent stem cells (iPS)<br />

with variant mutations in NTCP in the genomic NTCP locus.<br />

Wild-type and variant NTCP iPS were differentiated into iHeps<br />

and then inoculated with HBV containing serum. NTCP variants<br />

had variable permissiveness for HBV infection. In particular the<br />

Ser267Phe mutation impacts NTCP expression and HBV permissiveness<br />

to HBV infection. These results exemplify the utility<br />

of a physiologically relevant infectious system for studying HBV<br />

interactions with relevant host cell genetics and physiology.<br />

Combining these systems with gene editing technologies will<br />

enable the dissection of clinically relevant mutations in physiologically<br />

relevant human infectious systems.<br />

Disclosures:<br />

The following authors have nothing to disclose: Angela Frankel, Robert E.<br />

Schwartz<br />

1681<br />

Genetic association study failed to confirm an association<br />

between the NTCP S267F mutation and persistence<br />

of hepatitis B virus or development of hepatocellular<br />

carcinoma in the Japanese population<br />

Daiki Miki 1,2 , Hidenori Ochi 1,2 , C. Nelson Hayes 1,2 , Hiromi Abe 1,2 ,<br />

Sakura Akamatsu 1,2 , Atsushi Ono 1,2 , Takashi Nakahara 1,2 , Yizhou<br />

Zhang 1,2 , Keiichi Masaki 1,2 , Hatsue Fujino 1,2 , Eisuke Miyaki 1,2 ,<br />

Hiromi Kan 1,2 , Takuro Uchida 1,2 , Nobuhiko Hiraga 1,2 , Masataka<br />

Tsuge 1,2 , Tomokazu Kawaoka 1,2 , Michio Imamura 1,2 , Yoshiiku<br />

Kawakami 1,2 , Hiroshi Aikata 1,2 , Kazuaki Chayama 1,2 ; 1 Laboratory<br />

for Digestive Diseases, RIKEN Center for Integrative Medical<br />

Sciences, Hiroshima, Japan; 2 Department of Gastroenterology and<br />

Metabolism, Hiroshima University Hospital, Hiroshima, Japan<br />

Background & Aims: Sodium taurocholate cotransporting<br />

polypeptide (NTCP) has long been known as a liver bile acid<br />

transporter but was also recently identified as a hepatocyte<br />

receptor for hepatitis B virus (HBV). The NTCP S267F mutation<br />

has been reported to reduce taurocholate transporting activity<br />

and result in defective HBV receptor function. This amino acid<br />

substitution corresponds to a single nucleotide polymorphism<br />

(SNP), rs2296651 is found in the Asian population but not in<br />

European, African and American populations, according to the<br />

1000 Genomes database. Moreover, this SNP frequency varies<br />

widely even among Asian populations: 12% (Chinese Dai),<br />

11% (Kinh (Vietnam)), 8% (Southern Han Chinese), 3% (Han<br />

Chinese), and 2% (Japanese). In this study, we investigated<br />

the importance of NTCP in HBV persistence and the development<br />

of HBV-induced hepatocellular carcinoma (HCC) by<br />

focusing on the S267F mutation using a large scale case-control<br />

association analysis. Methods: We genotyped SNPs in<br />

2,687 Japanese patients with chronic HBV infection (631 with<br />

HCC and 2,056 without HCC) and 5,489 control individuals<br />

without liver disease or cancer. Results: We found that minor<br />

allele (T) frequencies of rs2296651 were 0.013 and 0.010 in<br />

controls and chronic HBV patients, respectively. We observed<br />

3 minor allele homozygotes (TT genotype) in controls. In contrast,<br />

we did not observe any HBV patients with TT genotype.<br />

However, we found no significant differences in rs2296651<br />

frequencies between chronic HBV patients and controls in any<br />

genetic models. We also compared chronic HBV patients with<br />

and without HCC and found that their T allele frequencies were<br />

0.007 and 0.011, respectively. Finally, we found no significant<br />

association between the SNP genotype and development<br />

of HCC with or without adjustment for age and gender. These<br />

results might be partially affected by a low allele frequency<br />

of rs2296651. Therefore, we examined another NTCP SNP,<br />

rs4646287, which was reported as an intronic SNP associated<br />

with HBV-induced HCC susceptibility in the Han Chinese<br />

population and is more frequent in Japanese population than<br />

rs2296651. Previously reported risk allele (A) frequencies of<br />

rs4646287 were 0.144 and 0.142 in Japanese HBV patients<br />

with and without HCC, respectively, and we again found no<br />

significant association between SNP genotype and development<br />

of HCC. Conclusions: We found no genetic association<br />

between the S267F mutation of NTCP and HBV persistence or<br />

the development of HBV-induced HCC in the Japanese population.<br />

Further genetic analysis in populations with a higher<br />

frequency of rs2296651 would improve understanding about<br />

the role of NTCP in chronic HBV infection.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1029A<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Daiki Miki, Hidenori Ochi, C.<br />

Nelson Hayes, Hiromi Abe, Sakura Akamatsu, Atsushi Ono, Takashi Nakahara,<br />

Yizhou Zhang, Keiichi Masaki, Hatsue Fujino, Eisuke Miyaki, Hiromi Kan, Takuro<br />

Uchida, Nobuhiko Hiraga, Masataka Tsuge, Tomokazu Kawaoka, Michio<br />

Imamura, Yoshiiku Kawakami, Hiroshi Aikata<br />

1682<br />

Serial Changes of Th1 and Th17 Cytokines and a SNP<br />

in HLA class II DPB1 Gene Associated with Hepatitis B<br />

Virus Reactivation in Patients Treated with Immunomodulatory<br />

Agents<br />

Hidetaka Matsuda, Tomoyuki Nemoto, Yoshihiko Ozaki, Tatsushi<br />

Naito, Masahiro Ohtani, Katsushi Hiramatsu, Hiroyuki Suto,<br />

Yasunari Nakamoto; Second Department of Internal Medicine,<br />

University of Fukui, Fukui, Japan<br />

BACKGROUND: Hepatitis B virus (HBV) reactivation can be<br />

triggered by various chemotherapies, in which the patients’<br />

immunological status may be suppressive. As the patients’<br />

immunity are influenced by human leukocyte antigen (HLA)<br />

class II mediated interactions between CD4 + helper T cells and<br />

antigen presenting cells, polymorphisms of HLA class II genes<br />

might affect the reactivation of HBV. In this study, we aimed to<br />

assess the cytokine levels reflecting immunosuppressive status<br />

and the candidate single nucleotide polymorphisms (SNPs) of<br />

HLA class II genes associated with the risks of HBV reactivation<br />

in patients treated with immunomodulatory therapies. METH-<br />

ODS: STUDY 1: Five patients with malignant lymphoma (ML)<br />

who undergone R-CHOP or CHOP and achieved complete<br />

remission (CR) in University of Fukui Hospital between 2013<br />

and 2014 were enrolled. The sera collected consecutively from<br />

the subjects were measured for a total of 27 cytokines, chemokines<br />

and growth factors by multiplex cytokine analysis. Immune<br />

profiles after the initiation of chemotherapies were investigated,<br />

and levels of the 27 factors were compared with those<br />

of 20 healthy controls. STUDY 2: A total of 43 patients with<br />

resolved HBV infection treated with immunosuppressive chemotherapies<br />

(e.g. R-CHOP, CHOP, methotrexate, infliximab, and<br />

cyclosporin) during 1999 to 2014 were investigated. All were<br />

followed up for more than 2 years; HBV had reactivated in 10<br />

of the 43 subjects, but not in the other 33. Genotyping of 22<br />

SNPs located within HLA class II DPA1, DPB1, DQA1, DQB1,<br />

and DRB1 genes was conducted by quenching probe PCR<br />

(geneCube). Genotype frequencies of the tested SNPs were<br />

compared between the two groups, whose latently infected<br />

HBV had reactivated and not. RESULTS: STUDY 1: In the tested<br />

patients with ML, serial levels of IL-17, IL-1β, IFN-γ, and G-CSF<br />

at the timing of CR with R-CHOP and CHOP were significantly<br />

decreased in contrast to those of the controls (123.1, 3.7,<br />

96.0, and 194.1 pg/ml vs. 158.2, 6.8, 146.2, and 306.5<br />

pg/ml, respectively; p 22.5% was<br />

significant higher than those with PD-1 expression increasing<br />

≤ 22.5% (46.15% vs 13.64%, P=0.033) from week 12 to<br />

week 24. Interestingly, the patients with both TLR2 expression<br />

< 18% at week 12 and PD-1 expression increasing > 22.5%<br />

from week 12 to week 24 had a significantly higher rate of<br />

virological breakthrough than the patients who met either of<br />

conditions mentioned above (83.33% vs. 9.09%, P=0.002),<br />

and the patients who failed to meet any condition (83.33%<br />

vs. 11.11%, P=0.005). Conclusions Downregulation of TLR2<br />

on CD14 + monocytes and upregulation of PD-1 on CD8 + T cells<br />

correlate with virological breakthrough during the treatment<br />

with Peg IFN alfa-2a after switching them from ETV therapy.<br />

Combination of TLR2 expression < 18% at week 12 and PD-1<br />

expression increasing > 22.5% from week 12 to week 24 has<br />

a high positive predictive value (PPV) for virological breakthrough<br />

in entecavir suppressed CHB patients with sequential<br />

Peg IFN alfa treatment.


1030A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Qin Ning - Advisory Committees or Review Panels: ROCHE, NOVARTIS, BMS,<br />

MSD, GSK; Consulting: ROCHE, NOVARTIS, BMS, MSD, GSK; Grant/Research<br />

Support: ROCHE, NOVARTIS, BMS; Speaking and Teaching: ROCHE, NOVAR-<br />

TIS, BMS, MSD, GSK<br />

The following authors have nothing to disclose: Di Wu, Weiming Yan, Guang<br />

Chen, Meifang Han<br />

1684<br />

IFN-α-mediated Base Excision Repair Pathway Correlates<br />

with Antiviral Response Against Hepatitis B Virus<br />

Infection<br />

Yong Li 1 , Yuchen Xia 3 , Meifang Han 2 , Guang Chen 2 , Wolfgang<br />

E. Thasler 4 , Xiaoping Luo 1 , Ulrike Protzer 3 , Qin Ning 2 ; 1 Tongji<br />

Hospital, Tongji Medical College, Huazhong University of Science<br />

and Technology, Wuhan, China; 2 Department and Institute<br />

of Infectious Diseases, Tongji Hospital, Tongji Medical College,<br />

Huazhong University of Science and Technology, Wuhan, China;<br />

3 Institute of Virology, TechnischeUniversität München–Helmholtz<br />

Zentrum München, Munich, Germany; 4 Department of General,<br />

Visceral, Ludwig Maximilians University, Munich, Germany<br />

Previous <strong>studies</strong> identified APOBEC family as base excision<br />

repair (BER) proteins with enzymatic activity through deamination<br />

of a cytidine base in DNA and/or RNA. As an<br />

enzyme, APOBEC is therefore part of an antiviral effector repertoire<br />

against hepatitis B virus (HBV) infection by interferon<br />

(IFN). Additionally, other BERs, namely NEIL3 and TDG, are<br />

expressed in liver and are regulated by IFN-α in hepatocytes.<br />

We thus hypothesized that the responses to IFN-α treatment of<br />

chronic hepatitis B (CHB) patients are relevant to IFN-induced<br />

deaminases and BER genes. Ten CHB patients treated with<br />

PEGylated IFN-α for 48 or 96 weeks, and six CHB-treatment<br />

naïve patients as controls were recruited. Response to IFN<br />

treatment was defined as HBV DNA 1 log10IU/<br />

ml. Blood and liver samples were collected, and APOBEC3<br />

and other BER genes measured by real-time PCR. The correlations<br />

between BER gene expression levels and IFN treatment<br />

responses were studied in these patients as well as in primary<br />

human hepatocytes (PHH) and terminally differentiated HepRG<br />

cells in vitro. Compared to treatment-naive patients, APO-<br />

BEC3-A,-B, -C, -DE, and -G mRNA levels were up-regulated in<br />

IFN-treated subjects. APOBEC3-A was significantly increased<br />

in IFN-treated responders than non-responders. In contrast,<br />

other BER genes, NEIL3 and TDG, were down-regulated in both<br />

IFN-treated patients and in IFN-treated PHH and HepRG cells.<br />

APOBEC3 and BER gene expression at treatment endpoints<br />

partially correlated with the corresponding degree of HBsAg/<br />

HBV DNA decline and DNA level. Our study suggests that the<br />

expression levels of editing enzymes APOBEC3-A, -C, -F, -G,<br />

and NEIL3 and TDG correlates with IFN treatment responses in<br />

CHB patients, which may serve as biomarkers for CHB disease<br />

management.<br />

Disclosures:<br />

Ulrike Protzer - Advisory Committees or Review Panels: GILEAD, Roche, MedImmune;<br />

Board Membership: University Hospital Cologne; Grant/Research Support:<br />

Janssen, Roche<br />

Qin Ning - Advisory Committees or Review Panels: ROCHE, NOVARTIS, BMS,<br />

MSD, GSK; Consulting: ROCHE, NOVARTIS, BMS, MSD, GSK; Grant/Research<br />

Support: ROCHE, NOVARTIS, BMS; Speaking and Teaching: ROCHE, NOVAR-<br />

TIS, BMS, MSD, GSK<br />

The following authors have nothing to disclose: Yong Li, Yuchen Xia, Meifang<br />

Han, Guang Chen, Wolfgang E. Thasler, Xiaoping Luo<br />

1685<br />

Hepatitis B virus down regulates the expression levels of<br />

Vitamin D receptor in hepatocellular carcinoma (HepG2)<br />

transfected cells thereby preventing the inhibition of<br />

viral transcription and production by Vitamin D<br />

Neta Gotlieb 1 , Irina Tachlytski 1 , Maya Sultan 1 , Michal Safran 1 ,<br />

Ziv Ben Ari 1,2 ; 1 Liver Disease Center, Sheba Medical Center, Tel<br />

Hashomer, Israel; 2 The Sackler School of Medicine, Tel Aviv University,<br />

Tel-Aviv, Israel<br />

Vitamin D is an important immune modulator that plays an<br />

emerging role in both the innate and adaptive immune systems.<br />

Certain viruses as the HIV-1 are capable to impair the innate<br />

immune defense by down regulating the vitamin D receptor<br />

(VDR) pathway. Recently it has been demonstrated that low<br />

25(OH)D3 serum levels are associated with high levels of hepatitis<br />

B Virus (HBV) replication in patients with HBV infection.<br />

Our aim was to study in vitro the relationship between HBV<br />

transcription and production and Vitamin D signaling pathway<br />

and to explore the associated mechanism(s). Methods: To measure<br />

HBV transcription and production we used qRT-PCR on<br />

RNA purified from HBV-transfected hepatocellular carcinoma<br />

HepG2 cells (HepG2 215) using primers specific to RNA-<br />

HBV. In addition, the level of HBV replication was evaluated<br />

by determining the concentrations of HBsAg levels in the cell<br />

culture medium of HepG2 215 using specific enzyme-linked<br />

immunosorbent (Elisa) assay. cccDNA levels were quantified<br />

using qRT-PCR on DNA extracted from HepG2 215 cells. The<br />

level of HBV transcription and production was assessed with<br />

and without the administration of vitamin D (0-200mM) or its<br />

active metabolite calcitriol (0-100 nM) for 48 or 72 hours. VDR<br />

and IFNβ transcripts expression level were assessed by qRT-<br />

PCR. Levels were compared in RNAs purified from HepG2 cells<br />

(control) versus HepG2 215. Results: The expression level of<br />

VDR transcript in HepG2 215 cells was statistically much lower<br />

compared with non-transfected HepG2 cells. Furthermore, the<br />

administration of vitamin D or calcitriol did not suppress the<br />

secretion of HBsAg, the cccDNA expression and the HBVRNAs<br />

level in HepG2 215 compared to cells which were not treated<br />

with Vitamin D. Increasing the dose of both vitamin D and<br />

calcitriol did not affect these findings. In addition, the administration<br />

of Vitamin D or calcitriol induced interferon signaling<br />

pathway, resulting in up-regulation of IFNβ expression level in<br />

HepG2 cells, however, no such increase was observed in the<br />

HepG2 215 HBV transfected cells. Conclusions: Our results<br />

indicate for the first time that VDR expression level is down-regulated<br />

in HBV-transfected HepG2 cells thereby preventing vitamin<br />

D to affect transcription and production of the HBV in<br />

vitro. Furthermore, these findings suggest that HBV might use<br />

this mechanism to avoid the immunological defense system by<br />

affecting the interferon signaling pathway and the expression<br />

of the IFNβ.<br />

Disclosures:<br />

The following authors have nothing to disclose: Neta Gotlieb, Irina Tachlytski,<br />

Maya Sultan, Michal Safran, Ziv Ben Ari<br />

1686<br />

The role of HBV quasispecies evolution in HBeAg-negative<br />

reactivation<br />

Guan Huei Lee 1,2 , Hui Heng Chong 2 , Seng Gee Lim 2 ; 1 Medicine,<br />

National University Health System, Singapore, Singapore; 2 Medicine,<br />

Yong Loo Lin School of Medicine, Singapore, Singapore<br />

Background: HBeAg negative chronic hepatitis B reactivation is<br />

characterized by increased levels of HBV DNA accompanied<br />

by abnormal ALT, and occasionally severe hepatitis B flare.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1031A<br />

We have shown previously that viral quasispecies evolution<br />

and diversity may play a pathogenic role in development of<br />

HBeAg seroconversion. However, the role of viral quasispecies<br />

in pathophysiology of reactivation in HBeAg-negative chronic<br />

hepatitis B is uncertain. Methods: Twenty-two HBeAg-negative<br />

chronic hepatitis B patients who had serial serum samples<br />

collected before and after HBV reactivation were analysed.<br />

Full length HBV genomes were amplified from serum extracted<br />

DNA and were cloned into pCR®2.1-TOPO vector. At least<br />

20 positive clones were randomly selected and sequenced<br />

using Big Dye Terminator. The samples were analyzed for point<br />

mutation in precore stop codon, core promoter- Enhancer II, S1<br />

promoter, S2 promoter and Enhancer I- X promoter regions.<br />

Ten inactive HBV carriers were also analysed as controls. Fulllength<br />

HBV genome collected before and after reactivation<br />

were transfected into Huh7 cells for functional analysis of their<br />

replication fitness. Results: In subjects with clear clinical reactivation<br />

profile, clonal selection was demonstrated (9 out of<br />

10 subjects) by DNA phylogenetic analysis (p=0.005). The<br />

genetic diversity and viral evolution rates were not statistically<br />

different from the inactive controls except except at the precore/core<br />

ORF (p=0.035, Mann-Whitney U test). 80% of reactivation<br />

patients had precore stop codon G to A substitution at<br />

nucleotide position 1896. In addition, other mutations in the<br />

promoter/enhancer regions were identified with more than fifty<br />

percent difference between time point 1 and time point 2 (and<br />

not found in control subjects). HBV clones isolated during HBV<br />

reactivation showed higher HBV DNA and HBsAg production<br />

rates than the clones isolated during inactive phase of the same<br />

patient in transfected Huh7 cell line. Conclusions: Significant<br />

clonal selection occurred during HBeAg-negative reactivation.<br />

Precore stop codon substitution G1896A occurs significantly<br />

in majority of the patients. The nucleotide variability in the precore<br />

promoter, and possibly other promoter/enhancer regions<br />

may play an important role in the progression of HBV disease,<br />

although more functional analyses of individual mutations are<br />

required.<br />

Disclosures:<br />

Seng Gee Lim - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals, Gilead<br />

Pharmaceuticals, Roche Pharmaceuticals; Speaking and Teaching: Bristol-Myers<br />

Squibb, Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals,<br />

Gilead Pharmaceuticals<br />

The following authors have nothing to disclose: Guan Huei Lee, Hui Heng Chong<br />

1687<br />

Hepatic macrophage IL-1 beta expression and bile and<br />

sterol transporter suppression precede cholestasis in a<br />

parenteral nutrition mouse model<br />

Karim C. El Kasmi, Aimee Anderson, Michael W. Devereaux,<br />

Ronald J. Sokol; Pediatrics, UC Denver, Aurora, CO<br />

Background: Parenteral Nutrition Associated Cholestasis<br />

(PNAC) is a serious complication of long term PN infusion in<br />

children and adults with intestinal failure. We have previously<br />

reported in a PNAC mouse model that activation of hepatic<br />

macrophages and up-regulation of pro-inflammatory cytokines<br />

(IL-1 beta) was associated with suppression of hepatic bile<br />

(Abcb11/BSEP, Abcc2/MRP2) and sterol (ABCG5/8/sterolin)<br />

transporters. IL-1R-/- mice were protected from PNAC, with<br />

normalization of hepatic transporter expression, supporting<br />

a causal role for IL-1 beta. The objective of this study was<br />

to determine the role of IL-1 beta at early stages of PNAC<br />

by examining the temporal relationship at early time points<br />

between IL-1 beta, transporter expression, and cholestasis.<br />

Methods and Results: PNAC model: wild type C57/B6 mice<br />

were exposed to dextran sulfate sodium (DSS) in drinking<br />

water (to induce intestinal injury) for 4 days followed by infusion<br />

of soy lipid emulsion-based PN through a central venous<br />

catheter for 3 or 7 days (DSS-PN mice); appropriate controls<br />

were conducted. Cholestasis (increased serum bile acids and<br />

bilirubin) and hepatocyte injury (increased AST and ALT)<br />

were absent in DSS/PN3d mice but developed in all DSS/<br />

PN7d mice. IL-1 beta transcription was significantly induced<br />

in liver homogenate and in isolated hepatic mononuclear cells<br />

(hMNCs) at 3 days in DSS/PN mice and persisted at 7 days.<br />

Concomitantly, transcription of Abcb11, Abcc2, and Abcg5/8<br />

was suppressed at both of these time points in DSS/PN mice,<br />

indicating that macrophage activation and transporter suppression<br />

preceded the onset of biochemical cholestasis. To further<br />

determine if IL-1 beta was responsible for the transcriptional<br />

suppression, wild type mice were injected i.p. with recombinant<br />

IL-1 beta (200ng/mouse) and sampled after 4 hrs, and<br />

HuH7 and HepG2 cells (human hepatocyte cell lines) were<br />

incubated with IL-1b (10-50ng/ml) for 4 hrs, and gene expression<br />

measured. IL-1 beta treatment significantly suppressed<br />

expression of Abcb11, Abcc2, and Abcg5/8 in the mice and<br />

in both hepatocyte cell lines. Incubation of these cell lines with<br />

LPS (100-1000ng/ml) had no effect on these transporters. Conclusions:<br />

Activation of hepatic macrophages and generation of<br />

IL-1 beta, as well as suppression of bile and sterol transporters,<br />

occurred after only 3 days of PN infusion and preceded onset<br />

of cholestasis. Furthermore, IL-1 beta was sufficient to suppress<br />

these transporters in cultured hepatocytes and in vivo. These<br />

data support IL-1 beta as a critical early mediator in the pathogenesis<br />

of PNAC and suggest IL1-beta signaling as a potential<br />

therapeutic target in this disorder.<br />

Disclosures:<br />

Ronald J. Sokol - Advisory Committees or Review Panels: Yasoo Health, Inc.; Consulting:<br />

Roche, Ikaria, Otsuka American Pharmaceuticals, Alnylam, Retrophin;<br />

Grant/Research Support: Mead Johnson Nutritionals, Lumena, FFF Enterprises<br />

The following authors have nothing to disclose: Karim C. El Kasmi, Aimee Anderson,<br />

Michael W. Devereaux<br />

1688<br />

Molecular mechanism of IL-1 beta mediated transcriptional<br />

suppression of the sterol transporter Abcg5/8 in<br />

a parenteral nutrition mouse model<br />

Karim C. El Kasmi, Aimee Anderson, Padade M. Vue, Michael<br />

W. Devereaux, Natarajan Balasubramaniyan, Frederick J. Suchy,<br />

Ronald J. Sokol; Pediatrics, UC Denver, Aurora, CO<br />

Background: Parenteral nutrition associated cholestasis (PNAC)<br />

is the leading indication for combined intestine/liver transplant.<br />

A possible role of intravenous soy lipid emulsions in PNAC<br />

pathogenesis has led to concern about the hepatic toxicity of<br />

phytosterols contained in soy lipid emulsions. We have demonstrated<br />

in a novel PNAC mouse model that soy lipid emulsions<br />

in PN are associated with induction of hepatic macrophage<br />

IL1b expression concomitant with accumulation of hepatic phytosterols<br />

and transcriptional suppression of hepatocyte Abcg5/<br />

g8, coding for the canalicular transporter of sterols, as well<br />

as BSEP and MRP2. The aim of this study was to elucidate the<br />

role of IL1b signaling in suppression of ABCG5/8 transcription<br />

during PNAC. Methods and Results: Wild type C57/B6 mice<br />

were exposed to dextran sulfate sodium (DSS) (to induce intestinal<br />

injury) for 4d followed by infusion of phytosterol-containing<br />

(soy lipid) PN solution through a central venous catheter for 14<br />

d (DSS-PN mice) and developed cholestasis (increased serum<br />

bile acids and bilirubin) and hepatocyte injury (increased AST<br />

and ALT). DSS-PN mice had hepatic IL1b induction and markedly<br />

reduced hepatic mRNA for Abcb11, Abcc2, Abcg5/8<br />

compared to control mice. To further elucidate the role of IL1b<br />

in this transcriptional suppression, mice with genetic deletion


1032A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

of the receptor for IL-1 (IL1R-/-) were exposed to DSS-PN for<br />

14d. DSS-PN/IL1R-/- mice had significantly reduced serum<br />

AST, ALT, bile acids, and bilirubin compared to DSS-PN mice.<br />

Hepatic gene mRNA for Abcb11, Abcc2, and Abcg5/8 was<br />

not reduced in DSS-PN IL1R-/- mice. Chromatin-immuno-precipitation<br />

assays (ChIP) on liver demonstrated that in WT DSS-PN<br />

mice binding of NFkB (downstream of IL1 signaling) to the<br />

Abcg5/8 promoter was increased concomitant with reduced<br />

binding of LXR (the transcription factor that promotes Abcg5/8<br />

expression). In contrast, in DSS-PN IL1R-/- mice, binding of<br />

NFkB was reduced concomitant with increased binding of LXR<br />

to the Abcg5/8 promoter. Supporting these in vivo <strong>studies</strong>,<br />

incubation of HepG2 and HuH7 cells with IL1b significantly<br />

suppressed Abcg5/8 mRNA while markedly increasing NFkB<br />

reporter activity. Finally, incubation of HepG2 cells with IL1b<br />

suppressed the ability of an LXR agonist to induce activity<br />

ABCG5/8 promoter linked luciferase. Conclusion: Hepatic IL1b<br />

is a critical mediator in the pathogenesis of PNAC by promoting<br />

suppression of the phytosterol transporter Abcg5/8 through<br />

increased binding of NFkB and decreased binding of LXR to<br />

the Abcg5/8 promotor. Reduced Abcg5/8 causes accumulation<br />

of phytosterols which then interfere with hepatocyte FXR<br />

signaling, thus promoting cholestasis.<br />

Disclosures:<br />

Ronald J. Sokol - Advisory Committees or Review Panels: Yasoo Health, Inc.; Consulting:<br />

Roche, Ikaria, Otsuka American Pharmaceuticals, Alnylam, Retrophin;<br />

Grant/Research Support: Mead Johnson Nutritionals, Lumena, FFF Enterprises<br />

The following authors have nothing to disclose: Karim C. El Kasmi, Aimee Anderson,<br />

Padade M. Vue, Michael W. Devereaux, Natarajan Balasubramaniyan,<br />

Frederick J. Suchy<br />

1689<br />

Enteral Obeticholic Acid Promotes Intestinal Growth in<br />

Total Parenteral Nutrition Fed Neonatal Pigs<br />

Yanjun Jiang 1 , Zhengfeng Fang 2 , Barbara Stoll 1 , Gregory J. Guthrie<br />

1 , Jens Holst 3 , Bolette Hartmann 3 , Douglas Burrin 1,4 ; 1 USDA<br />

Children’s Nutrition Research Center, Department Pediatrics, Baylor<br />

College of Medicine, Houston, TX; 2 Animal Nutrition Institute,<br />

Sichuan Agricultural University, Chengdu, China; 3 The NNF Center<br />

for Basic Metabolic Research and the Department of Biomedical<br />

Sciences, University of Copenhagen, Copenhagen, Denmark;<br />

4 Pediatric Gastroenterology, Hepatology and Nutrition, Department<br />

Pediatrics, Baylor College of Medicine, Houston, TX<br />

Intestinal atrophy is an adverse outcome associated with prolonged<br />

total parenteral nutrition (PN) partly due to disruption<br />

of normal enterohepatic circulation of bile acids. Previously<br />

we showed that enteral treatment with chenodeoxycholic acid<br />

(CDCA), a dual agonist for the nuclear receptor, farnesoid X<br />

receptor (FXR) and G protein-coupled receptor TGR-5, induced<br />

intestinal mucosal growth in a parenteral nutrition associated<br />

liver disease (PNALD) piglet model. We hypothesized that the<br />

intestinal trophic CDCA effects were mainly mediated by TGR5<br />

receptor-mediated glucagon-like peptide 2 (GLP-2) released<br />

from enteroendocrine cells. However, CDCA may also exert<br />

trophic effects via FXR signaling in intestine. The aim of the current<br />

study was to compare the physiological effects of a selective<br />

and potent FXR agonist, obeticholic acid (OCA) vs CDCA<br />

on intestinal growth in TPN-fed pigs. Term, newborn pigs were<br />

assigned to receive complete TPN (PN), PN + enteral CDCA<br />

(30 mg/kg), or PN +enteral OCA (0.5, 5, 15 mg/kg) daily for<br />

19 d. The daily parenteral lipid was Intralipid given at 10 g/<br />

kg. Intestinal growth and crypt cell proliferation (in vivo BrdU<br />

labeling) were measured. We found that both CDCA and OCA<br />

treatments significantly increased small intestinal weight, compared<br />

to PN pigs, but OCA15 was higher than CDCA (146%<br />

vs. 118%). The OCA-induced increase in jejunal and Ileal<br />

weight was dose-dependent, yet the trophic effects of OCA and<br />

CDCA were greater in the ileum than jejunum. Ileal villus height<br />

and crypt depth were increased by OCA and CDCA. The percentage<br />

of BrdU positive crypt cells in PN, CDCA, OCA0.5,<br />

OCA5 and OCA15 groups were 31%, 40%, 40%, 46% and<br />

46% respectively, suggesting OCA and CDCA increased crypt<br />

cell proliferation. Portal plasma GLP-1 and -2 concentrations<br />

were significantly increased in pigs treated with CDCA, but not<br />

OCA, compared to PN pigs suggesting differential intestinal<br />

activation of TGR5 signaling. OCA, but not CDCA, treatment<br />

dose-dependently increased ileal transcriptional expression of<br />

FXR target genes, small heterodimer partner (SHP), ileal lipid<br />

binding protein (ILBP), and fibroblast growth factor 19 (FGF19).<br />

The expression of fibroblast growth factor receptor 4 (FGFR4)<br />

and β-Klotho mRNA were relative abundant in ileal tissue in all<br />

groups. We conclude that the intestinal trophic effects of OCA<br />

are greater than CDCA in TPN-fed pigs. The trophic effects of<br />

CDCA appear to occur via TGR5-mediated GLP-2 secretion<br />

rather than FXR signaling. We show novel evidence that OCA<br />

exerts trophic effects in the neonatal intestine and this appears<br />

to act mainly via FXR-dependent mechanisms.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yanjun Jiang, Zhengfeng Fang,<br />

Barbara Stoll, Gregory J. Guthrie, Jens Holst, Bolette Hartmann, Douglas Burrin<br />

1690<br />

Human iPSC-derived hepatocyte-like cells generated<br />

from patients with bile salt export pump deficiency<br />

recapitulate the phenotype of progressive familial intrahepatic<br />

cholestasis type 2<br />

Kazuo Imagawa 1,2 , Kazuo Takayama 2,3 , Shigemi Isoyama 1 , Ken<br />

Tanikawa 4 , Masato Shinkai 5 , Masayoshi Kage 4 , Kenji Kawabata<br />

6,7 , Ryo Sumazaki 1 , Hiroyuki Mizuguchi 2,3 ; 1 Department of<br />

Child Health, Faculty of Medicine, University of Tsukuba, Tsukuba,<br />

Japan; 2 Laboratory of Hepatocyte Regulation, National Institute of<br />

Biomedical Innovation, Osaka, Japan; 3 Laboratory of Biochemistry<br />

and Molecular Biology, Graduate School of Pharmaceutical<br />

Sciences, Osaka University, Osaka, Japan; 4 Department of<br />

Diagnostic Pathology, Kurume University Hospital, Kurume, Japan;<br />

5 Department of Surgery, Kanagawa Children’s Medical Center,<br />

Yokohama, Japan; 6 Laboratory of Stem Cell Regulation, National<br />

Institute of Biomedical Innovation, Osaka, Japan; 7 Laboratory of<br />

Biomedical Innovation, Graduate School of Pharmaceutical Sciences,<br />

Osaka, Japan<br />

Background: The bile salt export pump (BSEP) is a key molecule<br />

that transports bile acid into the biliary canaliculi in<br />

humans. Progressive familial intrahepatic cholestasis type 2<br />

(PFIC2) is mainly characterized by BSEP deficiency, which is<br />

followed by intrahepatic cholestasis. In many cases, liver transplantation<br />

is the most effective therapy available. Hence, it is<br />

necessary to identify novel therapeutic alternatives. To clarify<br />

the underlying disease mechanisms and discover the novel<br />

therapeutic options, we generated iPSCs from BSEP deficient<br />

(BD) patients and analyzed the differentiated hepatocyte-like<br />

cells (HLCs). Methods: One healthy donor and two BD patients<br />

participated in present study. Patient 1 was diagnosed with<br />

normal-gamma glutamyl transferase PFIC and presented with<br />

the PFIC2 phenotype in infancy. Although her clinical features,<br />

liver histology results, and BSEP deficiency were indicative<br />

of the PFIC2 phenotype, exonic mutations were not found in<br />

the BSEP gene. Patient 2 was diagnosed with PFIC2 based<br />

on the presence of compound heterozygous mutations in the<br />

BSEP gene (c.-24C>A and c.2416G>A) and liver histology<br />

results. The iPSCs were generated from blood cells obtained<br />

from all three participants (Control-iPSC and BD-iPSC) by using


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1033A<br />

a Sendai virus vector expressing Yamanaka factors. To examine<br />

whether these iPSCs expressed pluripotent markers, realtime<br />

RT-PCR and immunostaining analyses were performed.<br />

The iPSCs were then differentiated into hepatocyte-like cells.<br />

To examine the BSEP expression levels in HLCs derived from<br />

BD-iPSC (BD-iPS-HLCs) and their biliary excretion capacity, we<br />

performed real-time RT-PCR, immunostaining analyses, and<br />

cholyl-lysyl-fluorescein (CLF) excretion experiments. Results: We<br />

successfully obtained iPSCs from blood cells. These cells were<br />

positive for OCT4 and expressed markers specific to human<br />

embryonal stem cells. The HLCs were obtained on day 25 after<br />

hepatocyte differentiation. Bile canaliculi between the cells<br />

were observed by electron microscopy. BSEP was expressed<br />

in the membranes of Control-iPSC derived HLCs (Control-iPS-<br />

HLCs), while BD-iPS-HLCs did not show localized expression of<br />

BSEP in its membranes. The accumulation of CLF in bile canaliculi<br />

was observed in Control-iPS-HLCs, but not in BD-iPS-HLCs,<br />

suggesting that the biliary excretion capacity was impaired<br />

in BD-iPS-HLCs. Thus, the data implied that BD-iPS-HLCs could<br />

reproduce the pathophysiology of PFIC2. Conclusions: BD-iPS-<br />

HLCs recapitulated the PFIC2 phenotype, mislocalization of<br />

BSEP, and impairment of biliary excretion. Thus, this PFIC2<br />

model can be applied to elucidate disease mechanisms and to<br />

determine novel therapeutic options.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kazuo Imagawa, Kazuo<br />

Takayama, Shigemi Isoyama, Ken Tanikawa, Masato Shinkai, Masayoshi Kage,<br />

Kenji Kawabata, Ryo Sumazaki, Hiroyuki Mizuguchi<br />

1691<br />

King’s-Variceal Prediction Score (K-VaPS); a non invasive<br />

assessment tool of severe portal hypertension in<br />

cirrhotic children<br />

Peter Witters, Dominic A. Hughes, Palaniswamy Karthikeyan,<br />

Alastair J. Baker, Mark Davenport, Anil Dhawan, Tassos Grammatikopoulos;<br />

Paediatric Liver GI & Nutrition Centre, King’s College<br />

Hospital NHS Foundation Trust, London, United Kingdom<br />

Background & aim: Variceal hemorrhage is a serious complication<br />

of chronic liver disease(CLD) in children. There is limited<br />

knowledge around the best prophylactic management and<br />

selection criteria of children undergoing surveillance endoscopy<br />

(OGD). We derive a new prediction model of clinically<br />

significant varices (CSV) and validate it in a separate cohort.<br />

Methods: All patients presenting in our centre with suspected<br />

PHT or gastrointestinal (GI) bleeding due to CLD and undergoing<br />

a first OGD between January 2005- December 2012 were<br />

included. A validation cohort from 2013-2015 was obtained.<br />

Clinical, biochemical and radiological data were collected.<br />

Results: Data on 124(67M) treatment-naïve patients were collected.<br />

Mean age was 8.81 ± 5.60 years. 50% had biliary<br />

atresia and the remainder various other CLD types. 35(28%)<br />

presented with GI bleeding and overall 79(64%) were in<br />

CSV+ve. Mean values in CSV+ve vs. CSV-ve group were for<br />

haemoglobin(Hb)(mg/dL) [10,58 vs. 11,62 p=0.008], platelet<br />

count(10 9 /ml) [111 vs. 167, p=0.001], white cell count(10/<br />

ml) [5.3 vs. 6.9, p=0.019], INR[1.22 vs. 1.13, p=0.004], bilirubin(mg/dL)<br />

[53.33 vs. 28.22, p=0.014], albumin(g/L) [37<br />

vs. 41, p=0.0001], AST(IU/L) [121 vs. 104, p=0.815], spleen<br />

size(cm) [16.41 vs. 14.49, p=0.073], spleen size(z-score)<br />

[9.1 vs. 6.35, p=0.009] and equivalent adult spleen size(E-<br />

ASS) (cm) [23.21 vs. 19.3, p=0.003], respectively. Previously<br />

published predictions rules had, at optimal cut-off, a sensitivity<br />

and specificity of 76% and 59%(CPR), 60% and 55%(APRI),<br />

80% and 59%(VPR), respectively. To construct a new prediction<br />

score Pearson’s bivariate correlation was performed.<br />

Variables significantly correlating with CSV were Hb(R=-.250,<br />

p= 0.005), platelet count(R=-.276, p=0.002), INR(R=0.223,<br />

p=0.010), albumin(R=-.334, p


1034A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Th1 cytokines (TNF-α and IFN-γ) and innate cytokines (IL-1, IL-6)<br />

in WT-RRV mice but not BSS controls or Ig-α -/- RRV groups. Conclusions:<br />

The RRV-induced mouse model of BA is associated<br />

with an abnormal decrease in the expression of the majority<br />

of transporters involved in bile excretion. The high levels of<br />

retained bile acids may contribute to liver injury. The lack of<br />

inflammation within the Ig-α -/- RRV mice is associated with normal<br />

expression of the majority of bile acid transporter proteins,<br />

suggesting that the inflammation present in RRV-induced BA<br />

mice inappropriately inhibits bile acid transporter expression.<br />

Disclosures:<br />

The following authors have nothing to disclose: Asokan Rengasamy, Karim C.<br />

El Kasmi, Brianna Traxinger, Jonathan Roach, Natarajan Balasubramaniyan,<br />

Cara L. Mack<br />

1693<br />

Exome Sequencing and de novo Mutation Analysis<br />

Reveal a Highly Connected Gene Network in Isolated<br />

Biliary Atresia<br />

Kathleen M. Loomes 5,1 , Ramakrishnan Rajagopalan 2 , Christopher<br />

Grochowski 2 , Ying Chen 3,4 , Melissa Gilbert 2 , Henry C. Lin 5,1 , Marcella<br />

Devoto 4,1 , Nancy B. Spinner 2,6 ; 1 Department of Pediatrics,<br />

University of Pennsylvania Perelman School of Medicine, Philadelphia,<br />

PA; 2 Department of Pathology and Laboratory Medicine, The<br />

Children’s Hospital of Philadelphia, Philadelphia, PA; 3 Genomics<br />

and Computational Biology Graduate Group, University of Pennsylvania<br />

Perelman School of Medicine, Philadelphia, PA; 4 Division<br />

of Human Genetics, The Children’s Hospital of Philadelphia,<br />

Philadelphia, PA; 5 Division of Gastroenterology, Hepatology and<br />

Nutrition, The Children’s Hospital of Philadelphia, Philadelphia,<br />

PA; 6 Department of Pathology and Laboratory Medicine, University<br />

of Pennsylvania Perelman School of Medicine, Philadelphia, PA<br />

Biliary atresia (BA) is a severe pediatric liver disease resulting<br />

in necroinflammatory obliteration of the extrahepatic biliary<br />

tree. The majority of BA patients develop biliary cirrhosis and<br />

50% will require liver transplantation in the first two years of<br />

life. Incidence of BA is reported to be between 1 in 18,000<br />

and 1 in 8,000 depending on ethnicity. The etiology of BA<br />

is unknown, with evidence for infectious, environmental, and<br />

genetic risk factors described. The purpose of this study was to<br />

investigate possible genetic etiologies of BA by whole exome<br />

sequencing of non-syndromic BA probands and their parents<br />

(trios). DNA samples were obtained from 10 well-characterized<br />

non-syndromic BA probands and their parents from the<br />

NIDDK-funded Childhood Liver Disease Research Network<br />

(ChiLDReN). We performed whole-exome sequencing in these<br />

10 trios and looked for de novo mutations under the assumption<br />

that mutations only seen in probands, but not the parents,<br />

are more likely to be involved in the etiology of the disease.<br />

We used a custom in-house computational analysis pipeline to<br />

call variants and identify de novo mutations. We found a total<br />

of 33 exonic de novo variants in 33 different genes. These<br />

were 2 loss-of-function mutations, 3 in-frame insertions/deletions,<br />

2 splice-site mutations and 26 missense mutations. The<br />

number of de novo changes per proband ranged from 1 to 9<br />

with a mean of 3.3. Seven of the 10 probands carried more<br />

than 1 de novo mutation. However, none of the mutations was<br />

seen in more than one proband. We used protein-interaction<br />

network analysis tools such as GeneMANIA, DAPPLE and Ingenuity<br />

Pathway Analysis (IPA) to assess 1) whether the genes<br />

with de novo mutations are connected with each other directly<br />

or through a single common partner, 2) whether the genes<br />

with de novo mutations form a network with genes previously<br />

implicated in BA. Gene network analysis with GeneMANIA<br />

suggested a highly interconnected network of proteins that<br />

are co-expressed. Protein interaction analysis with DAPPLE<br />

suggested that 17 out of 33 genes (51%) interact with each<br />

other directly or through a single common partner to form a<br />

tight network. We used IPA to explore the paths between the<br />

genes with de novo mutations and genes previously implicated<br />

in BA. Network connectivity analyses with IPA suggest that<br />

29 out of 33 genes (88%) are at most 2 nodes away from a<br />

group of 19 selected BA candidate genes. Results from network<br />

analysis suggest a shared etiology to seemingly sporadic de<br />

novo mutations observed in probands with isolated BA. To our<br />

knowledge, this is the first report of trio analysis for isolated BA<br />

using next-generation sequencing technology.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kathleen M. Loomes, Ramakrishnan<br />

Rajagopalan, Christopher Grochowski, Ying Chen, Melissa Gilbert, Henry<br />

C. Lin, Marcella Devoto, Nancy B. Spinner<br />

1694<br />

DXA Bone Density in Alagille Syndrome Correlates with<br />

Fracture History and Degree of Cholestasis<br />

Kathleen M. Loomes 2 , Cathie Spino 14 , Nathan P. Goodrich 1 ,<br />

Thomas N. Hangartner 15 , Amanda E. Marker 15 , James E. Heubi 16 ,<br />

Binita M. Kamath 13 , Benjamin Shneider 17 , Philip Rosenthal 3 ,<br />

Paula M. Hertel 17 , Saul J. Karpen 4 , Nanda Kerkar 5,19 , Jean P.<br />

Molleston 6 , Karen F. Murray 7 , Kathleen B. Schwarz 8 , Jeffrey Teckman<br />

18 , Yumirle P. Turmelle 9 , Peter F. Whitington 20 , Averell H.<br />

Sherker 10 , John C. Magee 11 , Ronald J. Sokol 12 ; 1 Arbor Research<br />

Collaborative for Health, Ann Arbor, MI; 2 Pediatric Gastroenterology,<br />

Hepatology and Nutrition, Children’s Hospital of Philadelphia,<br />

Philadelphia, PA; 3 Division of GI, Hepatology and Nutrition,<br />

UCSF-University of California, San Francisco, San Francisco, CA;<br />

4 Pediatric Gastroenterology, Hepatology and Nutrition, Emory<br />

University School of Medicine/Children’s Healthcare of Atlanta,<br />

Atlanta, GA; 5 Children’s Hospital of Los Angeles, Los Angeles, CA;<br />

6 Pediatric Gastroenterology, Hepatology and Nutrition, Indiana<br />

University School of Medicine/Riley Hospital for Children, Indianapolis,<br />

IN; 7 Division of Gastroenterology and Hepatology, University<br />

of Washington Medical Center, Seattle Children’s, Seattle,<br />

WA; 8 Johns Hopkins School of Medicine, Baltimore, MD; 9 Washington<br />

University School of Medicine, Saint Louis, MO; 10 Liver Disease<br />

Research Branch, National Institute of Diabetes and Digestive<br />

and Kidney Diseases, National Institutes of Health, Bethesda, MD;<br />

11 University of Michigan Medical School, Ann Arbor, MI; 12 Section<br />

of Pediatric Gastroenterology, Hepatology and Nutrition,<br />

Department of Pediatrics, University of Colorado, Aurora, CO;<br />

13 Division of Gastroenterology, Hepatology and Nutrition, Hospital<br />

for Sick Children and University of Toronto, Toronto, ON, Canada;<br />

14 University of Michigan, Ann Arbor, MI; 15 Department of<br />

Biomedical, Industrial, & Human Factors Engineering, Wright State<br />

University, Dayton, OH; 16 Division of Pediatric Gastroenterology,<br />

Hepatology and Nutrition, Cincinnati Children’s Hospital Medical<br />

Center, Cincinnati, OH; 17 Pediatric Gastroenterology, Hepatology<br />

and Nutrition, Baylor College of Medicine, Houston, TX; 18 Saint<br />

Louis University, Cardinal Glennon Children’s Medical Center,<br />

St. Louis, MO; 19 Mount Sinai, New York, NY; 20 Department of<br />

Pediatrics, Ann & Robert H. Lurie Children’s Hospital of Chicago,<br />

Northwestern University Feinberg Medical School, Chicago, IL<br />

Osteopenia and bone fractures (fx) are significant causes of<br />

morbidity in children with cholestatic liver disease. We performed<br />

dual energy X-ray absorptiometry (DXA) analysis in<br />

a cohort of children with inherited cholestatic diseases and<br />

explored associations with anthropometrics, laboratory measurements<br />

and fx history. Methods: Subjects were enrolled<br />

in the Longitudinal Study of Genetic Causes of Intrahepatic<br />

Cholestasis (LOGIC) in the NIDDK-funded Childhood Liver


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1035A<br />

Disease Research Network (ChiLDReN). DXA was performed<br />

on subjects age > 5 years (with native liver), diagnosed with<br />

Bile Acid Synthetic Disorder (BAD; n=14), Alpha-1 antitrypsin<br />

deficiency (A1AT; n=44) and Alagille syndrome (ALGS;<br />

n=49). Pearson correlation coefficients examined associations<br />

between subject reference (adjusted for age, gender and race)<br />

and anthropometrically-adjusted DXA Z-scores and laboratory<br />

values (total bilirubin (TB), serum bile acids (SBA) and 25-OH<br />

vitamin D (vitD)) and fx history. Results: Weight, height and<br />

BMI Z-scores were all lower, and TB and SBA higher, in the<br />

ALGS group (p1 fx; 21% and 7% in BAD;<br />

20% and 2% in A1AT). Bone mineral density (BMD) reference<br />

measures did not differ among disease groups except for<br />

total hip BMD and femoral neck BMD, which were lower in<br />

ALGS (p=0.007 and p=0.0002). Total body and total body<br />

minus head bone mineral content (BMC) reference measures<br />

were lower in the ALGS subjects (p


1036A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1696<br />

Diagnostic determination system for high-risk screening<br />

for inborn errors of bile acid metabolism: Results during<br />

20 years in East Asia<br />

Nakayuki Naritaka 1 , Mitsuyoshi Suzuki 1 , Hajime Takei 2 , Akihiko<br />

Kimura 3 , Takao Kurosawa 4 , Takashi Iida 5 , Hiroshi Nittono<br />

2 , Toshiaki Shimizu 1 ; 1 Pediatrics, Juntendo University Faculty<br />

of Medicine, Tokyo, Japan; 2 Junshin Clinic bile acid institute,<br />

Tokyo, Japan; 3 Pediatrics, Kurume University School of Medicine,<br />

Fukuoka, Japan; 4 Faculty of Pharmaceutical Sciencies, Health Services<br />

University of Hokkaido, Hokkaido, Japan; 5 Nihon University<br />

College of Humanities and Sciences, Tokyo, Japan<br />

BACKGROUND: Some patients with cholestasis of unknown<br />

cause may have inborn errors of bile acid metabolism (IEBAM),<br />

leading to abnormalities of bile acid biosynthesis. Although<br />

many types of bile acid synthesis defects have been reported<br />

for this disorder, no detailed information on its incidence and<br />

other aspects in East Asia has been available. To elucidate the<br />

current status of IEBAM, in July 1996, a diagnostic determination<br />

system was established for high-risk screening for IEBAM.<br />

METHODS: The target patients included children with hepatic<br />

disorders for which the cause could not be identified on conventional<br />

liver function testing or hepatitis virus testing, patients<br />

with liver cirrhosis of unknown etiology, sibling cases, and<br />

patients with cholestasis who exhibited serum levels of direct<br />

bilirubin (D-Bil) was ^2.0mg/dL,besides total bile acid and<br />

γ-glutamyltranspeptidase (GGT) were within the normal range.<br />

Urine samples were sent to the Bile Acid Institute which located<br />

in Tokyo, via refrigerated delivery service or impregnated filter<br />

paper with sufficient urine volume. The samples were analyzed<br />

during the 20 years between July 1996 and May 2015.Urinary<br />

bile acids were analyzed via gas chromatography–mass<br />

spectrometry (GC/MS) or liquid chromatography-tandem mass<br />

spectrometry (LC-MS/MS). RESULTS: In 960 cases analyzed<br />

over a 20-year period, 12 samples were differentially diagnosed<br />

with IEBAM, including: 3β-hydroxy-Δ 5 -C 27<br />

-steroid dehydrogenase/isomerase<br />

deficiency in 5 (2 in Japan, 2 in China,<br />

1 in Thailand), 3-oxo-Δ 4 -steroid 5β-reductase deficiency in 3<br />

(all in Japan), oxysterol 7α-hydroxylase deficiency in 3 (1 in<br />

Japan, 2 in Taiwan), bile acid-CoA amino acid N-acyltransferase<br />

deficiency in 1 (speculated in Thailand). Especially in<br />

Japanese cases, 2 cases each of 3β-hydroxy-Δ 5 -C 27<br />

-steroid<br />

dehydrogenase/isomerase deficiency and 3-oxo-Δ 4 -steroid<br />

5β-reductase deficiency were effectively treated with oral bile<br />

acid therapy. The third case of 3-oxo-Δ 4 -steroid 5β-reductase<br />

deficiency had spontaneous remission without oral therapy. The<br />

Japanese case of oxysterol 7α-hydroxylase deficiency underwent<br />

successful heterozygous living donor liver transplantation.<br />

CONCLUSIONS: Twelve patients were identified with cholestasis<br />

of unknown cause in East Asia during 20 years. IEBAM is<br />

a rare hereditary disease; therefore, much cost and labor is<br />

spent to make a definitive diagnosis via genetic analysis. The<br />

diagnostic determination system with urinalysis is considered<br />

useful for diagnosis and treatment planning, and furthermore,<br />

contributes to improved treatment efficacy for IEBAM.<br />

Disclosures:<br />

The following authors have nothing to disclose: Nakayuki Naritaka, Mitsuyoshi<br />

Suzuki, Hajime Takei, Akihiko Kimura, Takao Kurosawa, Takashi Iida, Hiroshi<br />

Nittono, Toshiaki Shimizu<br />

1697<br />

Role of transient elastography to differentiate children<br />

with biliary atresia from other causes of neonatal<br />

cholestasis<br />

Bikrant B. Lal, Rajeev Khanna, Vikrant Sood, Dinesh Rawat, Seema<br />

Alam; DEPARTMENT OF PEDIATRIC HEPATOLOGY, INSTITUTE<br />

OF LIVER AND BILIARY SCIENCES, New Delhi, India<br />

Transient elastography is an important non-invasive tool to<br />

detect fibrosis. Children with biliary atresia need to be diagnosed<br />

early as outcome post Kasai surgery depends on age at<br />

operation. Fibroscan can serve as an important non-invasive<br />

tool and an alternative to liver biopsy. Objective: To study the<br />

role of transient elastography to differentiate biliary atresia<br />

from neonatal hepatitis. To evaluate its efficacy in predicting<br />

outcome post kasai portoenterostomy. Methodology: All<br />

patients with neonatal cholestasis were included in the study<br />

and evaluated as per the instituitional protocol. Transient elastography<br />

(TE) was done in a 4-hour fasting state using Fibro-<br />

Scan502, Echosens, Paris, France. Median of 10 successful<br />

readings was taken and values with IQR>30% or success rate<br />

of


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1037A<br />

in experimental and human BA. Methods: Neonatal mice were<br />

challenged with saline or RRV soon after birth. Gene expression<br />

was determined by real-time PCR and immune cells were quantified<br />

by flow cytometry. Histology was performed using H/Estained<br />

sections. Immunofluorescence was performed using<br />

antibodies against C5aR, CD68, pan-cytokeratin (CK) and<br />

Vimentin. Liver wedge biopsies were obtained from patients at<br />

Kasai and explants at transplantation. Genome-wide expressions<br />

were obtained from GEO Series: GSE46995. Results: In<br />

diseased mice, expression levels of Cd68, Cd14 and Mpeg1<br />

progressively increased from the inception of epithelial injury<br />

to complete duct atresia (days 3–14, 1.7–4.5 fold above<br />

saline-injected mice; P=0.001). Proinflammatory macrophage<br />

polarization correlated with increased expressions of Ccl2,<br />

Ccl3, Ccl4, Ccr7, Tlr2 and Tlr6 (1.3–8.3 fold above controls;<br />

P


1038A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and 1764 (p=0.04) in the basal core promoter region. In<br />

patients >12 years (N=75), 3 groups of baseline mutations<br />

were significantly associated with long-term ALT normalization<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1039A<br />

1703<br />

Dissecting cellular entry mechanisms by enveloped and<br />

nonenveloped hepatitis E virus<br />

xin yin, Zongdi Feng; Research Institute at Nationwide Children’s<br />

Hospital, Columbus, OH<br />

Hepatitis E virus (HEV) infection is a significant global health<br />

issue. Recent <strong>studies</strong> show that while nonenveloped HEV virions<br />

are shed in feces, the virus circulating in blood is cloaked<br />

in host membrane and hidden from neutralizing antibodies.<br />

To better understand HEV infection and pathogenesis, we<br />

investigate the mechanism by which enveloped HEV (eHEV)<br />

enters the cell. Gradient purified enveloped and non-enveloped<br />

HEV (genotype 3, Kernow C1/P6) were inoculated to<br />

HepG2 cells and cellular pathways and factors involved in<br />

HEV entry were studied by using pharmacological inhibitors<br />

and siRNA mediated gene depletion. Compared to non-enveloped<br />

HEV, cellular attachment of eHEV is much less efficient<br />

and its entry is also much slower. Entry of eHEV is also more<br />

sensitive to dynamin-2 and clathrin depletion, suggesting eHEV<br />

internalization involves clathrin-mediate endocytosis. Knockdown<br />

of Rab7A and Rab5A had a greater inhibitory effect<br />

on the entry of eHEV than nonenveloped HEV. In addition,<br />

treatment of cells with lysosomotropic agents bafilomycin A or<br />

NH 4<br />

Cl almost completely blocked the entry of eHEV, but had<br />

no effect on the entry of nonenveloped HEV, indicating eHEV<br />

also requires endosomal acidification for its entry. Finally,<br />

depletion of NPC-1 (Niemann-Pick disease, type C1), a host<br />

protein involved in cholesterol transport in late endosomes/<br />

lysosomes and Ebolavirus entry, greatly reduced entry of eHEV,<br />

but not nonenveloped HEV. Conclusion: Entry of eHEV involves<br />

dynamin-2 and clathrin-mediated endocytosis and requires<br />

endosomal acidification. eHEV envelope is likely degraded in<br />

endolysosomes in order for the viral capsid to interact with its<br />

cellular receptor. These results have important implications for<br />

understanding HEV infection and pathogenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: xin yin, Zongdi Feng<br />

1704<br />

Pre-core region mutations analysis helps to delineate<br />

immune stage in HBeAg+ children with chronic hepatitis<br />

B<br />

Ivana Carey, Sanjay Bansal, Sarah Tizzard, Matthew J. Bruce,<br />

Mary Horner, Kosh Agarwal, Diego Vergani, Giorgina Mieli-Vergani;<br />

Institute of Liver Studies, Kings College School of Medicine at<br />

King’s College Hospital, London, United Kingdom<br />

Interaction between the hepatitis B virus (HBV) and the immune<br />

system is crucial for the outcome of chronic hepatitis B (CHB). A<br />

recent study shows that presence of pre-core region mutations<br />

helps distinguishing between the immune stages of HBeAg+<br />

CHB in adults, immunotolerant patients having no mutations<br />

and higher HBsAg plasma levels. There are no data on pre-core<br />

region mutations in relation to the immune stage in HBeAg+<br />

CHB children. In this retrospective cross-sectional study on<br />

HBeAg+ CHB children with different alanine-aminotransferase<br />

(ALT) profiles, we aimed to evaluate the frequency of precore<br />

region mutations – basal core promoter (BCP) (A1762T/<br />

G1764A) and stop codon (A1896G) - and its relationship with<br />

semi-quantitative HBeAg and HBsAg levels, HBV DNA viral<br />

load, HBV genotypes and levels of interferon-gamma-inducible<br />

protein 10 (IP10), a chemokine reflecting the activation status<br />

of the innate immune system in CHB. Patients: 137 children<br />

(median age 9.55 years) with HBeAg+ CHB were followed<br />

for at least 2 years (median 5 years) and, based on their ALT<br />

profiles, were divided into 3 groups: persistently normal (N)<br />

ALT (n=64), fluctuating (F) ALT (n=43) and persistently abnormal<br />

(A) ALT (n=30). Material and methods: Plasma samples at<br />

diagnosis were used to evaluate BCP and stop codon 1896<br />

mutations and HBV genotypes by direct population sequencing;<br />

semi-quantitative HBeAg and HBsAg plasma levels were<br />

measured by Abbott ARCHITECT® assay [S/CO & log 10<br />

IU/<br />

ml]; HBV DNA viral load by real-time PCR [log 10<br />

IU/ml] and<br />

IP10 plasma levels by ELISA [pg/ml]. Results: Frequencies of<br />

BCP and stop codon mutations were higher in patients with<br />

fluctuating (F) and abnormal (A) ALT profiles than in patients<br />

with normal (N) ALT levels (p


1040A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Statistical analysis was performed using STATA software<br />

version 12.0. Statistical significance was defined as P-value<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1041A<br />

on established exposure-safety analyses. Other PK parameters<br />

were similar between the two populations. Conclusion In<br />

HCV-infected adolescents (12 to< 18 years), SOF 400 mg<br />

(+RBV) or LDV/SOF 90 mg/400 mg were well tolerated and<br />

provided comparable plasma exposures to those observed in<br />

adults. These data confirm the appropriateness of SOF or LDV/<br />

SOF adult clinical doses in adolescents.<br />

Disclosures:<br />

Brian Kirby - Employment: Gilead Sciences<br />

Polina German - Employment: GIlead Sciences, Inc; Stock Shareholder: GIlead<br />

Sciences, Inc<br />

Bittoo Kanwar - Employment: Gilead Sciences<br />

Isobel Lakatos - Management Position: Gilead Sciences, INC; Stock Shareholder:<br />

Gilead Sciences, Inc<br />

Anita Mathias - Employment: Gilead Sciences Inc.,<br />

The following authors have nothing to disclose: Liyun Ni, John Ling<br />

1708<br />

HCV infection shifts CD8+ T cell Naïve and Memory<br />

Subsets in the Maternal Fetal Interface (MFI)<br />

Melissa A. Sheiko 1 , Rachel McMahan 2 , Lucy Golden-Mason 2 ,<br />

Christine Waasdorp Hurtado 1 , Mona Krull 3 , Silvia Giugliano 2 ,<br />

Angela M. Mitchell 2 , Michael R. Narkewicz 1 , Hugo R. Rosen 2 ;<br />

1 Pediatrics, Digestive Health Institute, University of Colorado Denver<br />

School of Medicine, Aurora, CO; 2 Department of Medicine,<br />

Gastroenterology, University of Colorado Denver School of Medicine,<br />

Aurora, CO; 3 Ob/Gyn, DenverHealth Medical Center, Denver,<br />

CO<br />

Introduction: Women with Hepatitis C (HCV) infection have low<br />

rates of vertical transmission (VT) to their infants (2-6%) compared<br />

to HIV and hepatitis B. It has been postulated that this is<br />

due to an immune response at the maternal fetal interface (MFI).<br />

The MFI is comprised of the maternal blood, the decidua (the<br />

part of the placenta that contacts the maternal blood), the placenta,<br />

and the infant’s blood. Methods: Samples were collected<br />

from mother and infant pairs with and without HCV. Mononuclear<br />

cells were isolated from the maternal blood, decidua,<br />

placenta, and cord blood. Multi-parameter flow cytometry<br />

was performed, using antibodies to CD3, CD8, CCR7, and<br />

CD45RA. All cells that were analyzed were CD3+ and CD8+ T<br />

cells. Naïve cells were also CCR7+CD45RA+, central memory<br />

cells were CCR7+CD45RA, effector memory cells were CCR7-<br />

CD45RA-, and terminally-differentiated effector memory cells<br />

were CCR7-CD45RA+. GraphPad Prism software was used<br />

for analysis: medians are reported, and p values were determined<br />

by the Mann Whitney U test. Results: The CD8+ T cells<br />

from the decidua showed differences in all subsets of CD8+<br />

T cells between controls and HCV-exposed subjects: there<br />

were fewer naïve cells in HCV-exposed (15.7% versus 36.0%;<br />

p


1042A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

for inflammatory bowel disease (250-450). Our study did not<br />

reveal specific goal levels to aim for in the treatment of AIH, but<br />

would suggest that patients should be maintained at the lowest<br />

level that keeps them in remission.<br />

of PSC and 64% of Overlap. Conclusion: The nomenclature of<br />

autoimmune liver disease remains inadequate and in need of<br />

consensus. Of young adult PSC patients, half presented with an<br />

AISC phenotype, and of young adult AIH-PSC overlap, nearly<br />

half presented as AIH. Evaluating whether immunosuppressive<br />

therapy in childhood prevents ultimate disease progression<br />

remains an important goal for prospective <strong>studies</strong>.<br />

Characteristics of children with AIH, PSC and AISC at diagnosis<br />

Disclosures:<br />

The following authors have nothing to disclose: Melissa A. Sheiko, Kelley E.<br />

Capocelli, Shikha Sundaram, Zhaoxing Pan, Annette McCoy, Cara L. Mack<br />

Disclosures:<br />

James W. Ferguson - Advisory Committees or Review Panels: Astellas, Novartis<br />

David H. Adams - Advisory Committees or Review Panels: GSK, Proximagen;<br />

Grant/Research Support: Takeda, Biotie Therapies, Novimmune, ChemoCentryx,<br />

Novimmune, Biotie Therapies; Patent Held/Filed: biotie therapies, Biotie Therapies,<br />

Biotie Therapies, Biotie Therapies, Biotie Therapies, Biotie Therapies, Biotie<br />

Therapies, Biotie Therapies<br />

Deirdre A. Kelly - Consulting: Sanofi Pasteur; Grant/Research Support: BMS,<br />

Astellas, Acitllion, MSD, Roche<br />

Gideon Hirschfield - Advisory Committees or Review Panels: Intercept Pharma;<br />

Consulting: Dignity Sciences, GSK, NGM Bio, Lumena, J & J; Grant/Research<br />

Support: BioTie; Speaking and Teaching: Falk Pharma<br />

The following authors have nothing to disclose: Jeremy K. Rajanayagam, Ye H.<br />

Oo<br />

1710<br />

Long-term follow up of childhood autoimmune liver disease:<br />

the importance of resolving nomenclature<br />

Jeremy K. Rajanayagam 1,3 , James W. Ferguson 2 , Ye H. Oo 2,3 ,<br />

David H. Adams 2,3 , Deirdre A. Kelly 1 , Gideon Hirschfield 2,3 ; 1 Liver<br />

Unit, BIrmingham Children’s Hospital, Birmingham, United Kingdom;<br />

2 Liver Unit, Queen Elizabeth Hospital, Birmingham, United<br />

Kingdom; 3 Centre for Liver Research, University of Birmingham,<br />

Birmingham, United Kingdom<br />

Background: The nomenclature and long term outcomes for<br />

patients with autoimmune liver diseases transitioned to adult<br />

practice lacks consensus. Aim: To study the nature and progression<br />

of childhood onset autoimmune liver disease in patients<br />

transitioned from pediatric to adult liver care. Methods: Clinical<br />

review of transitioned patients with standard definitions<br />

of AIH and PSC was performed. AISC was defined as cholangiographic<br />

and/or histologic evidence of biliary involvement,<br />

clinical features of autoimmunity (increased IgG, and positive<br />

autoantibodies), and histological findings of AIH. The term AIH-<br />

PSC overlap was used for concomitant occurrence of clinical,<br />

biochemical, serological and/or histological features of PSC<br />

and AIH. Results: 73 children with pediatric autoimmune liver<br />

disease were transitioned over a median follow-up of 13 years<br />

(IQR 8-16 years). At presentation, 52 were diagnosed with AIH<br />

(38F), 4 with PSC (1F); 17 with AISC (8F) (Table). IBD occurred<br />

in 4% of AIH, 25% of PSC and 65% of AISC at diagnosis with<br />

liver disease. Median IAIHG score at diagnosis was 19 (IQR<br />

17-21), 4 (IQR 3.5-5) and 12 (IQR 10-14) for AIH, PSC and<br />

AISC, respectively. Children with AIH & AISC were treated<br />

with corticosteroids and azathioprine, with ursodeoxycholic<br />

acid added in AISC. 10/52 (19%) patients initially presenting<br />

as AIH had phenotypic progression by last adult review: 6<br />

(11.5%) developed an overlap syndrome whilst 4 (7.6%) had<br />

clinically dominant PSC. 8/17 (47%) AISC patients retained<br />

an adult overlap phenotype, but 9 (53%) had clinically dominant<br />

PSC. Those presenting as childhood PSC (n=4) remained<br />

phenotypically as PSC in adulthood. Thus by last review in<br />

adult care the clinical diagnoses were AIH n=42, PSC n=17<br />

and Overlap AIH/PSC n=14. IBD occurred in 5% of AIH, 65%<br />

1711<br />

Autoimmune Liver disease in Children with Sickle Cell<br />

Disease<br />

Suttiruk Jitraruch 2 , Emer Fitzpatrick 2 , Maesha Deheragoda 2 , Giorgina<br />

Mieli-Vergani 2 , Sue Height 1 , Nedim Hadzic 2 , Marianne<br />

Samyn 2 ; 1 Paediatric Haematology, King’s College Hospital, London,<br />

United Kingdom; 2 Institute of Liver Studies, King’s College<br />

Hospital, London, United Kingdom<br />

Sickle cell disease (SCD) is an autosomal recessive haemoglobinopathy<br />

resulting in intermittent haemolysis and microvascular<br />

occlusions. Hepatic involvement varies from asymptomatic gallstones<br />

to life-threatening acute liver failure (ALF) and cirrhosis.<br />

Aim: To characterise the clinical features, natural history and<br />

outcome of autoimmune liver disease (AILD) in patients with<br />

SCD. Methods: We performed a retrospective review of SCD<br />

patients with hepatic dysfunction referred to our centre from<br />

1999-2015. The demographic, clinical, laboratory, histological<br />

and radiological features, management and outcome were<br />

studied. The diagnostic criteria included: positive serum autoantibodies,<br />

hypergammaglobulinemia, compatible histology<br />

and absence of viral/metabolic causes. Results: Of 83 SCD<br />

patients with hepatic dysfunction, 13 (16%), (10 female) were<br />

diagnosed with AILD at a median age of 11 (range, 3.4-16)<br />

years. Eleven were homozygote for HBSS and 2 required regular<br />

transfusions. Family history for autoimmune disease was<br />

positive in 2. Two patients presented with ALF (INR 2.7 and<br />

1.8). In two patients parvo B19 induced aplastic crisis preceded<br />

AILD diagnosis, 2 and 6 months, respectively. At presentation,<br />

anti-nuclear and anti-smooth muscle autoantibodies<br />

were in range 1/20-1/2560 and 1/10-1/320, respectively,<br />

median AST 294 (range, 67-814) U/L, IgG 33.5 (range, 13.7-<br />

43.7) g/L and INR 1.32 (range, 1.01-2.7). Ultrasonography<br />

showed enlarged lymph nodes at porta/superior to pancreas<br />

in 4, gall stones in 3, and splenomegaly in 5 patients. On<br />

MRCP five children had radiological features of cholangiopathy;<br />

4 at presentation and 1 three years later. Liver biopsy<br />

was performed in 11 (6 via transjugular route) without complications;<br />

9 showed interface hepatitis without cholangiopathy,


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1043A<br />

one obtained after treatment and another one was inadequate.<br />

Treatment included ursodeoxycholic acid (12), prednisolone<br />

(12), azathioprine (8) and mycophenolate mofetil (1). After a<br />

median follow up of 3.8 (range, 0.2-14.3) years, 10 patients<br />

are alive with 2 lost to follow up. One patient died following<br />

intracranial haemorrhage. One patient required liver transplantation,<br />

6.4 years after diagnosis due to recurrent biliary<br />

sepsis. Four patients were in full and five in partial remission.<br />

Four patients (2 males) were diagnosed with ulcerative colitis<br />

2 before and 2 after AILD. Conclusion: AILD is not uncommon<br />

in patients with SCD with a strong female preponderance. It<br />

responds well to standard treatment. Liver biopsy can be helpful<br />

to confirm the diagnosis. Ulcerative colitis was more common<br />

in the boys and should be excluded.<br />

Disclosures:<br />

Nedim Hadzic - Consulting: Alnylam, Cambridge, MA; Speaking and Teaching:<br />

Synegeva UK<br />

The following authors have nothing to disclose: Suttiruk Jitraruch, Emer Fitzpatrick,<br />

Maesha Deheragoda, Giorgina Mieli-Vergani, Sue Height, Marianne<br />

Samyn<br />

1712<br />

“Hepatic Granulomas: 18 Year Experience of a Tertiary<br />

Paediatric Liver Centre in the UK”<br />

Marumbo P. Mtegha, Katie Hunt, Nedim Hadzic, Maesha Deheragoda,<br />

Sanjay Bansal; Department of Paediatric Hepatology,<br />

Gastroenterology and Nutrition, King’s College Hospital, London,<br />

United Kingdom<br />

Objective Hepatic granulomas are reported in 2-15% of all<br />

liver biopsies in adult <strong>studies</strong> . Up to 30% have no identifiable<br />

etiology. The incidence in the paediatric population is not well<br />

established. We describe the prevalence, character and aetiology<br />

of this pathology at our centre. Methods Retrospective<br />

analysis of all liver biopsies showing granulomas and their<br />

clinical data between 1996-2014. Results Granulomas were<br />

identified in 44 (0.32%) of 13542 paediatric liver biopsies<br />

perfomed. Native Liver 31 (70.4%) patients had granuloma<br />

within their native liver (Male 52%). The median age at presentation<br />

was 3.7yrs (range 0.3–11.8). The commonest presentation<br />

was abnormal liver function tests (LFTs) with or without<br />

cholestasis. 4 patients presented with isolated or a combination<br />

of hepatomegaly, splenomegaly and lymphadenopathy.<br />

Etiologies are listed in Table 1. Histologically, non-caseating<br />

and epithelioid granulomas were found in 26 cases. Lipogranuloma<br />

were described in 2 cases, one each in Non-alcoholic<br />

fatty liver disease and HIV infection. There were 3 cases of<br />

caseating granuloma (2 Sarcoidosis and Tuberculosis). Transplanted<br />

Liver 13 (29.6%) patients had granulomas within their<br />

allograft. The median age at liver transplant was 1.4 years<br />

(range 0.3-15.9). Median time from transplant to granuloma<br />

detection was 0.2yrs (range 8 days to-8.1yrs). All granulomas<br />

were non-caseating and epithelioid. Acute cellular rejection<br />

was noted in 3 cases, chronic rejection in 2. 7 cases showed<br />

non-specific lobular hepatitic changes, the last showed early<br />

erythrophagocytosis. Conclusions Hepatic granulomas in children<br />

are rare with idiopathic neonatal hepatitis being the most<br />

common condition in the native liver. In the post transplant<br />

setting, they are more common in the younger age group and<br />

in the first few months after transplant. In this setting, they may<br />

be associated with acute or chronic rejection.<br />

*= Persistent biochemical or radiological abnormalities<br />

**=Normalization of biochemistry and radiology<br />

***=Incomplete follow-up data<br />

Disclosures:<br />

Nedim Hadzic - Consulting: Alnylam, Cambridge, MA; Speaking and Teaching:<br />

Synegeva UK<br />

The following authors have nothing to disclose: Marumbo P. Mtegha, Katie Hunt,<br />

Maesha Deheragoda, Sanjay Bansal<br />

1713<br />

Differentiating autoimmune hepatitis from lupus-associated<br />

hepatitis: A paradigm shift in approach to elevated<br />

liver enzymes<br />

Andreanne Benidir 1 , Brian M. Feldman 3 , Anna Goldenberg 2 , Earl<br />

Silverman 3 , Sindhu Johnson 4 , Binita M. Kamath 1 ; 1 Gastroenterology,<br />

Hepatology and Nutrition, The Hospital for Sick Chidlren,<br />

Toronto, ON, Canada; 2 Computer Science, Research Institute - The<br />

Hospital for Sick Children, Toronto, ON, Canada; 3 Rheumatology,<br />

The Hospital for Sick Children, Toronto, ON, Canada; 4 Rheumatology,<br />

Toronto Western Hospital, Toronto, ON, Canada<br />

Background:Whether autoimmune hepatitis(AIH), originally<br />

“lupoid hepatitis”, and lupus(SLE)-associated hepatitis correspond<br />

to two distinct entities or are part of the same disease<br />

spectrum is controversial . Both are diagnosed based on<br />

expert-consensus and complex constellations of clinical, serological<br />

and histological features. There is the potential for an<br />

inherent flaw in defining these diseases based on human-driven<br />

criteria. We performed a rigorous hypothesis-generating study<br />

that removes patients’ current diagnostic labels and clusters<br />

them based on their broad clinical data. Aim:To determine if<br />

distinct disease entities can be objectively determined from a<br />

heterogeneous group of children with non-infectious, -toxic,<br />

-genetic causes of elevated liver enzymes. Methods:A retrospective<br />

cohort of pediatric patients(age


1044A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

grated clusters with diagnosis-based groups was 0.44(perfect<br />

similarity =1). Within the 8 groups, children with PSC were<br />

divided into 2. Those with AIH clustered across multiple groups<br />

that mostly did not include lupus patients. Conclusion:Our data<br />

suggests AIH and SLE-associated hepatitis exist as 2 distinct<br />

conditions that cluster differently with other causes of elevated<br />

liver enzymes. Identifying key differentiating groups of features<br />

that drive the similarities between children with elevated liver<br />

enzymes may assist in better future management and understanding<br />

of their disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Andreanne Benidir, Brian M.<br />

Feldman, Anna Goldenberg, Earl Silverman, Sindhu Johnson, Binita M. Kamath<br />

1714<br />

Screening for Mitochondrial Disease in Pediatric Acute<br />

Liver Failure: The Role of Lactate and Pyruvate<br />

Amy G. Feldman 1,5 , Ronald J. Sokol 1,5 , Regina M. Hardison 4 ,<br />

Estella M. Alonso 2,6 , Robert H. Squires 3,7 , Michael R. Narkewicz<br />

1,5 ; 1 Pediatric Gastroenterology, Hepatology and Nutrition,<br />

Children’s Hospital Colorado, Aurora, CO; 2 Department of Pediatrics,<br />

Ann and Robert H Lurie Children’s Hospital of Chicago,<br />

Chicago, IL; 3 Division of Gastroenterology, Children’s Hospital<br />

of Pittsburgh, Pittsburgh, PA; 4 Graduate School of Public Health,<br />

University of Pittsburgh, Pittsburgh, PA; 5 University of Colorado<br />

School of Medicine, Aurora, CO; 6 Northwestern University Feinberg<br />

School of Medicine, Chicago, IL; 7 University of Pittsburgh<br />

School of Medicine, Pittsburgh, PA<br />

Objectives: Mitochondrial disorders can cause pediatric acute<br />

liver failure (PALF). Screening for mitochondrial disorders using<br />

serum lactate and lactate/pyruvate (L:P) molar ratio has been<br />

recommended, yet the diagnostic utility of these tests in PALF<br />

has not been systemically analyzed. The goals of this study<br />

were to determine whether 1) the L:P ratio is a sensitive and<br />

specific test to screen for a mitochondrial, respiratory chain, or<br />

fatty acid oxidation (Mito) disorder in children with PALF and<br />

(2) the L:P ratio correlates with clinical outcome at 21 days.<br />

Study Design: A retrospective review was conducted of demographic,<br />

clinical, laboratory, and outcome data for patients in<br />

the NIH-supported PALF Study dataset who had lactate and<br />

pyruvate levels collected on the same day (within 7 days of<br />

enrollment) and a serum lactate level ≥2.5 mmol/L. The Kruskal-Wallis<br />

test and Pearson chi-square were used to compare<br />

continuous variables and proportions, respectively, among the<br />

diagnosis groups. Spearman correlations were used to estimate<br />

the association between L:P and clinical labs. Multinomial<br />

logistic regression was used to estimate the effect of L:P group<br />

on outcome. Results: Of 986 subjects, 110 had lactate and<br />

pyruvate levels collected on the same day, 74 (67 %) of whom<br />

had a lactate level ≥2.5 mmol/L. Of these 74 participants,<br />

39 (53%) were male; median age was 2.0 years (IQR 0.36-<br />

5.9). Six (8%) had a Mito disorder, 22 (29.7%) had another<br />

confirmed diagnosis, and 46 (62.2%) had an indeterminate<br />

diagnosis. There were no significant differences in the lactate<br />

levels, pyruvate levels, or L:P ratios between these three groups.<br />

Although 6 Mito group subjects had lactate ≥2.5mmol/L, only<br />

2 had L:P ratio ≥25. The percent with elevated L:P ratio≥ 25 did<br />

not differ by diagnostic group (33% Mito, 64% other diagnosis<br />

and 54% indeterminate diagnosis). There was no significant<br />

relationship between the L:P ratio and the AST/ALT ratio, INR,<br />

ammonia, creatinine kinase, acylcarnitine profile, or glucose<br />

levels nor was there a significant relationship between initial<br />

L:P ratio and clinical outcomes at 21 days (alive with native<br />

liver, alive with transplanted liver, dead). Conclusions: A high<br />

serum lactate and/or elevated L:P ratio were common in all<br />

causes of PALF, were not diagnostic for a mito cause, and did<br />

not predict 21 day clinical outcome. This suggests that either<br />

mito dysfunction is common in all causes of PALF, that mito<br />

disorders are misdiagnosed, or that L:P ratio is not specific for<br />

mito disorders. Patient age and clinical features remain the<br />

most helpful diagnostic clues for mito disorders, which must be<br />

confirmed by genetotyping or tissue enzymology.<br />

Disclosures:<br />

Ronald J. Sokol - Advisory Committees or Review Panels: Yasoo Health, Inc.; Consulting:<br />

Roche, Ikaria, Otsuka American Pharmaceuticals, Alnylam, Retrophin;<br />

Grant/Research Support: Mead Johnson Nutritionals, Lumena, FFF Enterprises<br />

Michael R. Narkewicz - Consulting: Vertex; Grant/Research Support: Novartis,<br />

Vertex; Stock Shareholder: Merck<br />

The following authors have nothing to disclose: Amy G. Feldman, Regina M.<br />

Hardison, Estella M. Alonso, Robert H. Squires<br />

1715<br />

Nocturnal Hypoxia is associated with Hedgehog Signaling<br />

and Severity of Pediatric Non-Alcoholic Fatty Liver<br />

Disease (NAFLD)<br />

Shikha Sundaram 1 , Marzena Swiderska-Syn 2 , Ronald J. Sokol 1 ,<br />

Zhaoxing Pan 1 , Ann C. Halbower 1 , Kelley E. Capocelli 1 , Kristen<br />

N. Robbins 1 , Anna Mae Diehl 2 ; 1 Children’s Hospital Colorado,<br />

Aurora, CO; 2 Duke University, Durham, NC<br />

Chronic intermittent nocturnal hypoxia is associated with<br />

NAFLD progression. Deregulation of the Hedgehog (Hh) pathway<br />

has also been implicated in NAFLD pathogenesis and<br />

progression. Objective: To determine relationships between<br />

obstructive sleep apnea (OSA), hypoxia, and Hh signaling in<br />

pediatric NAFLD. Methods: 31 adolescents with biopsy-proven<br />

NAFLD (mean age: 13 ± 1.9 yrs; mean BMI z score: 2.2 ±<br />

0.3, 65% male, 87% Hispanic) underwent polysomnogram,<br />

liver histology scoring, and laboratory testing. Sonic hedgehog<br />

(SHh), Indian Hh (IHh), Glioblastoma associated oncogene<br />

2 (Gli2), and markers of progenitors (keratin 7- K7) and<br />

myofibroblasts (α-SMA) were evaluated by immunohistochemistry.<br />

Results: Subjects with (68%) and without (32%) OSA/<br />

hypoxia had similar aminotransferases, serum lipids, inflammatory<br />

and insulin resistance markers. As expected, α-SMA<br />

generally correlated with fibrosis stage (r=0.44, p=0.02). Steatosis<br />

was inversely correlated with IHh (r=-0.48, p=0.009)<br />

and the NAS score correlated with SHh (r=0.39, p=0.03) and<br />

α- SMA (r=0.42, p=0.02). SHh correlated with AST (r=0.64,<br />

p=0.0001) and ALT (r=0.51, p=0.003). Gli2 also correlated<br />

with AST (r=0.47, p=0.007) and ALT (r=0.43, p=0.02). AST<br />

correlated with K7 (r=0.45, p=0.01), and ALT with α-SMA<br />

(r=0.51, p=0.005). Subjects with OSA/hypoxia had higher<br />

mean Apnea Hypopnea Index (AHI), % time oxygen saturation<br />

(SaO 2<br />

)


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1045A<br />

and relates to NAFLD disease severity. We speculate that tissue<br />

hypoxia allows for functional activation of Hypoxia Inducible<br />

Factor 1α with subsequent induction of genes important in<br />

epithelial-mesenchymal transition, including SHh, and NAFLD<br />

progression.<br />

Disclosures:<br />

Ronald J. Sokol - Advisory Committees or Review Panels: Yasoo Health, Inc.; Consulting:<br />

Roche, Ikaria, Otsuka American Pharmaceuticals, Alnylam, Retrophin;<br />

Grant/Research Support: Mead Johnson Nutritionals, Lumena, FFF Enterprises<br />

The following authors have nothing to disclose: Shikha Sundaram, Marzena<br />

Swiderska-Syn, Zhaoxing Pan, Ann C. Halbower, Kelley E. Capocelli, Kristen N.<br />

Robbins, Anna Mae Diehl<br />

1716<br />

Risk Factors Associated with Cystic Fibrosis-associated<br />

Liver Disease (CFLD)<br />

Tamir A. Miloh 1,2 , Amanda Pope 1 , John Daye 3 , Chengcheng Hu 3 ,<br />

Peggy Radford 1 ; 1 Phoenix Children, Phoenix, AZ; 2 Mayo Clinic,<br />

Scottsdale, AZ; 3 University of Arizona, Tucson, AZ<br />

Cystic fibrosis-associated liver disease (CFLD) affects approximately<br />

30% of patients and is a significant cause of morbidity<br />

and mortality. Aims: Analyze data from the CF registry to<br />

determine the predictors significant towards the development<br />

and progression of CFLD. We analyzed 39642 unique and<br />

valid observations documented on routine annual visits from<br />

1986 to present. Categorical, quantitative and ordinal variables<br />

were analyzed using Pearson’s Chi-squared test, 2-group<br />

t-test with Welsh degree of freedom modification and the<br />

independent 2-group Mann-Whitney U-test, respectively. The<br />

following phenotypes were further analyzed (n and %): any<br />

liver disease (11,002, 28%), gallstones (480, 1.7%), cirrhosis<br />

(1735, 4.5%), steatosis (150, 0.5%), non-cirrhotic liver disease<br />

(3513, 9%) and other (5124, 13.2%). Complications of cirrhosis<br />

(n and % of patients with cirrhosis) include gastric or esophageal<br />

varices (191, 22.9%), gastrointestinal bleeding (292,<br />

22.3%), splenomegaly and hypersplenism (302, 36%) and<br />

ascites (82, 9.8%). Variables were statistically significant at<br />

the p-valueG (54) and 3876delA (8.4).<br />

The following reduced risk of CFLD: Caucasian (0.65) and<br />

diagnosis on newborn screening (NS) (0.4). (b) Gallstones: FTT<br />

(1.43), steatorrhea (1.73), respiratory infection with pseudomonas<br />

(2.79), burkholderia cepacia (1.7), fungal (1.84). Protective:<br />

male gender (0.6) and NS (0.17). (c) Cirrhosis: male<br />

gender (1.4), FTT (1.43), meconium ileus (1.75), steatorrhea<br />

(1.33), respiratory infection with staphylococcal (1.4), pseudomonas<br />

(3.35), burkholderia cepacia (1.65), fungal (1.7) and<br />

F508 mutations (1.44). Mean chloride sweat concentration<br />

was 5.62 mmol/L higher than those without cirrhosis. Protective:<br />

NS (0.28) and mutations: R117H (0.23), 2789+5G>A<br />

(0.05) and 3849+10kbC>T (0.16). (d) Steatosis: mutations<br />

Q890X (26.66) and 218delA (8.86). The mean age of death<br />

for patients with cirrhosis was 2.24 years less than those without<br />

cirrhosis. Conclusion: Analyzing the largest CF database<br />

revealed that CFLD leads to earlier mortality and male gender,<br />

non-Caucasian, meconium ileus, malnutrition, increased sweat<br />

chloride, certain respiratory infections and mutation were associated<br />

with increased risk of CFLD and other mutations were<br />

protective.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tamir A. Miloh, Amanda Pope,<br />

John Daye, Chengcheng Hu, Peggy Radford<br />

1717<br />

Hepatic Manifestations in Adenosine Deaminase-Deficient<br />

Severe Combined Immune Deficiency<br />

Shilpa Lingala 1 , Elizabeth Garabedian 2 , Fabio Candotti 2 , David<br />

E. Kleiner 4 , Kenneth J. Wilkins 3 , Theo Heller 1 , Christopher Koh 1 ;<br />

1 Liver Diseases Branch, National Institute of Diabetes and Digestive<br />

and Kidney Diseases, NIH, Bethesda, MD; 2 National Human<br />

Genome Research Institute, NIH, Bethesda, MD; 3 Office of the<br />

Director, National Institute of Diabetes, Digestive and Kidney Diseases,<br />

Bethesda, MD; 4 Laboratory of Pathology, National Cancer<br />

Institute, Bethesda, MD<br />

Background: Adenosine deaminase-deficient severe combined<br />

immune deficiency (ADA-SCID) is a rare autosomal recessive<br />

metabolic disorder that occurs in


1046A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1718<br />

Autosomal Recessive Polycystic Kidney Disease: Not just<br />

a Renal Disease<br />

Sarika Rohatgi 1 , William E Sweeney 2 , Emma Schwasinger 2 , Nicholas<br />

Kampa 2 , Ellis D. Avner 2 ; 1 Pediatric Gastroenterology, Medical<br />

college of Wisconsin, Milwaukee, WI; 2 Pediatric Nephrology,<br />

Medical college of Wisconsin, Wauwatosa, WI<br />

Introduction: Autosomal Recessive Polycystic Kidney Disease/<br />

Congenital Hepatic Fibrosis (ARPKD/CHF) is an inherited hepatorenal<br />

fibrocystic disease with a wide spectrum of dual organ<br />

developmental abnormalities, secondary to PKHD1 mutations.<br />

Renal pathophysiology of ARPKD/CHF is well understood and<br />

clinical trials targeting renal disease are near. CHF is not well<br />

studied and warrants investigation due independent disease<br />

progression in the two organs. We hypothesize that proliferation<br />

and fibrosis are two separate but overlapping processes<br />

and identification of the ‘Critical stage’ of disease is important<br />

to achieve goal therapeutic results in CHF. Aim and Methods:<br />

PCK rat, an orthologous model with PKHD1 mutation, consistently<br />

develops dual organ abnormalities as in human disease,<br />

providing the best model for investigation. We performed a<br />

systematic characterization of morphological changes in development<br />

and progression of CHF in PCK Rats, correlating them<br />

with biliary ductal ectasia, proliferation, apoptosis and periportal<br />

fibrosis at progressive stages of disease. Our goal was to<br />

identify ‘critical stages’ of the disease to develop liver specific<br />

therapies. Five male rats each at 0, 30, 60, 90, 120 and<br />

150 days of age were evaluated and physiologic data (body,<br />

liver, kidney, intestine and spleen weights) collected. Serial<br />

sections of these livers were stained with Masson’s Trichrome<br />

stain to score fibrosis and immunohistochemistry (IHC) was<br />

done to identify cholangiocytes, proliferation and apoptosis.<br />

Morphometric analysis was done using ImageJ to quantitate<br />

disease progression. Results: Liver weight to body weight ratios<br />

increased as disease progressed. Morphometric analyses of<br />

liver tissues show consistent enlargement and deformation of<br />

biliary ducts with increasing age. A 50% increase in cyst numbers<br />

and cyst area per surface area of the liver with increased<br />

proliferation of cholangiocytes and mesenchymal cells was<br />

observed between ages of 30 and 90 days. Thereafter, proliferation<br />

gradually decreased to negligible by day 150. Periportal<br />

fibrosis index increased to 39% at day 90 and to 47%<br />

at 150 days. Fibrosis score increased from 0 to 2 between<br />

30 and 90 days and to 4 at 150 days. Fibrosis was more<br />

prominent in hilum around larger cysts. Apoptosis increased<br />

steadily with age. Conclusion: Rapid increase in proliferation<br />

and fibrosis at day 90 and decrease thereafter suggest that<br />

‘critical stage’ of the disease is before 90 days. Given the wide<br />

variability in rate of disease progression in CHF, therapies<br />

aimed at reducing proliferation will require a histopathological<br />

assessment of the liver disease to provide maximum benefit.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sarika Rohatgi, William E Sweeney,<br />

Emma Schwasinger, Nicholas Kampa, Ellis D. Avner<br />

1719<br />

Recurrent recessive mutation in DGUOK causes non-cirrhotic<br />

intrahepatic portal hypertension in the absence of<br />

liver synthetic dysfunction<br />

Silvia Vilarinho 4,1 , Sinan Sari 2 , Guldal Yilmaz 3 , Buket Dalgic 2 ,<br />

Richard P. Lifton 1,5 ; 1 Department of Genetics, Yale University<br />

School of Medicine, New Haven, CT; 2 Department of Pediatrics,<br />

Gazi University - Faculty of Medicine, Ankara, Turkey; 3 Department<br />

of Pathology, Gazi University - Faculty of Medicine, Ankara,<br />

Turkey; 4 Department of Internal Medicine, Section of Digestive<br />

Diseases, Yale University School of Medicine, New Haven, CT;<br />

5 Department of Internal Medicine, Yale University School of Medicine,<br />

New Haven, CT<br />

Liver cirrhosis is the most frequent cause of portal hypertension,<br />

but a variety of other diseases may cause elevated pressure<br />

in the portal venous system. However, when all known causes<br />

of portal hypertension have been ruled out, the diagnosis of<br />

idiopathic non-cirrhotic intrahepatic portal hypertension is<br />

established. In the literature, the terminology of this medical<br />

condition is ambiguous and several other terms such as hepatoportal<br />

sclerosis, obliterative venopathy, idiopathic portal<br />

hypertension, incomplete septal cirrhosis and nodular regenerative<br />

hyperplasia are used interchangeably. Despite progress<br />

in the diagnosis and management of this clinical entity,<br />

its etiopathogenesis remains elusive. Interestingly, idiopathic<br />

non-cirrhotic intrahepatic portal hypertension has also been<br />

reported in infancy and childhood, occasionally affecting more<br />

than one family member, suggesting that an unrecognized<br />

Mendelian trait may underpin the onset and development of<br />

portal hypertension in a subset of these subjects. To investigate<br />

this hypothesis, we applied unbiased whole-exome capture<br />

and DNA sequencing, in contrast to a hypothesis-driven<br />

approach, to nine Turkish subjects with onset of intrahepatic<br />

portal hypertension of indeterminate etiology during infancy<br />

or early childhood. Whole-exome sequencing allows us to<br />

determine the existence of recurrent mutations across samples<br />

and to assess the burden of rare variants in individual gene(s)<br />

greater than expected by chance. We found three individuals,<br />

in two unrelated families, that shared an identical rare<br />

homozygous mutation in deoxyguanosine kinase, encoded by<br />

the DGUOK nuclear gene, at a highly conserved amino acid<br />

position (p.Asn46Ser). The probability of finding two unrelated<br />

instances of this identical homozygous variant in DGUOK by<br />

chance is conservatively estimated to be 5.2x10 -12 , providing<br />

extremely strong statistical support for the role of this mutation<br />

in non-cirrhotic intrahepatic portal hypertension. DGUOK is an<br />

enzyme responsible for phosphorylation of purine deoxyribonucleosides<br />

in the mitochondrial matrix and its dysfunction has<br />

been previously associated with mitochondrial depletion syndrome<br />

type 3. In contrast to this syndrome, none of our index<br />

cases demonstrated signs of hepatic synthetic dysfunction for<br />

a follow-up period of 6 to 17 years. In conclusion, our study<br />

provides a new insight into the mechanisms mediating non-cirrhotic<br />

intrahepatic portal hypertension, defines a previously<br />

unrecognized form resulting from recessive N46S mutation in<br />

DGUOK, and expands the phenotypic spectrum of DGUOK<br />

deficiency beyond our current understanding.<br />

Disclosures:<br />

The following authors have nothing to disclose: Silvia Vilarinho, Sinan Sari,<br />

Guldal Yilmaz, Buket Dalgic, Richard P. Lifton


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1047A<br />

1720<br />

Liver Involvement in Turner’s Syndrome, Is Associated<br />

with Extra Hepatic Disease Manifestations<br />

Ohad Etzion 1 , Theo Heller 1 , Ma Ai Thanda Han 1 , Ruchi Patel 1 ,<br />

Christopher Koh 1 , David E. Kleiner 1,2 , Omar Mousa 1 , Rohit<br />

Loomba 1 , Vladimir Bakalov 2 , Carolyn Bondy 2 ; 1 Liver Disease<br />

Branch, NIH, Rockville, MD; 2 Heritable Disorders Branch, NICHID-<br />

NIH, Bethesda, MD<br />

Introduction: Turner’s syndrome (TS) is a genetic disorder<br />

caused by complete, partial lack or structural abnormalities<br />

of the X chromosome. Liver abnormalities are commonly<br />

described in patients with TS, however, little is known about<br />

its association with other clinical features of this syndrome.<br />

Aim: To characterize the spectrum of hepatic involvement and<br />

its correlation with non-invasive biomarkers and systemic manifestations<br />

of disease, in a cohort of patient with TS. Methods:<br />

From 2004-2009, patients with TS followed at the NIH Clinical<br />

Center, were evaluated for suspected liver disease. Biochemical<br />

tests, imaging <strong>studies</strong> and liver biopsy were performed<br />

as clinically indicated, and assessed for its association with<br />

documented extra hepatic disease manifestations. Results: 133<br />

patients (median age 43 years) were evaluated during this<br />

period. Abnormal ALT, AST and ALP values were found in 47%,<br />

40% and 21% of subjects respectively. Sonographic signs of<br />

steatosis were evident in 42% of patients. Echocardiography<br />

detected congenital heart defects in 42%, major venous anomalies<br />

in 13.4%, and aortic or major arterial anomalies in 8.9%<br />

of subjects. Elevated ALT was associated with presence of<br />

major venous anomalies (p=0.01), but not with arterial vascular<br />

anomalies or congenital heart defects. Association between<br />

hepatic steatosis on ultrasound and major venous anomalies<br />

trended towards statistical significance (p=0.053). Twenty-four<br />

patients underwent liver biopsy. Only 1 biopsy was staged as<br />

Ishak 3, with the rest showing Ishak scores of 0 or 1. Steatosis<br />

was diagnosed in 11 (46%) cases, and nodular regenerative<br />

hyperplasia (NRH) in 6 (25%). Portal tracts with missing bile<br />

ducts, arteries or veins were observed in 78%, 69% and 65%<br />

of biopsies respectively. Presence of a major artery anomaly<br />

was associated with higher rate of missing veins in the portal<br />

tracts (p=0.02), but not with bile duct or artery dropout.<br />

Congenital heart defects, major venous anomalies and other<br />

clinical features such as webbed neck and horseshoe kidney<br />

were associated with neither liver enzyme abnormalities, nor<br />

with liver histological findings. Conclusion: Liver enzyme abnormalities<br />

are common in TS and may reflect unique patterns of<br />

liver injury. As previously reported, steatosis is a dominant form<br />

of liver injury in TS. Abnormalities in hepatic microvasculature<br />

are a common finding in TS that are associated with systemic<br />

vascular anomalies. These findings suggest that liver vascular<br />

changes in TS represent an organ specific manifestation of a<br />

systemic disorder of vascular development, which evolves from<br />

the underlying chromosomal abnormality.<br />

Disclosures:<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

The following authors have nothing to disclose: Ohad Etzion, Theo Heller, Ma<br />

Ai Thanda Han, Ruchi Patel, Christopher Koh, David E. Kleiner, Omar Mousa,<br />

Vladimir Bakalov, Carolyn Bondy<br />

1721<br />

NOTCH2 variants in cholestatic liver disease<br />

Tassos Grammatikopoulos 1 , S. Strautnieks 2 , Melissa Sambrotta 1 ,<br />

Pierre Foskett 2 , Maesha Deheragoda 2 , A. S. Knisely 2 , Richard J.<br />

Thompson 1 ; 1 Institute of Liver Studies, Division of Transplantation<br />

Immunology and Mucosal Biology, King’s College London, London,<br />

United Kingdom; 2 Institute of Liver Studies, King’s College<br />

London, London, United Kingdom<br />

Background: Alagille syndrome (AGS) is an autosomal dominant<br />

multisystem disorder associated with mutations in JAG1<br />

and NOTCH2 in respectively 94% and ~1% of affected<br />

patients. Aim: We report eight cholestatic patients with novel,<br />

or rare NOTCH2 variants. Methods: After biliary atresia and<br />

α-1-antitrypsin deficiency were ruled out, cholestatic patients<br />

in whom clinical features and/or liver histology findings could<br />

not exclude AGS were assessed. NOTCH2 screening was<br />

performed as part of diagnostic next generation sequencing.<br />

Results: Among 302 patients sequenced using a custom-designed<br />

cholestasis gene panel, heterozygous variants in<br />

NOTCH2 were identified in 8 (5 male) patients. Seven variants<br />

were missense changes. One was protein-truncating (Table).<br />

Median age at presentation was 5 weeks (range, 2wks-30yrs)<br />

with jaundice (7), hepatosplenomegaly (1), failure to thrive (4),<br />

pale stools (5), raised transaminases (7) and pruritus (2). Vertebral<br />

and ophthalmological and cardiac abnormalities were not<br />

found in screened patients. Facial characteristics typical of AGS<br />

were identified in patient 1 and in his maternal grandmother<br />

who had the same genetic variant. Liver microscopy showed<br />

lobular cholestasis (5), bile duct paucity (2), and fibrosis (3);<br />

cytokeratin 7, expressed in attenuated biliary radicles, was<br />

aberrantly expressed in hepatocytes (2). Abdominal sonography<br />

and magnetic resonance cholangiography showed bile<br />

duct irregularities (2), abnormal gallbladder (1) and renal cysts<br />

(1). Median serum bilirubin was 105 μmol/L [5-227], AST<br />

86 IU/L [31-319], ALT 90 IU/L [33-195], GGT 172 IU/L [27-<br />

389], albumin 38 g/L [35-30], creatinine 43 μmol/L [13-89],<br />

INR 0.98 [0.94-1.1] and cholesterol 3.89 mmol/L [1.9-5.1].<br />

Patients 2 and 6 underwent liver transplantation before their<br />

1 st birthdays. Family histories included jaundice and gallstones<br />

in maternal grandparents in patient 3, heart murmur in mother<br />

of patient 2 and similar findings on liver biopsy in father of<br />

patient 6. Conclusion: We report the identification of eight<br />

probands with cholestasis and variants in NOTCH2. We could<br />

find clinical features associated with AGS in only a minority of<br />

these patients. None met diagnostic criteria for AGS.<br />

Disclosures:<br />

Richard J. Thompson - Grant/Research Support: Shire; Speaking and Teaching:<br />

Shire<br />

The following authors have nothing to disclose: Tassos Grammatikopoulos,<br />

S. Strautnieks, Melissa Sambrotta, Pierre Foskett, Maesha Deheragoda, A. S.<br />

Knisely


1048A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1722<br />

Liver Disease in Subjects with N-Glycanse-1 Deficiency<br />

Shilpa Lingala 1 , David E. Kleiner 6 , Christina Lam 2 , Lynne A.<br />

Wolfe 3,5 , Carlos R. Ferreira 2 , Donna Krasnewich 4 , William A.<br />

Gahl 2,5 , Marc G. Ghany 1 ; 1 Liver Diseases Branch, National Institute<br />

of Diabetes and Digestive and Kidney Diseases, National<br />

Institutes of Health, Bethesda, MD, Betheda, MD; 2 Medical Genetics<br />

Branch, National Human Genome Research Institute, NIH,<br />

Bethesda, MD; 3 Undiagnosed Diseases Program, NIH, Bethesda,<br />

MD; 4 Division of Genetics and Developmental Biology,, National<br />

Institute of General Medical Sciences, NIH, Bethesda, MD; 5 Office<br />

of the Clinical Director, National Human Genome Research Institute,<br />

NIH, Bethesda, MD; 6 Laboratory of Pathology, National Cancer<br />

Institute, Bethesda, MD<br />

Background: N-glycanase 1 (NGLY1) deficiency is a newly<br />

described inherited disorder of glycoprotein degradation.<br />

N-glycanase is a cytosolic enzyme that cleaves intact glycans<br />

off of misfolded N-linked glycoproteins after they have been<br />

processed by the endoplasmic reticulum associated degradation<br />

pathway. Deficiency of NGLY1 in humans manifests as<br />

developmental delay, multifocal epilepsy, involuntary movements,<br />

low tear production and elevation of liver-associated<br />

enzymes. Aim: To describe the spectrum of liver disease in<br />

subjects with NGLY1 deficiency. Methods: 11of 27 subjects<br />

with confirmed NGLY1 deficiency worldwide were evaluated<br />

at the Clinical Center, NIH between June 2014-May 2015.<br />

Some have been previously reported. A careful history and<br />

physical exam were performed and all available retrospective<br />

clinical data including liver biopsy were reviewed. At the Clinical<br />

Center, subjects underwent a comprehensive evaluation<br />

for liver disease including liver-associated enzymes, serologic<br />

testing for known causes of chronic liver disease, abdominal<br />

ultrasound and transient elastography to assess liver stiffness.<br />

Liver biopsies, if available, were evaluated by an expert hepatopathologist.<br />

Results: There were 6 females and 5 males with<br />

a mean age 9.9 years (2.4-21 yrs.). All were Caucasian.<br />

6/11(55%) had a history of neonatal jaundice, 50% of whom<br />

required phototherapy. 8 subjects had ALT levels tested within<br />

the first 2 years of age, mean ALT 141 U/L (12-1000 U/L). ALT<br />

levels were elevated in 7/8 (87%) and remained persistently<br />

elevated in only 2 (51 & 137 U/L) over a mean follow-up of<br />

2.5 years. Alkaline phosphatase was elevated for age in 7 subjects.<br />

Other causes of chronic liver disease were excluded in all<br />

11 subjects. Liver tissue was available on 3/11 and revealed<br />

cirrhosis in 2 and bridging fibrosis in one subject. Mild to<br />

moderate steatosis was seen in all biopsies. Scattered PAS<br />

positive macrophages were seen in biopsies from 2 subjects<br />

but no inclusion bodies were noted. Transient elastography<br />

stiffness scores were elevated (17.1, 9.5 and 8.1 Kpa) in the<br />

3 subjects with advanced fibrosis but normal in the remainder.<br />

One subject with cirrhosis developed possible HCC at age 1.2<br />

years and underwent liver transplantation. There was no noted<br />

association of liver disease progression with specific genotype.<br />

Conclusion: Chronic liver disease was common among subjects<br />

with NGLY1 deficiency, particularly during infancy, but<br />

liver-associated enzymes improved over time reminiscent of<br />

alpha-1 antitrypsin deficiency. Progression of liver disease to<br />

cirrhosis was observed and HCC may occur. The pathogenesis<br />

of the liver disease is currently being evaluated.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shilpa Lingala, David E. Kleiner,<br />

Christina Lam, Lynne A. Wolfe, Carlos R. Ferreira, Donna Krasnewich, William<br />

A. Gahl, Marc G. Ghany<br />

1723<br />

Primary immunodeficiencies in cryptogenic acute liver<br />

failure in children: a single centre experience<br />

Rajeev Khanna 1 , Susan Height 3 , Kimberly Gilmour 2 , Nedim<br />

Hadzic 1 ; 1 Paediatric Liver unit, King’s College Hospital, London,<br />

United Kingdom; 2 Immunology, Great Ormond Street Hospital,<br />

London, United Kingdom; 3 Paediatric Haemato-Oncology Unit,<br />

King’s College Hospital, London, United Kingdom<br />

Background: Majority of paediatric acute liver failure (PALF)<br />

cases remain cryptogenic. Haemophagocytic lymphohistiocytosis<br />

(HLH) is an under-diagnosed clinical syndrome which is<br />

often associated with primary immunodeficiency disorders.<br />

HLH is occasionally observed in ALF. Aim: We aimed to identify<br />

primary immunodeficiency states associated with HLH,<br />

observed in PALF between January’99 and May’15. Methods:<br />

We retrospectively analysed our database for children with<br />

HLH and investigated their demographic, clinical, biochemical<br />

and immunological markers, management and outcome.<br />

We used standard international criteria for definition of PALF.<br />

Children with histological evidence of haemophagocytosis, or<br />

with deficient proteins involved in lymphocyte granule-mediated<br />

cytotoxic pathway including perforin, signaling lymphocyte-activating<br />

molecule-associated protein (SAP), granule<br />

release assay (CD107a), Munc-13-4, syntaxin 11, syntaxin<br />

binding protein-2 on flow cytometry were enrolled. Mutational<br />

analysis was performed for PRF1, UNC13D, STX11, STXBP2,<br />

RAB27A, XLP-1/SH2D1A, XIAP/BIRC4, CD27 and MAGT1<br />

genes. Their outcome was compared with historical controls<br />

with PALF from our centre. Results: We identified 16 children<br />

with HLH – 8 had histological evidence, whereas 10 had deficient<br />

proteins (6 perforin, 3 SAP and 1 Munc-13-4). Males<br />

(81%) predominated. Four (25%) belonged to consanguineous<br />

families. Ten (62%) were younger than 6 months. Jaundice,<br />

fever and encephalopathy were present in 12, 12 and 8 cases,<br />

while hepatomegaly, splenomegaly and lymphadenopathy<br />

were there in 15, 13 and 4 children, respectively. Median<br />

peak bilirubin, AST and INR were 158 micromol/L, 1870 IU/L<br />

and 3.47. Hyperferritinemia, hypertriglyceridemia and hypofibrinogenemia<br />

were present in 14, 8 and 5 children. Mutations<br />

were identified in PRF1, SAP/XLP-1 and RAB27A genes in<br />

7, 3 and 1 patients, respectively. All patients were managed<br />

as per standard intensive care therapy. Their clinical course<br />

was complicated by marrow failure, ventilatory requirement,<br />

systemic inflammatory response syndrome and multi-organ dysfunction<br />

syndrome in 13, 9, 6 and 5 patients, respectively. 2<br />

children underwent liver transplantation – one of them died few<br />

months later due to neurological events. One underwent bone<br />

marrow transplantation. 1-year mortality was 31% which was<br />

higher than historical cohort. Conclusion: Identification of this<br />

group of children with ALF is important for their prognostication<br />

and family screening. In PALF, diagnosis of HLH needs to<br />

be pursued by combination of biochemical, histological and<br />

immunological markers, particularly in young male children<br />

from consanguineous families.<br />

Disclosures:<br />

Nedim Hadzic - Consulting: Alnylam; Speaking and Teaching: Synageva<br />

The following authors have nothing to disclose: Rajeev Khanna, Susan Height,<br />

Kimberly Gilmour


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1049A<br />

1724<br />

Zinc monotherapy for young pediatric patients with presymptomatic<br />

Wilson disease: a two-center experience in<br />

Japan<br />

Keisuke Eda 1 , Itaru Iwama 2 , Takahito Takeuchi 1 , Yugo Takaki 1 ,<br />

Tatsuki Mizuochi 1 ; 1 Pediatrics, Kurume University School of Medicine,<br />

Kurume, Japan; 2 Pediatrics, Okinawa Chubu Hospital,<br />

Uruma, Japan<br />

Purpose: We previously reported that a reasonable goal in<br />

treating young pediatric patients with Wilson disease (WD)<br />

using zinc to be maintaining 24-hour urinary copper excretion<br />

between 1 and 3 μg/kg/day for the first 1-2 years (Mizuochi,<br />

et al. JPGN 2011;53:365). Here, we aimed at evaluating<br />

long-term efficacy and safety of zinc monotherapy for<br />

young pediatric patients with WD and establishing appropriate<br />

benchmarks of maintenance therapy. Methods: We performed<br />

a retro- and prospective study to examine 7 girls (median age<br />

5 years, range 4-8) who satisfied diagnostic criteria for WD<br />

and were treated solely with zinc acetate prior to symptom<br />

onset at Kurume University Hospital and Okinawa Chubu Hospital<br />

in Japan. No additional WD sequelas, such as jaundice,<br />

hepatomegaly, or neurologic abnormalities were noted. We<br />

monitored serum AST and ALT, nonceruloplasmin serum copper,<br />

and 24-hour urinary copper for 2-7 years after initiation<br />

of zinc monotherapy. Additional monitorings included white<br />

blood cell count, hemoglobin, platelet count, serum total bilirubin,<br />

albumin, iron, amylase, lipase, and prothrombin time, as<br />

well as 24-hour urinary zinc excretion. We performed abdominal<br />

ultrasonograpy and evaluated clinical WD manifestations,<br />

drug compliance, and adverse effects of zinc. The prescribed<br />

dosage of zinc acetate for patients ≤ 5 years 25 mg twice<br />

daily; for those children 6 years or older, the dose was 25 mg<br />

3 times daily. We increased the dosage of zinc if the patients<br />

had AST/ALT > 50 U/L, and decreased it if they had adverse<br />

effects of zinc such as iron-deficiency anemia or pancytopenia.<br />

Results: At the time of diagnosis, AST/ALT and 24-hour urinary<br />

copper were 207±182/292±211 U/L and 143±68 μg/day<br />

(7.2±3.5 μg/kg/day), respectively. All patients continued to<br />

take zinc without any evidence of zinc toxicity. None of our<br />

7 patients became clinically symptomatic. AST/ALT sharply<br />

decreased to 35±7/37±16 U/L at 1 year after treatment and<br />

was mostly maintained less than 50U/L for the remainder of<br />

the study (AST/ALT: 35±9/42±29 and 33±9/41±17 U/L at<br />

2 and 6 years after treatment, respectively). Twenty four-hour<br />

urinary copper decreased to 54±18 μg/day (2.3±0.6 μg/<br />

kg/day) at 1 year after treatment and was mostly maintained<br />

between 1 and 3 μg/kg/day for the remainder of the study<br />

(1.7±0.7 and 1.9±0.9 μg/kg/day at 2 and 6 years after treatment,<br />

respectively). Conclusions: Long-term zinc monotherapy<br />

for young pediatric patients with presymptomatic WD proved<br />

highly effective and safe. A reasonable goal in treating young<br />

children with WD using zinc appears to be maintaining both<br />

AST/ALT under 50 U/L and 24-hour urinary copper excretion<br />

between 1 and 3 μg/kg/day.<br />

Disclosures:<br />

The following authors have nothing to disclose: Keisuke Eda, Itaru Iwama, Takahito<br />

Takeuchi, Yugo Takaki, Tatsuki Mizuochi<br />

1725<br />

Congenital Lactic acidemia (CLA), Amino acidemia (AA),<br />

Urea cycle defects (UCD) and Organic acidemia (OA)<br />

presenting with hepatic dysfunction in sick infants and<br />

children: Diagnosis, Management and Outcome in a<br />

Pediatric hepatology unit at a specialized liver centre in<br />

India.<br />

Seema Alam, Vikrant Sood, Bikrant B. Lal, Dinesh Rawat; Pediatric<br />

Hepatology, ILBS, New Delhi, India<br />

CLA, AA, UCD, OA are a group of MLDs which present with life<br />

threatening illness and various grades of hepatic dysfunction.<br />

Early recognition and management is of utmost importance in<br />

countries with limited Newborn Screening Programme. Here<br />

we present diagnosis, management and outcome children with<br />

various causes of CLA, UCD and OA following an algorithmic<br />

approach (Fig 1). Lactate and pyruvate ratio cam differentiate<br />

between various congenital lactic academia. There were 7<br />

cases of CLA including 3 cases of Fructose 1, 6 biphosphatemia,<br />

1 case of respiratory chain defect (RCD) and one case of<br />

(PEPCK, phosphoeno pyruvate carboxykinase). The three cases<br />

of Fructose 1, 6 biphosphatemia were managed with fructose<br />

free diet and recovered completely. The infants with RCD (management<br />

with Carnitine, Coenzyme Q10, Riboflavine and low<br />

carbohydrate diet) and PEPCK (supportive management) died.<br />

A newborn presenting as neonatal liver failure and HCC was<br />

diagnosed to be Tyrosinemia type 1 and the parents refused<br />

the option of liver transplant and medical management. A 3<br />

years old girl with UCD (Partial Ornithine transcarbamyolase<br />

deficiency) presented as acute liver failure and completely<br />

recovered with ammonia scavengers and awaiting liver transplant.<br />

Another UCD presented with hepatic dysfunction at 2<br />

months of age and died of encephalopathy. The only newborn<br />

with hepatic dysfunction and Propionic academia improved<br />

from the metabolic acidosis and ketosis with appropriate diet<br />

and was sent home with the option of liver transplant in the<br />

plan. A detailed protocol and algorithmic approach can help<br />

in early diagnosis and management of these rare liver diseases<br />

which in turn improves their outcome.<br />

Figure 1. Protocol based approach to life threatnimg illness with<br />

hepatic dysfunction.<br />

Disclosures:<br />

The following authors have nothing to disclose: Seema Alam, Vikrant Sood,<br />

Bikrant B. Lal, Dinesh Rawat


1050A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1726<br />

Bile acid metabolism is altered in adolescents with type<br />

1 diabetes<br />

Zoehrer Evelyn 1 , Melanie Heckmann 1 , Elke Fröhlich-Reiterer 1 , Hildegard<br />

Jasser-Nitsche 1 , Günter Fauler 2 , Hubert Scharnagl 2 , Tatjana<br />

Stojakovic 2 , Jörg Jahnel 1 ; 1 Department of Paediatrics and<br />

Adolescent Medicine, Medical University Graz, Graz, Austria;<br />

2 Clinical Institute of Medical and Chemical Laboratory Diagnostics,<br />

Medical University Graz, Graz, Austria<br />

Objective: High levels of bile acids (BA) stimulate insulin release<br />

and therefore decrease serum glucose levels via the TGR5/<br />

GLP1 pathway; vice versa, high glucose levels can upregulate<br />

BA synthesis from cholesterol by stimulating transcription of the<br />

rate-limiting enzyme Cyp7a1. Type 1 diabetes (T1D) leads to<br />

hyperglycemia due to lack of endogenous insulin. We hypothesized<br />

that in children and adolescents with T1D BA levels<br />

vary coordinately with HbA1c, a longtime marker of glycemic<br />

control. Design: Concentrations of glucose, HbA1c, and serum<br />

total BA (tBA) were measured in 86 fasted children and adolescents<br />

with T1D (age 11 - 20 years) under insulin therapy.<br />

Patients were divided into three groups: Group 1: low HbA1c<br />

(< 59 mmol/mol; n=21); Group 2: medium HbA1c (59 – 75<br />

mmol/mol; n=49) and Group 3: high HbA1c (> 75 mmol/<br />

mol; n=16). TBA values were compared with normal values<br />

for adolescents (n=98; Jahnel et al. 2015, CCLM). Using 10<br />

μl of serum we determined by high-performance liquid chromatography<br />

– high-resolution mass spectrometry a BA profile<br />

including 15 unconjugated and taurine (T)- or glycine (G)-conjugated<br />

BA; summed, the values for these analytes yield the<br />

tBA value. Results: In Group 1, with low HbA1c values, mean<br />

tBA values lay below reference values with 3.1 mmol/l ± 2.4<br />

mmol/l (3.7 mmol/l ± 1.9 mmol/l). In Group 2, with normal<br />

HbA1c values, mean tBA values were significantly lower than<br />

in Group 1 and lay below normal ranges with 2.3 mmol/l ±<br />

1.8 mmol/l (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1051A<br />

1728<br />

Serum autoantibodies are associated with chronic hepatitis<br />

and graft fibrosis after pediatric liver transplantation<br />

Jeremy K. Rajanayagam 1 , Wolfram Haller 1 , Dominique Debray 2 ,<br />

Henkjan J. Verkade 3 , Valerie A. McLin 4 , Helen M. Evans 5 , Etienne<br />

M. Sokal 6 , Ekkehard Sturm 7 , Rene Scheenstra 3 , Annette S. Gouw 8 ,<br />

Vladimir Cousin 4 , Sharat Varma 9 , Steffen Hartleif 7 , James Hodson<br />

10 , Deirdre A. Kelly 1 ; 1 The Liver Unit, BIrmingham Children’s<br />

Hospital, Birmingham, United Kingdom; 2 Clinique chirurgicale<br />

Infantile, Necker Hospital, Paris, France; 3 Paediatric Hepatology,<br />

Beatrix Children’s Hospital, University Medical Center Groningen,<br />

Groningen, Netherlands; 4 Paediatrics, University Hospitals<br />

Geneva, Geneva, Switzerland; 5 Hepatology, Starship Children’s<br />

Hospital, Auckland, New Zealand; 6 Paediatric Hepatology and<br />

Cell Therapy Lab, Université Catholique de Louvain, Louvain,<br />

Belgium; 7 Paediatric Gastroenterology & Hepatology, University<br />

Children’s Hospital, Tuebingen, Germany; 8 Pathology and Medical<br />

Biology, University Medical Center Groningen, Groningen,<br />

Netherlands; 9 Cliniques universitaires St Luc, Brussels, Belgium;<br />

10 University Hospitals Birmingham, Birmingham, United Kingdom<br />

Background: Chronic hepatitis (CH) and fibrosis are frequent<br />

features of protocol liver biopsies after pediatric liver transplantation<br />

(LT). Previously reported associations with positive autoantibodies<br />

(AuAb) and abnormal immunoglobulins (Ig) suggest<br />

immunologic mechanisms of graft injury. Aim: To evaluate the<br />

role of AuAbs and Igs in the etiology of CH and fibrosis. Methods:<br />

International, multicenter, retrospective analysis of pediatric<br />

LT recipients with 5 and/or 10 yr protocol liver biopsies<br />

(normal aminotransaminases), paired serum AuAbs (ANA or<br />

SMA titre>1:40) and Ig levels, excluding those with pre-LT autoimmune<br />

liver disease, or other causes of graft injury. Results:<br />

467 children, from 7 centres underwent LT between 1985-<br />

2010, predominantly for biliary atresia (n=274), metabolic<br />

disease (n=58), other cholestatic disorders (n=53), acute liver<br />

failure (n=44), cryptogenic cirrhosis (n=20) and malignancy<br />

(n=7). CH was seen in 119/279 (43%) biopsies at 5yrs,<br />

and 121/229 (53%) biopsies at 10yrs. Fibrosis was present<br />

in 163/301 (54%) at 5yrs (mild=36%; moderate=12%;<br />

severe=6%) and 178/226 (79%) at 10yrs (mild=46%;, moderate=20%;<br />

and severe=13%). Children positive vs negative<br />

for AuAbs were more likely to exhibit not only CH (5yrs: 59%<br />

vs 34%, p


1052A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Kathleen B. Schwarz - Consulting: Novartis; Grant/Research Support: Bristol-Myers<br />

Squibb, Gilead, Roche/Genentech, Synageva, Vertex, Roche<br />

The following authors have nothing to disclose: Douglas Mogul, Eric Chow, Xun<br />

Luo, Allan Massie, Andrew M. Cameron, John F. Bridges, Dorry L. Segev<br />

Disclosures:<br />

Philip Rosenthal - Advisory Committees or Review Panels: Abbvie, Retrophin;<br />

Consulting: Gilead; Grant/Research Support: Gilead, Abbvie, Vertex, BMS,<br />

Roche/Genetech<br />

The following authors have nothing to disclose: Emily R. Perito, Robert H. Lustig<br />

1730<br />

Pediatric liver transplant recipients (PLTx) have a different<br />

form of pre-diabetes, with decreased β-cell function<br />

rather than insulin resistance<br />

Emily R. Perito 1,2 , Robert H. Lustig 1 , Philip Rosenthal 1,3 ; 1 Pediatrics,<br />

University of California San Francisco, San Francisco, CA; 2 Epidemiology<br />

and Biostatistics, University of California San Francisco,<br />

San Francisco, CA; 3 Surgery, University of California San Francisco,<br />

San Francisco, CA<br />

Introduction: PLTx have not been previously been evaluated<br />

for pre-diabetes (defined by the ADA as impaired glucose tolerance<br />

(IGT; ≥140mg/dL 2 hrs after glucose load) or high<br />

fasting glucose (>100mg/dL)), despite a known risk with calcineurin-inhibitors<br />

(CNI). Methods: Cross-sectional study of PLTx<br />

aged 8-30 yrs, on stable immunosuppression without diabetes,<br />

with oral glucose tolerance testing (OGTT) by NHANES protocol.<br />

Patients matched by age, sex, and race/ethnicity with<br />

≤4 controls from NHANES 2009-2012 cohorts. Results: PLTx<br />

(n=78) were studied at mean age 15.9 yrs (range 8.1-30.2)<br />

and mean 11.2 yrs since transplant (range 1-24). PLTx were<br />

more likely to have IGT (29%,n=72) than matched NHANES<br />

controls (6.5%,n=292, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1053A<br />

Disclosures:<br />

The following authors have nothing to disclose: Adam Arterbery, Awo Osafo-<br />

Addo, Maria Ciarleglio, Yanghong Deng, Mercedes Martinez, Steven J. Lobritto,<br />

Yaron Avitzur, David Hafler, Markus Kleinewietfeld, Udeme D. Ekong<br />

1732<br />

Distance Affects Mortality on the Waitlist in Pediatric<br />

Liver Transplantation<br />

Joel T. Adler, Yanik Bababekov, James F. Markmann, David C.<br />

Chang, Heidi Yeh; Massachusetts General Hospital, Boston, MA<br />

INTRODUCTION: The distance between patients and their liver<br />

transplant centers contributes to disparities in adult liver transplantation.<br />

We hypothesized that distance would adversely<br />

affect the time to transplantation and waitlist mortality for pediatric<br />

transplant candidates. METHODS: The Scientific Registry<br />

of Transplant Recipients was queried for isolated pediatric liver<br />

transplant registrants (under age 18) with chronic liver disease<br />

and valid ZIP code information from 2003 to 2012. Distance<br />

from home ZIP code to listing transplant center was calculated<br />

in miles. Competing events analysis, adjusted for demographic<br />

factors, rural/urban status, indication, and Pediatric End Stage<br />

Liver Disease score, was undertaken for transplantation and<br />

death while on the waitlist. RESULTS: 6,924 children were<br />

included. The median distance to listing transplant center was<br />

65 (IQR 17.5-189) miles. Median distance traveled increased<br />

by listing volume (73.9 vs. 33.8 miles, highest vs. lowest volume<br />

quartile, P


1054A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1734<br />

Dynamics of pediatric liver allograft histological<br />

changes and the role of HLA, non-HLA antibodies<br />

Sharat Varma 3 , Xavier Stephenne 3 , Françoise Smets 3 , Mina<br />

Komuta 1 , Jerome Ambroise 2 , Catherine de Magnée 4 , Dominique<br />

Latinne 5 , Raymond Reding 4 , Etienne M. Sokal 3 ; 1 Department of<br />

anatmopathology, Cliniques universitaires St Luc, Brussels, Belgium;<br />

2 Centre for Applied Molecular Technologies (CTMA), Institut<br />

de Recherche Expérimentale et Clinique (IREC),, Université<br />

Catholique de Louvain (UCL), Brussels, Belgium; 3 Service of Pediatric<br />

Gastroenterology and Hepatology, Cliniques universitaires<br />

St Luc, Brussels, Belgium; 4 Department of Pediatric Surgery and<br />

Transplantation, Cliniques universitaires St Luc, Brussels, Belgium;<br />

5 Clinical Transplant Immunology Department, Cliniques universitaires<br />

St Luc, Brussels, Belgium<br />

Study purpose:HLA and non-HLA antibodies are frequently<br />

seen post liver transplantation(LT) but their impact on the graft is<br />

unknown. We studied the impact of these in transplant recipient<br />

children with minimum confounding factors for causing fibrosis<br />

and histological changes. Methods: Retrospective study of children<br />

transplanted between 2004 - 2014 having known HLA<br />

and non-HLA antibody status and >1 protocol biopsy. Exclusion<br />

criteria were combined organ transplant, re-transplant,<br />

chronic rejection, vascular and biliary complications. Biopsies<br />

were evaluated for bile duct proliferation, portal tract inflammation<br />

and fibrosis by Metavir and Venturi score.Fibrosis rate<br />

was calculated as change in fibrosis score divided by duration<br />

(years) since previous biopsy. HLA antibodies were tested with<br />

solid single bead Luminex assay with MFI>1500 as cut off.<br />

Results: 103 children with 323 biopsies were included. Longer<br />

post LT duration was associated with higher grade of fibrosis<br />

(p 2000 (clinically relevant).<br />

We used the sum of class I and class II antibodies to calculate<br />

total MFI. Results: DSA status in patients with and without ACR<br />

are summarized in Table 1. Sixty-four patients had DSA measured<br />

sometime during follow up. Of these, 37 were positive<br />

for class I or class II at least once. Several patients presented<br />

with both pre-formed and de novo DSA. DSA were not associated<br />

with ACR during follow-up. In those patients with measurable<br />

DSA, MFI >2000 did not correlate with the occurrence<br />

of ACR. Discussion: In this small, single center retrospective<br />

study, 57% of children had measurable pre-formed or de novo<br />

DSA either pre- or post-LT. In spite of this finding, the presence<br />

of DSA at any time during the early follow up period was not<br />

associated with an ACR event. . MFI titer did not predict severity<br />

of rejection among patients who developed ACR.Among<br />

patients who experienced rejection, MFI titers did not correlate<br />

with severity of rejection.Prospective <strong>studies</strong> are required to<br />

understand better the relationship between DSA and ACR in<br />

pediatric LT.<br />

Table 1<br />

Disclosures:<br />

The following authors have nothing to disclose: Dominique Schluckebier, Laetitia<br />

Marie Petit, Vladimir Cousin, Anne-Laure Rougemont, Jean Villard, Dominique<br />

Belli, Barbara E. Wildhaber, Sylvie Ferrari-Lacraz, Valerie A. McLin


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1055A<br />

1736<br />

Surveillance biopsies in pediatric liver transplant recipients<br />

to guide immunosuppression (IS) reduction<br />

Amanda J. Posner 1 , Kuang-Yu Jen 3 , Linda Ferrell 3 , Anthony J.<br />

Demetris 5 , Sandy Feng 4 , Philip Rosenthal 1,4 , Emily R. Perito 1,2 ;<br />

1 Pediatrics, University of California San Francisco, San Francisco,<br />

CA; 2 Epidemiology and Biostatistics, University of California San<br />

Francisco, San Francisco, CA; 3 Pathology, University of California<br />

San Francisco, San Francisco, CA; 4 Surgery, University of California<br />

San Francisco, San Francisco, CA; 5 Pathology, University of<br />

Pittsburgh, Pittsburgh, PA<br />

Introduction: Recent AASLD guidelines on long-term management<br />

of pediatric liver transplant (LT) recipients recommend considering<br />

IS minimization in children with minimal inflammation<br />

and fibrosis on surveillance biopsy. We instituted surveillance<br />

biopsy and IS minimization protocols, following experience<br />

with Immune Tolerance Network (ITN) IS withdrawal trials.<br />

Methods: Single-center study of surveillance liver biopsies to<br />

assess eligibility for IS reduction (withdrawal in clinical trial,<br />

n=27; or minimization to 50% of baseline calcineurin-inhibitor<br />

[CNI] dose as standard of care [SOC], n=32) in subjects who<br />

underwent LT prior to age 18. At biopsy, all were ≥5 years<br />

post-LT on CNI (n=56) or CNI+mycophenolate (n=3). Biopsies<br />

reviewed by 1 of 2 pathologists using standardized criteria<br />

from Banff recommendations. Results: 59 pediatric LT recipients<br />

were included, at median (range) age 13.1 (5.1-22.1)<br />

years and 9.7 (3.6-20.5) years from LT. LT was whole liver in<br />

42%, split deceased-donor in 24%, and living-donor in 32%.<br />

Of 20 subjects with no fibrosis, 75% qualified for IS reduction.<br />

(FIG) Fibrosis was seen in 39 subjects, but 84% of portal fibrosis<br />

was Ishak stage 1-2 and 78% of perivenular fibrosis was<br />

mild. 38/59 (64%) subjects were eligible for IS withdrawal or<br />

SOC minimization. (FIG) Children eligible vs ineligible were<br />

younger at transplant (median [IQR] 0.8 [0.5-1.6] vs 1.6 [0.8-<br />

4.8] years, p=0.03) with no difference in years since transplant,<br />

whole vs split or living vs deceased donor, AR history,<br />

or AST/ALT/GGT at biopsy. Of 22 children eligible for SOC<br />

IS minimization, 5 completed and 5 are in process; none have<br />

had AR or complications. Conclusion: Surveillance biopsy is<br />

essential tp inform decision-making on IS reduction. Longitudinal<br />

<strong>studies</strong> are needed to establish benefits vs risks of biopsy-guided<br />

IS reduction.<br />

Surveillance biopsies and IS reduction eligibility in pediatric LT<br />

recipients<br />

Philip Rosenthal - Advisory Committees or Review Panels: Retrophin, Gilead; Consulting:<br />

Roche, Abbvie; Grant/Research Support: Roche, Bristol MyersSquibb,<br />

Gilead, Vertex, Abbvie<br />

The following authors have nothing to disclose: Amanda J. Posner, Kuang-Yu Jen,<br />

Linda Ferrell, Anthony J. Demetris, Emily R. Perito<br />

1737<br />

Epidemiology and outcomes of pediatric liver transplant<br />

recipients with multidrug resistant bacterial and fungal<br />

infections<br />

Gillian Noel, Mark J. Abzug, Donna Curtis, Zhaoxing Pan, Shikha<br />

Sundaram; University of Colorado/Children’s Hospital Colorado,<br />

Denver, CO<br />

Infection is a significant source of morbidity and mortality<br />

following liver transplantation. Adult transplant recipients<br />

experience increased colonization and infection with drug-resistant<br />

organisms including methicillin-resistant staphylococcus<br />

(MRSA), extended-spectrum beta-lactamase producers, vancomycin-resistant<br />

enterococci, and multidrug resistant pseudomonas<br />

aeruginosa. Limited data exist regarding the incidence of<br />

multidrug resistant organisms in the pediatric liver transplant<br />

(LT) population. Objective: To define the incidence and impact<br />

on outcomes of multidrug resistant organisms in pediatric LT<br />

recipients. Methods: A retrospective cohort study of pediatric LT<br />

recipients at Children’s Hospital Colorado between Jan 2006<br />

and Dec 2014 was conducted. Demographic, clinical and<br />

outcome data were collected. Results: Sixty subjects were identified<br />

and followed for one year post-LT (mean age at LT of 4.9<br />

± 6.1 years; 55% female, 67% white, 51% transplanted due<br />

to biliary atresia). Treatment for suspected bacterial infections<br />

occurred in 51% of subjects, but only 44% had a documented<br />

isolate. Major isolates included coagulase negative staphylococci,<br />

enterococci, and gram-negative enterics. Although<br />

MRSA colonization was common (6.5% MRSA positive at LT),<br />

MRSA infections did not occur. Two percent of subjects had<br />

multidrug resistant organisms (extended spectrum beta lactamase<br />

producers) identified. Of subjects treated for bacterial<br />

infections, 68% received 1 course, 18% received 2-3 courses,<br />

and 14% received 4 or more courses of antibiotics. Presumed<br />

cholangitis and bacteremia were the most common reasons for<br />

antibiotic treatment. Treatment for fungal infections occurred in<br />

5% of subjects, although no multidrug resistant fungal infections<br />

were identified. All treated subjects received one course of<br />

antifungal treatment. Of those treated, only 50% had a documented<br />

isolate, primarily candidal species. Fungemia was<br />

the most common reason for antifungal treatment. Subjects<br />

with clinically suspected bacterial or fungal infections post-LT<br />

had more ICU admissions (1-3) versus those without suspected<br />

infection (0) post-LT (p=0.05). There were no differences in<br />

overall numbers of hospitalizations, reoperation rates, rejection,<br />

graft loss, or death in the twelve months post-LT as related<br />

to infection status. Conclusions: While serious bacterial and<br />

fungal infections commonly occur in pediatric liver transplant<br />

recipients, multidrug resistant infections are uncommon in our<br />

institution. Infected subjects have more complicated courses,<br />

characterized by ICU admissions, but their overall outcomes<br />

are similar to those without infections.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gillian Noel, Mark J. Abzug,<br />

Donna Curtis, Zhaoxing Pan, Shikha Sundaram<br />

Disclosures:<br />

Sandy Feng - Consulting: Novartis


1056A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1738<br />

Immunization Practices Amongst Pediatric Transplant<br />

Hepatologists at SPLIT Centers<br />

Amy G. Feldman 1,4 , Shikha Sundaram 1,4 , Brenda Beaty 3 , Allison<br />

Kempe 2 ; 1 Pediatric Gastroenterology, Hepatology and Nutrition,<br />

Children’s Hospital Colorado, Aurora, CO; 2 General Pediatrics,<br />

Children’s Hospital Colorado, Aurora, CO; 3 Adult and Child Center<br />

for Health Outcomes Research and Delivery Science, Aurora,<br />

CO; 4 University of Colorado School of Medicine, Aurora, CO<br />

Objectives: To assess amongst US and Canadian pediatric<br />

transplant hepatologists: (1) current immunization practices<br />

before and after liver transplantation and (2) beliefs about the<br />

safety of different vaccines before and after liver transplantation.<br />

Study Design: An 80 item email survey of US and Canadian<br />

pediatric transplant hepatologists at centers participating<br />

in the Studies of Pediatric Liver Transplantation (SPLIT) consortium<br />

from December 2014 to March 2015. Results: The overall<br />

response rate was 80%, representing 97% of SPLIT centers.<br />

Thirty-four percent of programs did not have written protocols<br />

to guide immunizations before and after transplant and 50%<br />

of programs always involved an infectious disease physician<br />

as part of the formal liver transplant evaluation. Administration<br />

of Palivizumab was considered by 45% of hepatologists<br />

pre-transplant and by 64% post-transplant. Only 25% of hepatologists<br />

ever recommended live vaccines after transplant. Timing<br />

of inactive vaccines after transplant ranged from 1 month<br />

to greater than 6 months post-transplant. Only 6% of programs<br />

were able to provide all vaccines to patients in hepatology/<br />

transplant clinic pre and post-transplant. Similarly, 6% of programs<br />

were always able to provide influenza vaccine to family<br />

members/household contacts. Conclusions: There is wide variation<br />

in immunization practices before and after liver transplant<br />

amongst pediatric transplant hepatologists. Specific evidence<br />

based protocols are needed to guide immunization practices<br />

before and after transplant and ensure appropriate vaccination<br />

of pediatric liver transplant recipients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Amy G. Feldman, Shikha Sundaram,<br />

Brenda Beaty, Allison Kempe<br />

1739<br />

Detectable HBVDNA in liver allograft tissues has no clinical<br />

significance on long term follow up in liver transplant<br />

recipients under HBIg plus Antiviral prophylaxis<br />

Deniz Mut 1 , Ilker Turan 1 , Deniz Nart 2 , Funda Yilmaz 2 , Selda Erensoy<br />

1 , Ulus S. Akarca 1 , Galip Ersoz 1 , Fulya Gunsar 1 , Zeki Karasu 1 ;<br />

1 Department of Gastroenterology, Ege University, Izmir, Turkey;<br />

2 Pathology, Ege University, Izmir, Turkey<br />

Combination prophylaxis with hepatitis B immune globulin<br />

(HBIG) plus an oral anti-viral (OAV) is widely accepted regimen<br />

for the prevention of recurrent hepatitis B virus (HBV) infection<br />

after liver transplantation, although there is no consensus on<br />

dose of HBIg. Persistency of HBV infection in allograft tissue in<br />

spite of prophylaxis is speculative. We examined the presence<br />

of HBVDNA using real-time polymerase chain reaction (PCR) in<br />

liver allografts in liver transplant recipients who undergone liver<br />

transplantation because of HBV related liver disease. Patients<br />

who had HBIg+Lam prophylaxis and who have no detectable<br />

serum virologic markers (HBsAg and/or HBV DNA) for recurrent<br />

disease during follow up of at least 3-years were included<br />

study. Histologic examination is also performed. We investigated<br />

if there is any correlation between tissue HBVDNA and<br />

presence of pretransplant HCC, Delta co-infection, the type<br />

of graft donor (living/cadaveric), the type of post transplant<br />

immunosuppressive medication. Results: One-hundred and fifty<br />

(116 male, 34 female) patients with a mean age of 47-years<br />

(18-64 years) were included into the study. Median duration<br />

of follow up was 5 (3-9) years after transplantation. Sixty-four<br />

patient had delta co-infection and 44 patients had HCC before<br />

transplantation. Of the patients, 108 (72%) had their grafts<br />

from living donors. HBVDNA was detectable by PCR in liver<br />

tissue of 18 (12%) patients. Immunostaining for HBVwas negative<br />

on liver histology of all patients. Of the 18 patients with<br />

detectable HBVDNA by PCR, the liver graft showed no hepatitis<br />

on histologic examination. Detectable tissue HBVDNA had<br />

no correlation with age, sex, delta co-infection, HCC, type of<br />

donor and type of immunosuppressive medication. Conclusion:<br />

in spite of combination prophylaxis with HBIg+OAV, and of<br />

there is no detectable serologic markers in their serum, some<br />

(12%) cases may have detectable HBVDNA in their liver tissue.<br />

However presence of HBVDNA in grafts has no clinical impact.<br />

But in patients with detectable HBVDNA, the risk of reactivation<br />

may persist indefinitely.<br />

Disclosures:<br />

Ulus S. Akarca - Advisory Committees or Review Panels: GILEAD, BMS, MSD,<br />

AbbVie<br />

The following authors have nothing to disclose: Deniz Mut, Ilker Turan, Deniz<br />

Nart, Funda Yilmaz, Selda Erensoy, Galip Ersoz, Fulya Gunsar, Zeki Karasu<br />

1740<br />

HBV cccDNA evaluation in HBV-core antibody<br />

(HBcAb)-positive liver transplant donors<br />

Gian Paolo Caviglia 1 , Francesco Tandoi 2,3 , Maria Lorena Abate 1 ,<br />

Alessia Ciancio 1,4 , Renato Romagnoli 2,3 , Antonina Smedile 1,4 ;<br />

1 Medical Sciences, University of Turin, Turin, Italy; 2 Surgical Sciences,<br />

University of Turin, Turin, Italy; 3 Liver Transplant Center,<br />

Città della Salute e della Scienza, Molinette Hospital, Turin, Italy;<br />

4 Gastroenterology, Città della Salute e della Scienza, Molinette<br />

Hospital, Turin, Italy<br />

Introduction Occult hepatitis B virus infection (OBI) is defined<br />

as the presence of Hepatitis B Virus (HBV) DNA in the liver of<br />

subjects negative for hepatitis B surface antigen (HBsAg). A<br />

high OBI prevalence has been reported in the anti-HBV core<br />

antibody (HBcAb)-positive population as residue of a past<br />

infection. OBI-positivity may be responsible for recurrent or de<br />

novo hepatitis B in immunosuppressed subjects. In the setting<br />

of liver transplant, prophylaxis for hepatitis B recurrence is<br />

routinely performed in all HBcAb-positive liver graft recipients.<br />

However, the risk of HBV reactivation is linked to the presence<br />

of intrahepatic covalently-closed-circular DNA (cccDNA) being<br />

the replication-competent form. Aim To evaluate the presence<br />

of intrahepatic HBV cccDNA and its correlation with OBI status<br />

in a cohort of HBcAb-positive liver donors. Methods Liver<br />

biopsies from 78 consecutive HBsAg-negative/HBcAb-positive<br />

deceased donors (mean age 66 [31-91] years; M/F=50/28)<br />

were tested for OBI by 4 parallel nested-PCRs to detect HBV S,<br />

Core, Polymerase and X sequences. According to Taormina<br />

expert meeting statements, samples turned positive for at least<br />

2 targets were scored as OBI-positive. cccDNA presence was<br />

assessed by a specific nested-PCR after enzymatic digestion<br />

with S1 nuclease. Results Among HBcAb-positive liver donors,<br />

OBI prevalence was 56.4% (44/78). In detail, 28 liver biopsies<br />

were 4/4 targets positive, 7 were 3/4, 9 were 2/4, 11 were<br />

1/4 while 23 biopsies were completely negative. cccDNA<br />

prevalence was 32.1% (25/78) in the whole cohort and<br />

45.5% (25/55) in the OBI-positive group (including those positive<br />

for only 1 target). A significant direct correlation was found<br />

between OBI and cccDNA-positivity (r=0.594; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1057A<br />

ed-PCRs was significantly associated with cccDNA detection<br />

(OR=2.295, 95%CI 1.7574-4.5556; p10 ng/mL, 83% were male, and<br />

88% were non-Asian. Comorbidities included diabetes (29%),<br />

smoking (63%), HBV (2%), and HIV (12%). Median HCC size<br />

was 2.8 cm (range, 0.8-18.2), 83% had a single lesion on<br />

imaging and 51% had histologic vascular invasion. Median<br />

lab values at the time of HCC diagnosis indicated that liver<br />

function was generally well-preserved: albumin 4.2 g/dL (2.1-<br />

5.0), platelets 148 x 10 3 cells/μL (48-446), and total bilirubin<br />

1.3 mg/dL (0.2-8.8). HCC was diagnosed via surveillance<br />

(imaging and/or AFP) in 27 patients, and in response to symptoms<br />

in 13. In the surveillance group, 85% were within Milan<br />

criteria vs 23% in the symptomatic group (p


1058A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Chiara Rocha, M. Isabel Fiel, Erin<br />

H. Doyle, Nicolas Goossens, Yujin Hoshida, Myron E. Schwartz<br />

1743<br />

Treating HCV patients on the LT waiting list with Sofosbuvir<br />

plus Ribavirin is associated with clinical benefits<br />

and allows for safe assignement of older grafts to LT<br />

recipients<br />

Maria Francesca Donato 1 , Daniele E. Dondossola 2 , Federica Malinverno<br />

1 , Sara Monico 1 , Cristina Rigamonti 1 , Lucio Caccamo 2 ,<br />

Alessio Aghemo 1 , Margherita Cavenago 2 , Giovanna Lunghi 3 ,<br />

Barbara B. Antonelli 2 , Giorgio Rossi 2 , Massimo Colombo 1 ; 1 Division<br />

of Gastroenterology and Hepatology, Fondazione IRCCS<br />

Ca’ Granda Ospedale Maggiore Policlinico, University of Milan,<br />

Milan, Italy; 2 Division of Surgery and Liver Transplant, Fondazione<br />

IRCCS Ca’ Granda Ospedale Maggiore Policlinico, Milan, Italy;<br />

3 Department of Clinical Chemistry and Microbiology, Bacteriology<br />

and Virology Units, Fondazione IRCCS Ca’ Granda Ospedale<br />

Maggiore Policlinico, Milan, Italy<br />

Background and aim SOF/R treatment in HCV(+) cirrhotics<br />

on the liver transplant (LT) waiting list is a major breakthrough<br />

since HCV-graft recurrence might be prevented in 95% of<br />

recipients with >4 weeks viral suppression before transplant.<br />

We aimed to evaluate the clinical-virological outcome of<br />

pre-LT SOF/R regimen in patients who stopped treatment at<br />

LT because they were HCV-RNA undetectable for >4 weeks<br />

(SOF-pT) comparing them to patients who received bridging<br />

pre to post-transplant therapy (SOF-bppT) for lack of 4 weeks<br />

of HCV RNA undetectability at LT. Methods From July 2014<br />

an Italian compassionate use program by Gilead-Sciences<br />

endorsed by AISF-AIFA allowed us to treat all HCV consecutive<br />

transplant candidates at our Center with SOF 400 mg/day +<br />

R 200-1000 mg/day. Treated patients were compared with<br />

19 previously transplanted HCV-recipients (January-June 2014)<br />

who were left untreated before LT. Results 30 HCV-LT candidates<br />

received SOF/R pre-LT: 16 received LT while 13 are still<br />

in LT list and 1 died. 9/16 (56%) received SOF-pT whereas 7<br />

(44%) received SOF-bppT for a total duration of 24 weeks. The<br />

2 groups of SOF-treated patients were similar for sex, recipient<br />

age, donor age, HCV genotype or pretreatment viremia,<br />

MELD/Child and type of immunosuppression. The duration of<br />

SOF-therapy before LT was shorter in SOF-b-ppT compared to<br />

SOF-pT (32 vs 85 days, respectively; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1059A<br />

Varuna Aluvihare - Speaking and Teaching: Astellas<br />

John G. O’Grady - Advisory Committees or Review Panels: Astellas, Novartis;<br />

Speaking and Teaching: Astellas<br />

Kosh Agarwal - Advisory Committees or Review Panels: Gilead, BMS, Novartis,<br />

Janssen, AbbVie, Gilead; Consulting: MSD, Janssen; Grant/Research Support:<br />

Roche, Gilead, BMS, BMS; Speaking and Teaching: Astellas<br />

The following authors have nothing to disclose: Anna Mrzljak, Sarah Knighton,<br />

Matthew J. Bruce, Mary Horner, Aisling B. Considine, Michael A. Heneghan,<br />

Abid Suddle<br />

1745<br />

An Evidence-Based Algorithm to Optimize Patient Selection<br />

among Class III Obese Liver Transplant Waitlist<br />

Registrants<br />

Ryan B. Perumpail 1 , Andy Liu 2 , Channa R. Jayasekera 1 , James<br />

H. Marcus 4,5 , Swetha Tummala 1 , Sammy Saab 3 , Zobair M.<br />

Younossi 4,5 , Robert J. Wong 6 , Aijaz Ahmed 1 ; 1 Division of Gastroenterology<br />

and Hepatology, Stanford University Medical Center,<br />

Palo Alto, CA; 2 Medicine, Albert Einstein College of Medicine,<br />

Bronx, NY; 3 Medicine and Surgery, David Geffen School of<br />

Medicine at the University of California at Los Angeles, Los Angeles,<br />

CA; 4 Center for Liver Diseases, Inova Fairfax Hospital, Falls<br />

Church, VA; 5 Betty and Guy Beatty Center for Integrated Research,<br />

Inova Health System, Falls Church, VA; 6 Division of Gastroenterology<br />

and Hepatology, Alameda Health System - Highland Hospital<br />

Campus, Oakland, CA<br />

Background: Class III obesity (MOB), body mass index (BMI)<br />

≥40 kg/m2, is increasing in prevalence among U.S. waitlist<br />

registrants for liver transplantation (LT). MOB is considered a<br />

relative/absolute contraindication for LT at many centers and<br />

by some health insurers due to increased risk of perioperative<br />

morbidity. However, MOB is not independently associated<br />

with lower long-term survival following LT. An evidence-based<br />

algorithm is needed to optimize patient selection for LT in the<br />

setting of MOB. Aim: To establish an evidence-based algorithm<br />

to optimize patient selection for LT in patients with MOB.<br />

Methods: We conducted a retrospective cohort study using<br />

U.S. national data from the United Network for Organ Sharing<br />

registry from 2003-2012 (MELD era) to evaluate the impact of<br />

African American (AA) race, chronic hepatitis C virus (HCV)<br />

infection, diabetes mellitus (DM), hepatocellular carcinoma<br />

(HCC), and the presence of ascites on long-term survival<br />

among adult LT recipients with MOB. Survival following LT was<br />

evaluated with Kaplan Meier methods. Multivariate Cox proportional<br />

hazards models were used to evaluate the independent<br />

impact of the predetermined variables on post-LT survival.<br />

The model was adjusted for age, sex, race/ethnicity, year of<br />

LT, hepatic encephalopathy, Model for End-Stage Liver Disease<br />

score, cardiac disease, HCV status, DM, HCC, and the presence<br />

of ascites. Identified independent predictors of lower survival<br />

were incorporated into a stepwise algorithm for LT patient<br />

selection. Results: Overall, 1,845 adult patients with MOB<br />

underwent LT. When compared to non-HCV patients, 5-year<br />

survival in HCV patients was lower (68.3% vs 75.0%; 95% CI,<br />

63.4%- 72.7% vs. 70.8%-78.7%; p


1060A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Ching-Lung Lai - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead Sciences Inc; Consulting: Bristol-Myers Squibb, Gilead Sciences, Inc;<br />

Speaking and Teaching: Bristol-Myers Squibb, Gilead Sciences, Inc<br />

Man-Fung Yuen - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Bristol-Myers Squibb, Pfizer, GlaxoSmithKline, Bristol-Myers Squibb, Pfizer,<br />

GlaxoSmithKline, Bristol-Myers Squibb, Pfizer, GlaxoSmithKline, Bristol-Myers<br />

Squibb, Pfizer; Grant/Research Support: Roche, Bristol-Myers Squibb,<br />

GlaxoSmithKline, Gilead Science, Roche, Bristol-Myers Squibb, GlaxoSmith-<br />

Kline, Gilead Science, Roche, Bristol-Myers Squibb, GlaxoSmithKline, Gilead<br />

Science, Roche, Bristol-Myers Squibb, GlaxoSmithKline, Gilead Science<br />

The following authors have nothing to disclose: James Fung, Tiffany C. Wong,<br />

Cindy K. Cheung, Kenneth S. Chok, Wing-chiu Dai, Tan To Cheung, William<br />

Sharr, Albert Chan, Sui-Ling Sin, See-Ching Chan, Chung-Mau Lo<br />

Massimo Colombo - Advisory Committees or Review Panels: BRISTOL-MEY-<br />

ERS-SQUIBB, SCHERING-PLOUGH, ROCHE, GILEAD, BRISTOL-MEYERS-SQUIBB,<br />

SCHERING-PLOUGH, ROCHE, GILEAD, Janssen Cilag, Achillion; Grant/<br />

Research Support: BRISTOL-MEYERS-SQUIBB, ROCHE, GILEAD, BRISTOL-MEY-<br />

ERS-SQUIBB, ROCHE, GILEAD; Speaking and Teaching: Glaxo Smith-Kline,<br />

BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH, ROCHE, NOVARTIS, GILEAD,<br />

VERTEX, Glaxo Smith-Kline, BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH,<br />

ROCHE, NOVARTIS, GILEAD, VERTEX, Sanofi<br />

Alessio Aghemo - Advisory Committees or Review Panels: Gilead Sciences, MSD,<br />

AbbVie, Janssen; Speaking and Teaching: Jannsen, MSD, Abbvie, BMS, Gilead<br />

Sciences<br />

The following authors have nothing to disclose: Stella De Nicola, Maria F.<br />

Donato, Caterina Premoli, Giovanna Lunghi, Patrizia Bono, Matteo A. Manini,<br />

Federica Malinverno, Sara Monico, Paolo Reggiani<br />

1747<br />

Prevalence and Clinical Impact of Hepatitis E Virus in<br />

Immunosuppressed Patient after Liver Transplantation<br />

Stella De Nicola 1 , Maria F. Donato 1 , Caterina Premoli 1 , Giovanna<br />

Lunghi 2 , Patrizia Bono 2 , Matteo A. Manini 1 , Pietro Lampertico 1 ,<br />

Federica Malinverno 1 , Sara Monico 1 , Paolo Reggiani 3 , Massimo<br />

Colombo 1 , Alessio Aghemo 1 ; 1 Division of Gastroenterology and<br />

Hepatology, Fondazione IRCCS Ca’ Granda Ospedale Maggiore<br />

Policlinico, University of Milan, Milan, Italy; 2 Department of Clinical<br />

Chemistry and Microbiology, Bacteriology and Virology Units,<br />

Fondazione IRCCS Ca’ Granda, Ospedale Maggiore Policlinico,<br />

Milan, Italy; 3 Division of Surgery and Liver Transplant, Fondazione<br />

IRCCS Ca’ Granda Ospedale Maggiore Policlinico, Milan, Italy<br />

Backgruond: HEV infection is a rising health problem, as it is<br />

the cause of acute and chronic hepatitis especially in immunosuppressed<br />

patients. The prevalence of HEV infection in<br />

Italy and its clinical impact on liver transplanted (LT) patients<br />

is unknown. Aim: To investigate the seroprevalence of IgG-IgM<br />

anti-HEV in LT patients and to compare it with patients with<br />

chronic HCV or HBV infection, or healthy Volunteers. Moreover<br />

we evaluated the impact of HEV in transplanted patients<br />

on overall survival, rejection rates, fibrosis progression and<br />

incidence of extra hepatic manifestastions. Methods: Serology<br />

for HEV IgG and IgM was done with Wantai test (ELISA), HEV<br />

RNA (in-house RT-PCR) was tested in all LT patients in at least<br />

two timepoints during the first year following liver transplantation.<br />

Results: 79 patients who underwent liver transplantation<br />

at our center between 2010 and 2014 were enrolled.<br />

The median age at transplantation was 55 years, 11 (14%)<br />

patients were of foreign origins and HCV was the main etiology<br />

57%. 26 transplanted patients (33%) tested positive for<br />

IgG anti HEV, 3 patients, who were IgG anti-HEV negative<br />

at LT developed IgG during the period of observation. None<br />

of 79 patients were positive for HEV RNA. The seroprevalence<br />

in HCV patients 19% (19/100) (p=0.03), HBV patients<br />

12% (12/100) (p=0.0009) and healthy controls 5% (5/199)<br />

(p=0.0001) were significantly lower than in LT patients. The<br />

prevalence was not influenced by fibrosis stage in HBV or<br />

HCV patients. Overall none of the 79 patients had chronic<br />

HEV infection. In terms of clinical endpoints in the LT patients,<br />

no difference was seen in terms of survival after transplantation,<br />

onset of diabetes, onset of kidney disease, onset of<br />

neurological syndromes or fibrosis progression in IgG positive<br />

versus IgG negative patients. Conclusion: Anti-HEV IgG are<br />

common in the Italian general population and even more so in<br />

LT patients or patients with chronic HCV or HBV infection. However,<br />

in our cohort we do not report a major impact in terms of<br />

clinical endpoints in LT patients probably as a consequence of<br />

the 0% rate of chronic HEV infection.<br />

Disclosures:<br />

Pietro Lampertico - Advisory Committees or Review Panels: Bayer, Bayer; Speaking<br />

and Teaching: Bristol-Myers Squibb, Roche, GlaxoSmithKline, Novartis, Gilead,<br />

Bristol-Myers Squibb, Roche, GlaxoSmithKline, Novartis, Gilead<br />

1748<br />

Long-term outcomes after hepatic resection for hepatocellular<br />

carcinoma in patients with sustained virological<br />

responses to interferon therapy for chronic hepatitis C<br />

Shogo Tanaka 1 , Shigekazu Takemura 1 , Akihiro Tamori 2 , Masahiko<br />

Kinoshita 1 , Norifumi Kawada 2 , Shoji Kubo 1 ; 1 Department of<br />

Hepato-Biliary-Pancreatic Surgery, Osaka City University Graduate<br />

School of Medicine, Osaka, Japan; 2 Department of Hepatology,<br />

Osaka City University Graduate School of Medicine, Osaka,<br />

Japan<br />

Background: The long-term outcome after hepatic resection for<br />

hepatitis C virus (HCV)-related hepatocellular carcinoma (HCC)<br />

in patients who had previously achieved sustained virological<br />

responses (SVR) to interferon (IFN) therapy and who achieved<br />

SVR to adjuvant IFN therapy has been inclusive. Method:<br />

The clinical records of 277 patients who underwent hepatic<br />

resection for HCV-related early stage HCC between 1993<br />

and 2012 were retrospectively reviewed. Thirty-seven of the<br />

277 patients had achieved SVR at the detection of HCC (pre-<br />

SVR group). Twenty-three patients achieved SVR after hepatic<br />

resection (post-SVR group). The control group included the<br />

remaining 217 patients who had received IFN therapy but had<br />

not achieved SVR or who had not received IFN therapy. We<br />

investigated the effect of SVR on the postoperative recurrence<br />

and survival. Results: Twenty-two patients (59%) in the pre-SVR<br />

group, 23 (65%) in the post-SVR group, and 164 (59%) in the<br />

control group had postoperative recurrence during the follow-up<br />

periods, respectively (p=0.09). The disease-free survival rate<br />

was significantly better in the pre-SVR group (5/10/15 years:<br />

49%/29%/29%) and the post-SVR group (61%/36%/27%)<br />

than the control group (23%/7%/7%, p=0.000054). The<br />

overall survival rate at 10/15 years after hepatic resection<br />

was 68%/68% in the pre-SVR group, 78%/78% in the post-<br />

SVR group, and 13%/11% in the control group, respectively<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1061A<br />

pre- and post-operative achievement of SVR had better survival<br />

rates after hepatic resection for HCV-related HCC.<br />

Disclosures:<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

The following authors have nothing to disclose: Shogo Tanaka, Shigekazu<br />

Takemura, Akihiro Tamori, Masahiko Kinoshita, Shoji Kubo<br />

1749<br />

Effects of pre-transplant hepatitis C clearance on liver<br />

transplantation listing and wait time – a single center<br />

experience<br />

Po-Hung Chen 1 , Mary G. Bowring 2 , Lauren M. Kucirka 2 , Christine<br />

M. Durand 1 , Andrew Ofosu 2 , Alia S. Dadabhai 1 , James P. Hamilton<br />

1 , Andrew M. Cameron 3 , Dorry L. Segev 3 , Mark S. Sulkowski 1 ,<br />

Ahmet Gurakar 1 ; 1 Medicine, Johns Hopkins University, Baltimore,<br />

MD; 2 Johns Hopkins University, Baltimore, MD; 3 Surgery, Johns<br />

Hopkins University, Baltimore, MD<br />

Background: The use of hepatitis C virus (HCV)-positive<br />

allografts in viremic recipients helps expand the pool of<br />

potential donors – especially at high usage LT centers – but<br />

these organs are generally inappropriate for patients cleared<br />

of chronic HCV. Despite recent advancements using new<br />

direct-acting antivirals (DAA), no clear guidelines yet exist<br />

regarding the optimal timing of HCV treatment peri-LT. The<br />

present analysis aims to assess the associations of successful<br />

pre-LT HCV treatment in the era of interferon-sparing DAA regimens<br />

with LT listing and time on wait list. Methods: Data were<br />

collected retrospectively for consecutive adults with HCV-related<br />

liver disease who underwent cadaveric LT at the Johns Hopkins<br />

Hospital from January 1, 2014 to April 5, 2015. Recipients<br />

were grouped as “HCV Negative” or “HCV Positive” based on<br />

serum HCV RNA status at LT. The decision to treat HCV pre-LT<br />

was made by individual providers. Between-group comparisons<br />

were made of key recipient and donor characteristics.<br />

Results: There were 50 cadaveric LT for HCV-related liver disease,<br />

32 (64%) of which were in “HCV Positive” group. Within<br />

this group, 13 (41%) HCV-positive allografts were used. No<br />

significant differences were noted between “HCV Negative”<br />

and “HCV Positive” groups in age, gender, race, presence<br />

of hepatocellular carcinoma, and MELD at organ allocation.<br />

Among “Negative” recipients 11 (61%) had blood type O,<br />

versus 12 (38%) in “Positive” group. Sixteen (89%) “Negative”<br />

recipients were alive at 1 year post-LT, compared to 29<br />

(91%) “Positive” patients (P = 0.8). (Table) Finally, concordant<br />

HCV-negative recipient-donor pairs waited a median of<br />

158 days (Q25-Q75: 13-296) before LT; HCV-positive pairs<br />

waited 221 (26-241) days (P = 0.7). Conclusions: There were<br />

no significant between-group differences, including time on<br />

wait list, among recipients regardless of treatment with DAA<br />

pre-LT. The use of HCV-positive organs in viremic recipients did<br />

not adversely impact survival at one year. Larger <strong>studies</strong> are<br />

needed to confirm whether these relationships hold true across<br />

the entire spectrum of MELD and blood types at LT listing.<br />

Disclosures:<br />

Mark S. Sulkowski - Advisory Committees or Review Panels: Merck, AbbVie,<br />

Janssen, Gilead, BMS; Grant/Research Support: Merck, AbbVie, Janssen, Gilead,<br />

BMS<br />

Ahmet Gurakar - Advisory Committees or Review Panels: Gilead; Grant/<br />

Research Support: BMS<br />

The following authors have nothing to disclose: Po-Hung Chen, Mary G. Bowring,<br />

Lauren M. Kucirka, Christine M. Durand, Andrew Ofosu, Alia S. Dadabhai,<br />

James P. Hamilton, Andrew M. Cameron, Dorry L. Segev<br />

1750<br />

Protective Birth Cohort Effect in Baby Boomers Despite<br />

Presence of Poor Predictors<br />

Ryan B. Perumpail 1 , Andy Liu 2 , Channa R. Jayasekera 1 , Zobair M.<br />

Younossi 3,4 , Robert J. Wong 5 , Aijaz Ahmed 1 ; 1 Division of Gastroenterology<br />

and Hepatology, Stanford University Medical Center,<br />

Palo Alto, CA; 2 Medicine, Albert Einstein College of Medicine,<br />

Bronx, NY; 3 Department of Medicine, Center for Liver Diseases,<br />

Inova Fairfax Hospital, Falls Church, VA; 4 Betty and Guy Beatty<br />

Center for Integrated Research, Inova Health System, Falls Church,<br />

VA; 5 Division of Gastroenterology and Hepatology, Alameda<br />

Health System, Highland Hospital, Oakland, CA<br />

Background: Hepatitis C virus (HCV) is a leading etiology of<br />

chronic liver disease in the U.S. Seventy-five percent of HCV<br />

patients are of the baby boomer (BB) generation (born 1945-<br />

1965). Our study aims to analyze the impact of the BB generation<br />

on survival following liver transplantation (LT) in the<br />

U.S. Methods: We conducted a retrospective cohort study<br />

using population-based national data from the United Network<br />

for Organ Sharing 1995-2012 registry to evaluate birth<br />

cohort-specific (BB vs. non-BB) and etiology-specific (HCV vs.<br />

non-HCV) impact on outcomes. Overall post-LT survival was<br />

stratified by birth cohort and HCV status, and was evaluated<br />

with Kaplan Meier methods and multivariate Cox proportional<br />

hazards models. Results: From 1995 to 2012, the proportion<br />

of LT waitlist registrations represented by BB patients increased<br />

from 47.4% to 77.0%. These trends were primarily driven by<br />

HCV, and BB patients accounted for 90% of HCV-related LT<br />

waitlist registrants in 2012. Similar trends were seen among LT<br />

recipients: BB patients accounted for 78.1% of all LT in 2012<br />

compared to 21.9% among non-BB patients. Furthermore, BB<br />

patients accounted for 91.1% of HCV-related LT. Among both<br />

HCV and non-HCV patients, BB patients had significantly better<br />

5-year post-LT survival (HCV: 69.5% vs. 64.2%, p< 0.001;<br />

non-HCV: 79.8% vs. 74.2%, p


1062A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Aijaz Ahmed - Consulting: Bristol-Myers Squibb, Gilead Sciences Inc., Roche,<br />

AbbVie, Salix Pharmaceuticals, Janssen pharmaceuticals, Vertex Pharmaceuticals,<br />

Three Rivers Pharmaceuticals; Grant/Research Support: Gilead Sciences<br />

Inc.<br />

The following authors have nothing to disclose: Ryan B. Perumpail, Andy Liu,<br />

Channa R. Jayasekera, Robert J. Wong<br />

1751<br />

Prevention of post-transplant HCV recurrence by a 24<br />

week regimen of Sofosbuvir/ribavirin (SOF/R) bridging<br />

pre-post transplant (b-ppT): a real life strategy<br />

Maria F. Donato 1 , Sonia Berardi 2 , Rosa Maria Iemmolo 3 , Luca<br />

S. Belli 4 , Marzia Montalbano 5 , Sherrie Bhoori 6 , Giulia Pieri 7 ,<br />

Ilaria Lenci 8 , Alessandra Riva 9 , Veronica Bernabucci 10 , Sara<br />

Monico 1 , Chiara Mazzarelli 4 , Federica Malinverno 1 , Daniele E.<br />

Dondossola 11 , Chiara Taibi 5 , Giorgio Rossi 11 , Antonio Daniele<br />

Pinna 17 , Fabrizio Di Benedetto 3 , Luciano De Carlis 13 , Giuseppe<br />

M. Ettorre 14 , Vincenzo Mazzaferro 6 , Umberto Montin 12 , Giuseppe<br />

Tisone 15 , Marco Vivarelli 16 , Maria Rosa Tamè 2 , Paolo Caraceni 18 ,<br />

Maria Cristina Morelli 2 ; 1 Division of Gastroenterology and Hepatology,<br />

Fondazione IRCCS Cà Granda Ospedale Maggiore Policlinico,<br />

Università degli Studi di Milano, Milan, Italy; 2 Department<br />

of Digestive Disease and Internal Medicine, St Orsola-Malpighi<br />

University Hospital, Bologna, Italy; 3 Liver and Multivisceral Transplant<br />

Center, Azienda Ospedaliero Universitaria di Modena,<br />

Modena, Italy; 4 Hepatology and Gastroenterology Unit, Ospedale<br />

Niguarda Cà Granda, Milano, Italy; 5 Hepatology and Infectious<br />

Diseases Unit, National Institute for Infectious Diseases “L. Spallanzani”,<br />

Roma, Italy; 6 Gastroenterology, Surgery and Liver Transplantation,<br />

Fondazione Istituto Nazionale Tumori IRCCS, Milan,<br />

Italy; 7 Division of Hepatology, IRCCS AO San Martino-IST-Genova,<br />

Genova, Italy; 8 Hepatology Unit, Università Tor Vergata Roma,<br />

Roma, Italy; 9 Division of Infectious Diseases, AOU Ospedali Riuniti<br />

Ancona, Ancona, Italy; 10 Division of Gastroenterology, Azienda<br />

Ospedaliero Universitaria di Modena, Modena, Italy; 11 Division of<br />

General Surgery and Liver Transplantation, Fondazione IRCCS Ca’<br />

Granda Ospedale Maggiore Policlinico, Milan, Italy; 12 Epatobiliary<br />

and Liver Transplantation Surgery, AO Universitaria Integrata<br />

Verona, Verona, Italy; 13 Division of General Surgery and Liver<br />

Transplantation, Ospedale Niguarda Ca’ Granda, Milan, Italy;<br />

14 Division of General Surgery and Liver Transplantation, National<br />

Institute for Infectious Diseases “L. Spallanzani”, Roma, Italy;<br />

15 Department of Experimental Medicine and Surgery, Università<br />

Tor vergata Roma, Roma, Italy; 16 Hepatobiliary and Abdominal<br />

Transplantation Surgery, AOU Ospedali Riuniti Ancona, Ancona,<br />

Italy; 17 Liver Surgery and Transplantation Unit, St Orsola-Malpighi<br />

University Hospital, Bologna, Italy; 18 Division of Internal Medicine,<br />

St Orsola-Malpighi University Hospital, Bologna, Italy<br />

Background and aims: SOF/R treatment in HCV(+) cirrhotics<br />

waiting liver transplant (LT) is a major breakthrough since HCVgraft<br />

recurrence might be prevented in 95% of recipients with<br />

>4 weeks viral suppression before transplant. We aimed to<br />

evaluate if bridging SOF/R from pre to post-transplant phase<br />

(b-ppT) could be feasible, safe and prevent HCV-graft recurrence<br />

also in those HCV patients without HCV-RNA negativity<br />

for at least 4 weeks before LT. Methods: SOF 400 mg/day+R<br />

200-1200 mg/day was given for a total duration of 24 weeks<br />

pre-post-LT. Results: 25 HCV-LT candidates (2re-LT) received<br />

b-ppT SOF within an Italian compassionate use program (ITA-<br />

COPS study). LT indication was HCC in 17 (MELD 8-21) and<br />

decompensated cirrhotics in 8 (MELD 9-26). HCV-genotype<br />

was 1 or 4 in 18. At transplant: 14 patients were HCV-RNA(+)<br />

with a median viral load of 35 IU/ml (12-2584) and 11 were<br />

HCV-RNA undetected for less than 4 weeks following a median<br />

treatment of 36 days of SOF/R (5-112). Post-transplant: 16<br />

patients transiently discontinued SOF after-LT (median=1 day),<br />

but all cleared HCV within week-4 post-LT. At the time of the<br />

analysis, all recipients were HCV-free and alive, with a median<br />

post-transplant follow-up of 26 weeks (7-98). Eighteen recipients<br />

(72%) completed SOF/R treatment with SVR4 and SVR12<br />

rates of 100% (17/17 and 9/9, respectively). One patient was<br />

re-transplanted; 7 developed an episode of AKI; 3=arytmia;<br />

1=sepsis; 4 (Cya-group) cellular rejection requiring switch to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1063A<br />

Tacrolimus/steroid boli. Conclusions: A b-ppT SOF/R-regimen<br />

in HCV-listed cirrhotics with suboptimal virological response at<br />

the time of transplant, might represent an optimal strategy to<br />

avoid post-transplant graft reinfection.<br />

Disclosures:<br />

Sherrie Bhoori - Speaking and Teaching: Bayer, BTG<br />

Vincenzo Mazzaferro - Advisory Committees or Review Panels: Bayer; Grant/<br />

Research Support: BTG<br />

Paolo Caraceni - Advisory Committees or Review Panels: GSK; Consulting: BMS;<br />

Grant/Research Support: Grifols; Speaking and Teaching: Baxter, Kedrion<br />

The following authors have nothing to disclose: Maria F. Donato, Sonia Berardi,<br />

Rosa Maria Iemmolo, Luca S. Belli, Marzia Montalbano, Giulia Pieri, Ilaria Lenci,<br />

Alessandra Riva, Veronica Bernabucci, Sara Monico, Chiara Mazzarelli, Federica<br />

Malinverno, Daniele E. Dondossola, Chiara Taibi, Giorgio Rossi, Antonio<br />

Daniele Pinna, Fabrizio Di Benedetto, Luciano De Carlis, Giuseppe M. Ettorre,<br />

Umberto Montin, Giuseppe Tisone, Marco Vivarelli, Maria Rosa Tamè, Maria<br />

Cristina Morelli<br />

1752<br />

Hepatitis E infection in immunocompromised persons in<br />

Argentina<br />

Jose Debes 1 , Maribel Martinez Wassaf 2 , Maia Belen Pisano 3 , Leonardo<br />

Marianelli 4 , Martin Lotto 3 , Domingo Balderramo 5 , Hernan<br />

Coseano 5 , Viviana Re 3 ; 1 University of MInnesota, Minneapolis,<br />

MN; 2 Área de Virología y Biología Molecular, LACE Laboratorios,<br />

Cordoba, Argentina; 3 Instituto de Virología “Dr.J.M.Vanella”, Facultad<br />

de Ciencias Médicas, Universidad Nacional de Córdoba,<br />

Cordoba, Argentina; 4 Hospital Rawson, Cordoba, Argentina;<br />

5 Hospital Privado, Cordoba, Argentina<br />

Background: Hepatitis E virus (HEV) is a single-stranded RNA<br />

virus that can cause hepatitis in an epidemic fashion. In immunocompetent<br />

individuals, infection with HEV usually leads to<br />

silent seroconversion or to acute self-limited disease. Several<br />

reports suggest an increased rate of seroprevalence of HEV in<br />

immunosuppressed individuals, particularly those undergoing<br />

solid-organ transplantation. In this setting, HEV can develop<br />

into a chronic infection. The data is less clear in patients<br />

infected with human immunodeficiency virus (HIV). Moreover,<br />

information about prevalence of HEV in immunocopromised<br />

subjects outside of Europe or North America is scarce. In this<br />

study we addressed the seroprevalence of HEV in immunocompromised<br />

subjects in Argentina and associated risk factors.<br />

Methods: We performed third generation enzyme immunoassay<br />

for determination of IgG specific antibodies against HEV<br />

in 95 subjects infected with HIV, 81 subjects on hemodialysis<br />

(HD) and 58 solid-organ transplant recipients. On those samples<br />

that were positive for HEV IgG, further assessment of HEV<br />

IgM levels. HEV PCR was performed in all samples. Subjects on<br />

HD and transplant recipients were evaluated regarding social<br />

habits and potential risk factors. Results were compared to<br />

433 HIV-negative, immunocompetent controls from our center.<br />

Results: In our entire HIV-positive group we found 8 of<br />

95 samples to be positive for HEV IgG (8.8%), compared to<br />

19 of 433 samples (4.4%) in the control group. Interestingly,<br />

in a subgroup of patients (N27) with low CD4 counts 18%<br />

were positive for HEV IgG (p=0.03, compared to controls).<br />

This group had a much lower CD4 count when compared to<br />

the general HIV cohort (median CD4 count of 43/mm3 vs<br />

543/mm3 respectively). Eight out of 81 subjects (9.8%) on HD<br />

and 5 of 58 (8.6%) of transplant recipients were positive for<br />

HEV IgG. Interestingly, half of HEV seropositive patients in the<br />

HD group had positive IgM for HEV. There was no association<br />

between HEV serostatus and consumption of pork, fish, potable<br />

water or history of blood transfusion. Only 1 sample showed<br />

a positive PCR for HEV, within the HIV group. This patient<br />

had recent history of vomiting and diarrhea and likely experienced<br />

acute HEV infection. Conclusions: Our study found an<br />

increased seroprevalence of HEV IgG in subjects infected with<br />

HIV, on HD and solid-organ transplant recipients in Argentina.<br />

However, the only significant difference compared to controls<br />

was on HIV-infected patients with low CD4 counts.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jose Debes, Maribel Martinez<br />

Wassaf, Maia Belen Pisano, Leonardo Marianelli, Martin Lotto, Domingo Balderramo,<br />

Hernan Coseano, Viviana Re<br />

1753<br />

Oral Interferon-free Antiviral Regimens in Liver Transplant<br />

Recipients with Hepatitis C: a Real Life Experience<br />

Arun Mannem 2,1 , Shari S. Rogal 2 , Vivek Sharma 2 , Fadi Francis 2 ,<br />

Thomas V. Cacciarelli 2 , Obaid S. Shaikh 2,1 ; 1 Medicine/Gastroenterology,<br />

University of Pittsburgh School of Medicine, Pittsburgh,<br />

PA; 2 Medicine/Gastroenterology, Veterans Affairs Pittsburgh<br />

Healthcare System, Pittsburgh, PA<br />

Purpose: Oral interferon-free regimens are highly effective for<br />

hepatitis C (HCV) in the immunocompetent population. However,<br />

their role in transplant recipients is not well established.<br />

This study aimed to determine the real life experience with<br />

such regimens in liver transplant recipients. Methods: Beginning<br />

in 2013, all patients evaluated for HCV at VA Pittsburgh<br />

Healthcare System were prospectively entered in a quality<br />

of care database. One hundred twenty five liver transplant<br />

recipients with HCV were identified; 25 were excluded [treatment<br />

with interferon based regimens (13), transplant at a<br />

different center (6), lack of complete data (4), primary graft<br />

failure (2), and treatment in a clinical trial (1)]. Data regarding<br />

demographics, prior post-transplant treatment, HCC and<br />

MELD status and treatment were collected. Results: Among 100<br />

patients studied, 96 received transplant during 2001-2014.<br />

Most patients were adult, white males and received tacrolimus<br />

based immunosuppression. Eighty-six (86%) had genotype 1,<br />

1 (1%) genotype 2, 12 (12%) genotype 3 and 1 (1%) genotype<br />

4. Among genotype 1 patients, 60/86 (70%) underwent<br />

treatment post-transplant; 26/60 (43%) received combination<br />

sofosbuvir/simeprevir (SOF/SIM), 5 (8%) ledipasvir/sofosbuvir<br />

(LED/SOF) x 12 weeks, 6 (10%) LED/SOF x 24 weeks,<br />

21 (35%) LED/SOF/ribavirin (RBV) x 12 weeks and 1 (2%)<br />

received LED/SOF/RBV x 24 weeks. Rapid virologic response<br />

rates at 4 weeks of treatment (RVR) follow: - SOF/SIM 16/26<br />

(62%), LED/SOF-12 weeks 2/5 (40%), LED/SOF-24 weeks<br />

3/6 (50%), LED/SOF/RBV-12 weeks 14/21 (67%) and LED/<br />

SOF/RBV-24 weeks 0/1 (0%). The single genotype 2 patient<br />

was treated with SOF/RBV and response is awaited. Four genotype<br />

3 patients were treated- 1 (25%) received SOF/SIM,<br />

2 (50%) LED/SOF x 12 weeks and, 1 (25%) SOF/RBV. The<br />

single genotype 4 patient had treatment with LED/SOF/RBV x<br />

12 weeks and achieved RVR. The full treatment outcomes are<br />

awaited and will be reported at the meeting. Conclusion: Our<br />

study will shed light on real world treatment outcomes with the<br />

all oral interferon-free regimens in liver transplant recipients<br />

with HCV. It will help establish the tolerability, efficacy and<br />

duration of therapy in this special population.<br />

Disclosures:<br />

Shari S. Rogal - Grant/Research Support: Gilead Sciences


1064A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Obaid S. Shaikh - Grant/Research Support: Gilead Sciences, Shinongi Pharmaceuticals;<br />

Speaking and Teaching: Simply Speaking<br />

The following authors have nothing to disclose: Arun Mannem, Vivek Sharma,<br />

Fadi Francis, Thomas V. Cacciarelli<br />

1754<br />

Effect of simeprevir administration on dose-normalized<br />

concentration of calcineurin inhibitors in solid organ<br />

transplant recipients with hepatitis C infection<br />

Heather Johnson 1,2 , Maxwell A. Brown 2 , Alicia B. Lichvar 2 ,<br />

Michael A. Dunn 1 , Kapil B. Chopra 1 ; 1 University of Pittsburgh,<br />

PIttsburgh, PA; 2 UPMC Presbyterian, Pittsburgh, PA<br />

Background Simeprevir (SMV) is a second generation NS3/4A<br />

protease inhibitor that lacks the immunomodulatory and other<br />

adverse effects of interferon, representing an attractive therapeutic<br />

alternative for the treatment of hepatitis C virus infection<br />

in solid organ transplant recipients. The effect of P-glycoprotein<br />

(P-gp) inhibition by SMV on serum concentrations of the calcineurin<br />

inhibitors (CNI) cyclosporine (CsA) and tacrolimus (TAC)<br />

is not well described in the literature. The purpose of this analysis<br />

was to determine the magnitude of effect of P-gp inhibition<br />

with SMV on CNI blood levels. Methods Liver transplant recipients<br />

(LTRs) on maintenance immunosuppression with a CNI<br />

who initiated antiviral treatment with SMV between 1/2014<br />

and 4/ 2015 were included in this analysis. SMV was administered<br />

along with the NS5B polymerase inhibitor sofosbuvir for<br />

recurrent Genotype 1 allograft HCV. Dose-normalized levels<br />

were calculated by dividing CNI trough levels by the CNI total<br />

daily dose in order to account for CNI dose changes related<br />

to changes in CNI trough levels over time. Dose-normalized<br />

CNI levels in patients receiving 12 weeks of SMV were compared<br />

using Friedman’s ANOVA. Median percent change in<br />

dose-normalized levels were compared between LTRs with low<br />

(0–2) and high (3–6) Metavir scores using the Mann-Whitney<br />

U test. Results Nineteen patients were identified for inclusion in<br />

this analysis, of whom 15 (78.9%) received 12 weeks of SMV<br />

and 4 (21.1%) received 24 weeks of SMV. The majority of<br />

patients were Caucasian (100.0%), male (63.2%), liver transplant<br />

recipients (100.0%) on maintenance immunosuppression<br />

with TAC (n=17) with an average age of 59.2 years (SD±6.7).<br />

There was no significant difference between dose-normalized<br />

levels of TAC at baseline and during HCV treatment. Three<br />

patients (15.8%) had an increase in CNI dose and 4 patients<br />

(21.1%) had a decrease in CNI dose following initiation of<br />

SMV. There was no significant difference in median percent<br />

change in dose-normalized TAC levels between groups with<br />

low and high Metavir scores. Conclusions Patients on maintenance<br />

immunosuppression with TAC who were treated with 12<br />

weeks of SMV did not experience significant changes in their<br />

dose-normalized TAC trough levels. More frequent monitoring<br />

of TAC trough levels does not appear to be warranted in transplant<br />

recipients initiating SMV therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Heather Johnson, Maxwell A.<br />

Brown, Alicia B. Lichvar, Michael A. Dunn, Kapil B. Chopra<br />

1755<br />

Bioluminescence imaging of mesenchymal stem cells by<br />

over expression hepatocyte nuclear factor4α: tracking<br />

survival and biodistribution<br />

Pei-yi Xie, Xiaojun Hu, Xiaochun Meng, Hong Shan; Department of<br />

Radiology, The Third Affiliated Hospital of Sun Yat-sen University,<br />

Guangzhou, China<br />

Hepatocytes from human bone marrow-derived mesenchymal<br />

stem cells (hBM-MSCs) are expected to be a useful source<br />

for cell transplantation. However, relatively low efficiency of<br />

hepatic differentiation of hBM-MSCs remains unsolvable in clinical<br />

application. Hepatocyte nuclear factor 4alpha (HNF4α),<br />

a critical transcription factor, is essential for the entire process<br />

of liver development. The purpose of this study was to construct<br />

MSCs with over expression HNF4α and investigate their<br />

hepatic differentiation and therapeutic potential for treating<br />

Fulminant hepatic failure (FHF) rats, and track their survival<br />

and biodistribution after transplantation by bioluminescence<br />

imaging (BLI). HNF4α gene was transduced into hBM-MSCs by<br />

lentiviral vector (pLV/Final-puro-hHNF4α-hrGFP). HNF4α-transduced<br />

MSCs (E7-hHNF4α) and GFP-transduced MSCs cells<br />

(E7-hrGFP ) were labeled with pLENTi-CMV-luc2-mKate2 and<br />

analyzed for their hepatic functions such as measurements<br />

of albumin, urea, glucose, cytochrome P450 activity,Indocyanine<br />

green (ICG) uptake and release, and drug metabolization<br />

in vitro. FHF modals of Sprague-Dawley (SD) rats were<br />

established and divided into three groups: PBS, E7-hrGFP and<br />

E7-hHNF4α cells’ transplantation. After 2.0×10 6 cells transplantation,<br />

BLI was used to dynamically track the survival and<br />

biodistribution of implanted E7-hrGFP cells and E7-hHNF4α<br />

cells. The restoration of biological functions of the livers receiving<br />

transplantation was assessed via a variety of approaches<br />

such as mortality rate determination, serum biochemical analysis,<br />

and histological, immunohistochemical, and genetic analysis.<br />

In vitro, E7-hHNF4α cells showed mature hepatic functions<br />

including albumin secretion, urea production, glycogen storage,<br />

cytochrome P450 activity, ICG uptake and release and<br />

drug metabolization. They improved liver functions in vivo after<br />

transplantation into the D-galactosamine-injured rats’ liver as<br />

evidenced by the fact that AST, ALT, TBIL returned to normal<br />

levels in recipient FHF rats. The result of 30-day survival rates<br />

suggested that E7-hHNF4α cells’ transplantation via superior<br />

mesenteric vein (SMV) was able to significantly prolong the<br />

survival of FHF rats compared with the other two groups. BLI,<br />

histochemisty, and RT-PCR results confirmed the presence of<br />

transplanted E7-hrGFP cells and E7-hHNF4α cells in recipient<br />

rat livers. CONCLUSION: Our data revealed that E7-hHNF4α<br />

cells could not only differentiate into functional hepatocyte-like<br />

cells in vitro, but could also improve liver functions and prolong<br />

the survival time of FHF rats. And BLI is a useful tool to track the<br />

transplanted cells survival and biodistribution.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pei-yi Xie, Xiaojun Hu, Xiaochun<br />

Meng, Hong Shan


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1065A<br />

1756<br />

Dual CCR2/CCR5 Antagonist Cenicriviroc Leads to<br />

Potent and Significant Reduction in Proinflammatory<br />

CCR2+ Monocyte Infiltration in Experimental Acute Liver<br />

Injury<br />

Oliver Krenkel 1 , Tobias Püngel 1 , Jana C. Mossanen 1 , Can Ergen 1 ,<br />

Felix Heymann 1 , Eric Lefebvre 2 , Christian Trautwein 1 , Frank<br />

Tacke 1 ; 1 Medical Clinic III, University Hospital Aachen, Aachen,<br />

Germany; 2 Tobira Therapeutics Inc., San Francisco, CA<br />

Aim of the study: Acute liver failure (ALF) is a life-threatening<br />

condition with rapid deterioration of hepatic function and<br />

limited therapeutic options. In mouse models, liver injury by<br />

toxic agents like carbon tetrachloride (CCl4) or acetaminophen<br />

(APAP) leads to rapid pro-inflammatory monocyte infiltration<br />

into the liver via the CCR2–CCL2 (a.k.a. MCP-1) chemokine<br />

pathway. Cenicriviroc (CVC) is an oral, once-daily CCR2/<br />

CCR5 antagonist currently evaluated in a Phase-2b clinical trial<br />

in adults with NASH and liver fibrosis. We therefore evaluated<br />

CVC for inhibiting monocyte infiltration in CCl4 and APAP-induced<br />

acute liver injury in vivo. Methods: C57BL/6J (WT) and<br />

CCR2-deficient mice were subjected to ALF either by CCl4<br />

(0.6 ml/kg IP) or APAP (250 mg/kg IV). In both models, mice<br />

received either CVC (100 mg/kg) or vehicle by oral gavage.<br />

Liver injury and immune cell phenotypes were analyzed. For<br />

mechanistic <strong>studies</strong>, monocyte-derived macrophage subsets<br />

and resident macrophages (Kupffer cells) were sorted by FACS<br />

from injured livers and subjected to array-based Nanostring<br />

gene expression analysis. Results: Both CCl4 and APAP led<br />

to a rapid and massive accumulation of Ly6C+, monocyte-derived<br />

macrophages in injured livers, dependent on the chemokine<br />

receptor CCR2. Oral administration of CVC significantly<br />

reduced pro-inflammatory Ly6C+ monocytes in blood (p


1066A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1758<br />

Arcuate artery resistive index predicts Acute-on-Chronic<br />

Liver Failure<br />

Pablo Solis-Munoz 2,1 , Christopher Willars 2 , Julia Wendon 2 ,<br />

Georg Auzinger 2 , Michael A. Heneghan 2 , María De la Flor-Robledo<br />

3 , José A. Solís-Herruzo 1 ; 1 Research Center, Hospital “12 de<br />

Octubre”, Madrid, Spain; 2 Institute of Liver Studies, King’s College<br />

Hospital, London, United Kingdom; 3 University Hospital Severo<br />

Ochoa,, Madrid, Spain<br />

Patients with acutely decompensated (AD) liver cirrhosis are<br />

at risk for developing acute-on-chronic liver failure (ACLF) syndrome.<br />

This syndrome is associated with a high sort-term mortality<br />

rate. The aim of our study was to identify reliable and<br />

early predictors of developing ACLF in cirrhotic patients with<br />

AD. Methods: We assessed 84 cirrhotic patients admitted for<br />

AD without ACLF on admission. We performed routine blood<br />

testing and detailed ultrasound Doppler <strong>studies</strong> of systemic<br />

arteries and mayor abdominal veins and arteries. We also<br />

calculated liver and ICU predictive scores. The area under<br />

the ROC curve (AUROC) was calculated for all variables that<br />

were significantly different between patients who developed<br />

ACLF and those who did not. Sensitivity, specificity, positive<br />

and negative predictive values, and diagnostic accuracy for<br />

the prediction of the short-term ACLF development were determined.<br />

Results: Of the 84 patients, 23 developed ACLF and 61<br />

did not develop this complication. In the univariate analysis,<br />

serum levels of urea, creatinine, and bilirubin, prothrombin<br />

time, MELD score, portal vein, femoral and poplitea artery flow<br />

velocity, and renal and arcuate artery resistive indices (RI) had<br />

predictive value for short-term development of ACLF. However,<br />

only arcuate artery RI had independent predictive value in the<br />

multivariate analysis. The AUROC value for RI of the arcuate<br />

arteries was 0.9971. Conclusions: On the first day of admission,<br />

ultrasound measurement of the RI of the arcuate arteries<br />

recognizes with a high degree of certainty cirrhotic patients<br />

admitted because of AD who will develop ACLF.<br />

Disclosures:<br />

Julia Wendon - Consulting: Pulsion, Excalenz<br />

The following authors have nothing to disclose: Pablo Solis-Munoz, Christopher<br />

Willars, Georg Auzinger, Michael A. Heneghan, María De la Flor-Robledo, José<br />

A. Solís-Herruzo<br />

1759<br />

Fulminant Hepatitis B Acute Liver Failure due to immunosuppressant<br />

or chemo-therapy: A multicenter retrospective<br />

cohort analysis<br />

Filipe S. Cardoso 1 , Michelle Gottfried 3 , K. Rajender Reddy 2 , A.<br />

James Hanje 4 , Daniel Ganger 5 , William M. Lee 6 , Constantine J.<br />

Karvellas 1 ; 1 Hepatology/Critical Care Medicine, University of<br />

Alberta, Edmonton, AB, Canada; 2 University of Pennsylvania, Philadelphia,<br />

PA; 3 Medical University of South Carolina, Charlston,<br />

SC; 4 The Ohio State University, Columbus, OH; 5 Northwestern<br />

University, Chicago, IL; 6 University of Texas Southwestern, Dallas,<br />

TX<br />

Background/aims: Acute Liver Failure (ALF) due to reactivation<br />

of latent hepatitis B (HBVr) due to chemotherapy or immunosuppression<br />

(CTI) can be fatal and yet is preventable in many<br />

cases. Aims: To determine the frequency of HBV-ALF due to<br />

reactivation from CTI within the US Acute Liver Failure Study<br />

Group (ALFSG) registry and compare patient survival and complication<br />

rates to non-CTI HBV-ALF controls. Methods: Of a total<br />

of 149 patients with HBV induced ALF enrolled in the US ALFSG<br />

registry between 01/1998 and 04/2015, 26 (17%) developed<br />

HBV-ALF due to CTI and 123 (83%) did not (controls).<br />

Multivariable logistic regression was performed to determine<br />

predictors of 21-day spontaneous survival in HBV ALF, specifically<br />

examining the impact of HBVr ALF due to CTI. Results:<br />

Of the 26 CTI patients, 9 received corticosteroids, 15 cytotoxic<br />

chemotherapy (breast, lymphoma, testicular, leukemia,<br />

myeloma), 2 received rituximab and 1 received gemtuzumab.<br />

There were no differences between these groups in proportions<br />

of patients with coma grade 3 or 4 (67% vs. 63%, p=0.77),<br />

requirement for mechanical ventilation (32% vs. 46%, p=0.19)<br />

or renal replacement therapy (18% vs. 15%, p=0.70). Of CTI<br />

HBV-ALF patients, 10 (38%) were listed for liver transplant (LT)<br />

(vs. 46% controls, p=0.70) and 6 (26%) received LT (vs. 34%<br />

controls, p=0.37). Of these 6 LT patients, 2 died post-LT (survival<br />

post-LT 67% vs. 81% in controls, p=0.59). Overall unadjusted<br />

21-day survival was significantly lower (39% vs. 61%,<br />

p=0.015) in CTI HBVr- ALF patients but spontaneous survival<br />

was similar (25% vs. 35%, p=0.24) when compared with controls.<br />

Using multivariable logistic regression to examine clinical<br />

factors associated with 21-day spontaneous survival in HBV<br />

ALF patients, MELD (Odds Ratio 0.87 per increment; 95% CI<br />

0.80-0.95, p=0.0013) and maximum hepatic coma grade<br />

> 2 (OR 0.13, 95% CI 0.04-0.41, p=0.0005) were significant.<br />

There was a non-significant trend towards worse 21-day<br />

adjusted spontaneous survival in CTI HBV ALF patients (OR<br />

0.16, 0.02-1.54, p=0.11). The model performed well (c-statistic<br />

0.87). Summary/Conclusions: HBV ALF due to CTI has<br />

poor spontaneous survival (25%) and overall poorer outcomes,<br />

even after adjusting for MELD and Hepatic coma grade (>2).<br />

Guidelines from specialties that prescribe chemotherapy/immunosuppressant<br />

therapies need to more explicitly detail testing in<br />

who, how testing for those at risk should be performed, timing<br />

and duration of therapy, and surveillance for those at risk not<br />

on therapy. In this population of HBV-ALF patients where liver<br />

transplant may not be offered, prevention of reactivation is<br />

critical given the significant mortality observed.<br />

Disclosures:<br />

K. Rajender Reddy - Advisory Committees or Review Panels: Merck, Janssen,<br />

Vertex, Gilead, BMS, Abbvie; Grant/Research Support: Merck, BMS, Ikaria,<br />

Gilead, Janssen, AbbVie<br />

A. James Hanje - Consulting: Salix pharmaceutical<br />

Daniel Ganger - Advisory Committees or Review Panels: Bristol Myers, Abbvie;<br />

Grant/Research Support: Merck, Ocera; Speaking and Teaching: Gilead<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

Constantine J. Karvellas - Grant/Research Support: Merck; Speaking and Teaching:<br />

Gambro<br />

The following authors have nothing to disclose: Filipe S. Cardoso, Michelle<br />

Gottfried<br />

1760<br />

Mesenchymal stem cell using scaffold treatment<br />

increased survival on acute liver failure model<br />

Yong Kyun Cho 1 , Dae Won Jun 2 , Eun Chul Jang 3 , Seung Min<br />

Lee 3 , Joo Hee Kwak 2 ; 1 Department of Internal Medicine, Kangbuk<br />

Samsung Hospital, Sungkyunkwan University, School of Medicine,<br />

Seoul, Korea (the Republic of); 2 Department of Internal Medicine,<br />

Hanyang University College of Medicine, Seoul, Korea (the Republic<br />

of); 3 Department of Occupational and Environmental Medicine,<br />

Soonchunhyang University Cheonan Hospital, Cheonan, Korea<br />

(the Republic of)<br />

Background: Recently, the effects of mesenchymal stem cells<br />

(MSCs) have been extensively evaluated in acute liver failure<br />

model. However, the most commonly used methods are used<br />

injecting MSCs through a vessel, which has a major drawback<br />

of homing cells in other organs. The aim of current study was<br />

to evaluate the efficacy and degree of cell homing of scaffold<br />

loaded with MSCs on acute liver failure model. Methods: Acute


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1067A<br />

liver failure was induced in C57BL/6J mice using thioacetamide<br />

(TAA) (200mg/kg, i.p) once a day for 2 days. In phase<br />

I, animals were divided in three groups: 1) TAA+saline via tail<br />

vein; 2) TAA+MSCs via tail vein; 3) TAA+Scaffold loaded with<br />

MSCs to evaluate the mortality and changes in liver function.<br />

PLGA Poly (lactic-co-glycolic acid) scaffold alone, and loaded<br />

with 1x10 6 human MSCs were implanted on dorsum of animals.<br />

In phase II, animals were divided into three groups: 1)<br />

TAA; 2) TAA+Scaffold; 3) TAA+Scaffold loaded with MSCs<br />

to evaluate liver function and mortality. Animals were sacrificed<br />

3 days after implantation, serum and liver samples were<br />

collected. Liver tissue immunohistochemistry for proliferation<br />

marker (Ki67) was performed. Results: In phase I, the mortality<br />

rate of TAA+Scaffold group was 36% as compared to<br />

TAA+PBS tail vein (66%) and TAA+Scaffold only groups (54%).<br />

In phase II, TAA+Scaffold loaded with MSCs group had only<br />

11.8% mortality as compared to TAA (30.4%) and TAA+Scafold<br />

only (52.6%) (p=0.014) groups. In phase II, serum AST<br />

levels of TAA+Scaffold loaded with MSCs was lower than other<br />

two groups, while serum ALT levels of TAA+Scaffold group<br />

were the lowest. Furthermore, TAA+Scaffold loaded with MSCs<br />

group showed increased liver tissue staining for Ki67. Later, we<br />

performed PCR using human specific primer, to evaluate homing<br />

of MSCs in mouse liver. There was no detection of human<br />

cells residing in mouse liver. Conclusion: Scaffold loaded MSCs<br />

showed protective effects via paracrine signaling on acute liver<br />

failure model.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yong Kyun Cho, Dae Won Jun,<br />

Eun Chul Jang, Seung Min Lee, Joo Hee Kwak<br />

1761<br />

Indeterminate Etiology of Acute Liver Failure: Fact or<br />

Fiction?<br />

Jody A. Rule 1 , Jorge Rakela 2 , Norman L. Sussman 3 , Nathan M.<br />

Bass 4 , Holly Tillman 5 , Jaime L. Speiser 5 , R. Todd Stravitz 6 , Ryan<br />

Taylor 7 , William M. Lee 1 , Daniel Ganger 8 ; 1 Internal Medicine,<br />

University of Texas Southwestern Medical Center, Dallas, TX; 2 Gastroenterology<br />

and Hepatology, Mayo Clinic Hospital, Phoenix,<br />

AZ; 3 Division of Abdominal Transplantation, Baylor College of<br />

Medicine, Houston, TX; 4 Department of Medicine, University of<br />

California San Francisco, San Francisco, CA; 5 Data Coordination<br />

Unit, Department of Public Health Sciences, Medical University<br />

of South Carolina, Charleston, SC; 6 Internal Medicine, Virginia<br />

Commonwealth University, Richmond, VA; 7 Gastroenterology and<br />

Hepatology, University of Kansas Medical Center, Kansas City,<br />

KS; 8 Gastroenterology and Hepatology, Northwestern Memorial<br />

Hospital, Chicago, IL<br />

Background: Approximately 11% of acute liver failure or<br />

acute liver injury (ALF/ALI) cases are initially listed as “indeterminate”<br />

by attending hepatologists at academic transplant<br />

centers; this has been attributed in the past to an unidentified<br />

virus or toxic injury. Although nearly 20% of these cases<br />

demonstrate a positive acetaminophen (APAP)-adduct assay<br />

(Hepatology 2011;53:567), the remainder still are considered<br />

indeterminate. Methods: We examined 302 indeterminate<br />

cases from 2713 consecutively enrolled patients in the<br />

ALFSG Registry using pre-defined characteristics developed by<br />

an expert committee, to identify each etiology as either Highly<br />

Likely/Definite (HL/D: >75% certainty) or Probable (>50% certainty).<br />

Examples: Criteria for autoimmune ALF (AI-ALF): calculated<br />

globulins elevated >ULN or serum IgG >ULN; ANA,<br />

ASMA or LKMA-positive titer >1:40; biopsy features: plasma<br />

cells, centrilobular necrosis or interface hepatitis. Thus, Highly<br />

Likely/Definite (HL/D) AI-ALF: All 3 criteria met, Probable AIH:<br />

Any 2 of 3 met. For APAP, HL/D: Positive APAP-adduct assay<br />

or history of APAP ingestion or APAP level (or both) and AST<br />

or ALT >1000/Bili 1 etiology. Among the remaining<br />

86 cases, 51 had at least 1 primary diagnostic test missing<br />

(APAP-adducts, ANA, aHAVIgM, HBsAg, aHBcIgM, or aHCV).<br />

Only 35 cases had a complete set of medical history and laboratory<br />

tests, and no etiology defined, fulfilling our diagnosis of<br />

“indeterminate” ALF. Among patients initially diagnosed with<br />

indeterminate ALF or ALI, 216 (72%) could be assigned to a<br />

single defined etiology with additional testingand standardized<br />

criteria. Dual etiology (or uncertainty with two possibilities)<br />

will need further adjudication. Incomplete history or diagnostic<br />

testing precluded assignment of etiology in 17%. Conclusions:<br />

Only 11.5% of “indeterminates”, 1.2% of the overall ALF registry,<br />

fulfilled stringent criteria for “indeterminate etiology” (data<br />

complete, no cause discernible). With a defined/adjudicated<br />

approach to causality, true “indeterminate” ALF is much less<br />

common than previously estimated.<br />

Disclosures:<br />

Norman L. Sussman - Advisory Committees or Review Panels: BMS, Merck;<br />

Board Membership: HepaHope; Grant/Research Support: AbbVie, BMS, Biotest,<br />

Conatus, Esai, Genfit, Gilead, Hyperion, Immuron, Intercept, Lumena, Merck,<br />

Novartis, Ocera; Speaking and Teaching: AbbVie, Gilead, Jannsen<br />

Nathan M. Bass - Advisory Committees or Review Panels: Quest Diagnostics<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

Daniel Ganger - Advisory Committees or Review Panels: Bristol Myers, Abbvie;<br />

Grant/Research Support: Merck, Ocera; Speaking and Teaching: Gilead<br />

The following authors have nothing to disclose: Jody A. Rule, Jorge Rakela, Holly<br />

Tillman, Jaime L. Speiser, R. Todd Stravitz, Ryan Taylor<br />

1762<br />

High CXCR-1 and CXCR-2 Expressing Neutrophils<br />

Induce Early Hepatocyte Death and promote Liver Injury<br />

in Acute-on-Chronic Liver Failure<br />

Arshi Khanam 1 , Nirupma Trehanpati 1 , Peggy Riese 2 , Archana Rastogi<br />

1 , Carlos A. Guzman 2 , Shiv K. Sarin 1 ; 1 research, Institute of<br />

liver and biliary sciences, New Delhi, India; 2 Helmholtz Centre for<br />

Infection Research, Braunschweig, Germany<br />

Background and aims: Acute-on-chronic liver failure (ACLF) is a<br />

serious and often fatal liver disease. Mechanisms of liver injury<br />

and development of sepsis in these patients are unclear. Neutrophils<br />

contribute to disease pathogenesis in acute liver failure<br />

and serve as a biomarker of organ dysfunction. We investigated<br />

the mechanisms of CXCR-1 and CXCR-2 expressing<br />

neutrophils mediated hepatic injury in ACLF patients (alcohol<br />

and hepatitis B related) and compared with chronic hepatitis<br />

B (CHB) and healthy controls (HC). Methods: CXCR-1 and<br />

CXCR-2 expressions on neutrophils were measured by flowcytometry,<br />

immunohistochemistry and RT-PCR. In vitro co-culture<br />

assays were performed to determine the mechanisms of hepatic<br />

injury and cell death. Mechanism of contact dependant cell<br />

death was investigated by co-culture of neutrophils with HepG2<br />

and HepG2.2.15 cells. Contact independent cell death was<br />

checked by culture of HepG2 and HepG2.2.15 cells with LPS<br />

stimulated neutrophil’s supernatant. In patients, cell death was<br />

determined by intrahepatic caspase-3 (apoptosis) and recep-


1068A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tor-interacting-protein kinase 3 (RIP-3; necrosis marker) expressions.<br />

To confirm the involvement of CXCR-1 and CXCR-2 in<br />

hepatocyte death, CXCR-1 and CXCR-2 blockade assay was<br />

done using SCH527123 antagonist. Results: High CXCR-1<br />

and CXCR2 expressing neutrophils were markedly increased in<br />

ACLF patients (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1069A<br />

CK19, and γH2AX and p21 expression verified DNA damage<br />

and related perturbations with incomplete liver regeneration.<br />

Phosphoprotein analysis of differences in injured versus healthy<br />

livers using normalization with housekeeping gene products<br />

and >2-fold up- or downregulated phospho- versus total proteins<br />

indicated abnormalities in ATM signaling pathways<br />

(e.g., ATM, Chek2, CDKN1a, CDNK1b, E2F1, Gadd45,<br />

Rad51, p53, p73), cell cycle regulators (e.g.,CCNB1, CCNC,<br />

CCND3, CCNE1, CDC25c, Cdk1, Cdk2, Cdk, Cdk7) besides<br />

cell stress, inflammation, etc. IPA mapping of these and other<br />

proteins indicated dysregulations in DNA damage/ repair and<br />

G0/G1 and G2/M checkpoint controls. In cultured HuH-7<br />

cells, APAP cytotoxicity reproduced double-strand DNA breaks,<br />

dysregulation in ATM signaling and G0/G1 and G2/M checkpoint<br />

restrictions. Conclusions: APAP-induced ALF in human livers<br />

was characterized by replicative stress and arrest of cells in<br />

G0/G1 due to DNA damage with dysregulated ATM signaling<br />

and cell cycle checkpoint controls. These molecular alterations<br />

offer therapeutic targets for drug development with or without<br />

cell therapy to restore liver regeneration in ALF.<br />

Disclosures:<br />

The following authors have nothing to disclose: Preeti Viswanathan, Yogeshwar<br />

Sharma, Sriram Bandi, Sanjeev Gupta<br />

The baseline characteristics of all paracenteses episodes before<br />

and after propensity score matching<br />

Disclosures:<br />

The following authors have nothing to disclose: Su Lin<br />

1765<br />

Lower fibrinogen level predicts hemorrhagic complications<br />

following abdominal paracentesis in patients<br />

with acute-on-chronic liver failure: A propensity score<br />

analysis<br />

Su Lin; Liver Research Center of the First Affiliated Hospital of<br />

Fujian Medical University, Fuzhou, China<br />

Object Patients with acute on chronic liver failure (ACLF) usually<br />

present severe coagulopathy. Abdominal paracentesis is<br />

often performed in those patients. The aim of this study was to<br />

analyze the prevalence of hemorrhagic events after abdominal<br />

paracentesis and the predictive factors of it in ACLF populations.<br />

Methods Patients who were diagnosed as ACLF from<br />

1 January 2010 to 31 December 2014 were enrolled. Those<br />

who had any signs of ascites underwent routine paracentesis<br />

upon and after admission. A propensity score matching<br />

analysis (greedy nearest neighbor matching) was used to<br />

select matched cases from overall non-hemorrhagic group as<br />

a control group. Hemorrhagic complications and the risk factors<br />

were examined using bilinear logistic regression analysis.<br />

Results A total of 602 abdominal paracenteses were carried<br />

out on 218 ACLF patients within a 5-year period. A total of<br />

18 (2.99%) hemorrhagic complications were identified, with<br />

4 cases of abdominal wall hematomas and 14 cases of intraperitoneal<br />

hemorrhage. We finally matched 18 cases with<br />

bleeding events to 72 unique cases without bleeding cases.<br />

The patients with hemorrhagic events had significantly lower<br />

fibrinogen levels and higher PT level than non-hemorrhagic<br />

ones. Bilinear regression revealed that lower fibrinogen level<br />

could independently predict hemorrhagic complication (OR:<br />

0.128, 95% CI 0.023-0.697, p = 0.017). The best cut-off<br />

value of fibrinogen level f>or predicting bleeding events>was 0.70 g/L,<br />

with the sensitivity of 76.4% and specificity of 80.0%. The area<br />

under curve was 0.733 (95% CI 0.604-0.862, P =0.002).<br />

Conclusion Severe hemorrhagic complications happened in<br />

2.99% of all paracenteses performed for ACLF patients. Low<br />

fibrinogen level (≤0.70g/L) is an independent predictor.<br />

1766<br />

Does weight loss surgery predispose to acetaminophen-related<br />

acute liver failure?<br />

Shannan Tujios 2 , Michelle Gottfried 1 , Valerie L. Durkalski 1 , William<br />

M. Lee 2 ; 1 Medical University of South Carolina, Charleston,<br />

SC; 2 University of Texas Southwestern, Dallas, TX<br />

Background: Obesity and bariatric surgery are increasingly<br />

common with nearly 180,000 Americans undergoing weight<br />

loss surgery (WLS) annually; an estimated 4.5 million individuals<br />

in the United States have had bariatric surgery to date. A<br />

previous study (J Clin Gastro 2014 epub) suggested that prior<br />

WLS was over-represented among patients with acetaminophen<br />

(APAP) related acute liver injury or acute liver failure<br />

(ALI/ALF). Aims: To describe the presentation, etiology and<br />

outcome of ALI/ALF patients with prior WLS, compared to a<br />

large group of ALF patients with acetaminophen (APAP) overdose<br />

without prior WLS. Methods: 57 patients with prior WLS<br />

enrolled in the US ALFSG registry between 01/1998 and<br />

04/2015 were reviewed and compared to 1234 remaining<br />

non-WLS APAP patients. Results: During the early study era<br />

(1998-2009), 1.5% of all cases (N=1678) reported prior WLS,<br />

while 3.2% of all cases since (2010-2015) (N=997) had prior<br />

WLS. Those with prior WLS were largely female (96.5%), and<br />

white (84.2%), and most demonstrated advanced coma grades<br />

(3-4; 75%) during hospitalization; 43 (75.4%) were APAP-related,<br />

3 (5.3%) due to idiosyncratic drug injury and 4 (7%)<br />

indeterminate. Nearly 25% of WLS APAP patients were listed<br />

for liver transplant with 9 patients (64% of those) receiving a<br />

graft; 21 day spontaneous survival was 58.5%, compared to<br />

unadjusted overall survival of 61.4%. Comparing the cohort<br />

of 43 APAP patients with prior WLS to the remaining APAP<br />

patients, there were no differences in age, race or reported<br />

alcohol use. The WLS group was more often female (95.4% vs<br />

73.7%, p=0.0014), and was more anemic (median hemoglobin<br />

9.5 gm/dL vs 11.2 gm/dL p=


1070A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

coma grade (74.3% vs 63.5%), percent waitlisted for transplant<br />

(23.3% vs 20.3%), fraction transplanted (14% vs 7.3%,<br />

p=0.2564) or 21 day spontaneous survival (64.1% vs 70.2%),<br />

overall unadjusted survival (65.1% vs 65.2%). Summary/Conclusions:<br />

The fraction of patients with ALF who have had prior<br />

WLS has been increasing in recent years as WLS becomes<br />

more commonplace. When ALF occurs in a patient with prior<br />

WLS, APAP toxicity is the most likely but not the only diagnosis<br />

observed. Bariatric surgery may predispose women to severe<br />

unintentional liver injury at relatively low APAP doses although<br />

outcomes appear similar with and without WLS.<br />

Disclosures:<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

The following authors have nothing to disclose: Shannan Tujios, Michelle<br />

Gottfried, Valerie L. Durkalski<br />

1767<br />

Dual Role of Epidermal Growth Factor Receptor (EGFR)<br />

in Liver Injury and Regeneration after Acetaminophen<br />

Overdose in Mice<br />

Bharat Bhushan, Michael Manley, Mitchell R. McGill, Hartmut<br />

Jaeschke, Udayan Apte; Pharmacology, Toxicology & Therapeutics,<br />

University of Kansas Medical Center, Kansas City, KS<br />

Acetaminophen (APAP) overdose is the major cause of acute<br />

liver failure in the US with very limited treatment options.<br />

Compensatory liver regeneration following APAP toxicity is a<br />

critical determinant in final recovery and survival. Epidermal<br />

Growth Factor Receptor (EGFR) signaling is known to play an<br />

important role in liver regeneration after partial hepatectomy.<br />

However, role of EGFR signaling in APAP toxicity and accompanying<br />

compensatory regeneration is completely unknown.<br />

We investigated the role of EFGR in APAP-induced liver injury<br />

and subsequent liver regeneration using pharmacological inhibition<br />

of EFGR signaling in mice by treatment with Canertinib,<br />

a water soluble irreversible EGFR inhibitor. EGFR was remarkably<br />

activated as early as 30 minutes after APAP treatment in<br />

a dose-dependent manner and remained activated up to 24 hr<br />

after APAP treatment. Treatment with Canertinib 1 hr post-APAP<br />

caused robust inhibition of EGFR activation resulting in remarkable<br />

attenuation of APAP-induced liver injury. Canertinib treatment<br />

did not alter APAP-protein adduct formation indicating<br />

APAP metabolism was not altered by EGFR inhibitor treatment.<br />

Interestingly, APAP mediated JNK activation (a key mediator<br />

of APAP toxicity) was not altered by EGFR inhibitor treatment.<br />

In order to study role of EGFR in liver regeneration after APAP<br />

overdose, independent of injury, EGFR inhibitor was administered<br />

12 hr after APAP treatment, which resulted in remarkable<br />

impairment of liver regeneration response without any alteration<br />

of injury. Decreased liver regeneration was associated<br />

with decreased induction of cyclin D1 and decreased phosphorylation<br />

of Rb protein. In conclusion, our study revealed,<br />

for the first time, that EGFR receptor plays an important role in<br />

development of APAP injury but also has a direct role in stimulation<br />

of liver regeneration after APAP overdose. These data<br />

support the hypothesis that EGFR signaling plays dual role in<br />

APAP-induced hepatocyte death and proliferation in liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: Bharat Bhushan, Michael Manley,<br />

Mitchell R. McGill, Hartmut Jaeschke, Udayan Apte<br />

1768<br />

Heat Stroke Leading to Acute Liver Failure: Case Series<br />

from the Acute Liver Failure Study Group<br />

Brian C. Davis 1 , Holly Tillman 2 , Robert J. Fontana 3 , K. Rajender<br />

Reddy 4 , Raymond T. Chung 5 , Brendan M. McGuire 6 , R. Todd Stravitz<br />

7 , William M. Lee 1 ; 1 Division of Digestive and Liver Diseases,<br />

UT Southwestern Medical Center, Dallas, TX; 2 Department of Public<br />

Health Sciences, Medical University of South Carolina, Charleston,<br />

SC; 3 Division of Gastroenterology, University of Michigan Health<br />

System, Ann Arbor, MI; 4 Division of Gastroenterology, University<br />

of Pennsylvania Perelman School of Medicine, Philadelphia, PA;<br />

5 Department of Gastroenterology, Massachusetts General Hospital,<br />

Boston, MA; 6 Division of Gastroenterology and Hepatology,<br />

UAB School of Medicine, Birmingham, AL; 7 Division of Gastroenterology,<br />

VCU Medical Center, Richmond, VA<br />

Introduction: In the United States, a small percentage of cases<br />

of acute liver injury (ALI) and failure (ALF) have been associated<br />

with heat stroke, defined as elevated body temperature<br />

≥40°C with central nervous system dysfunction. This may occur<br />

as classical heat stroke (CHS), defined as exposure to high<br />

environmental temperatures, or exertional heat stroke (EHS),<br />

which occurs after strenuous activity. The aim of this study was<br />

to describe cases of ALI or ALF caused by heat stroke in a large<br />

ALF registry. Methods: Amongst 2,675 consecutive ALI and ALF<br />

subjects enrolled between January 1998 and April 1, 2015,<br />

there were 8 subjects with heat stroke, classified as having<br />

CHS or EHS. The clinical, laboratory, and outcome data were<br />

then compared. Results: Five patients had ALF and 3 patients<br />

had ALI. Six patients had EHS, one with CHS, and one was<br />

unclassified. Seven patients developed acute renal failure, 8<br />

had lactic acidosis, and 6 had rhabdomyolysis. Six patients<br />

underwent cooling treatments, 3 received N-acetyl cysteine<br />

(NAC), 3 required mechanical ventilation, 3 required renal<br />

replacement therapy, 1 underwent liver transplantation, and 2<br />

patients died—both within a day of presentation. All but one<br />

case occurred between June and August (May). No long-term<br />

sequelae were observed in survivors. Conclusions: In the US,<br />

heat stroke accounts for


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1071A<br />

Raymond T. Chung - Grant/Research Support: Gilead, Mass Biologics, Abbvie,<br />

Merck, BMS<br />

Brendan M. McGuire - Grant/Research Support: bayer healthcare, salix<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

The following authors have nothing to disclose: Brian C. Davis, Holly Tillman, R.<br />

Todd Stravitz<br />

1769<br />

Recovery of severe critically ill Acute Liver Failure (ALF)<br />

patients is associated with significantly better survival<br />

than patients undergoing liver transplantation<br />

Antonella Putignano 1 , Francesco Figorilli 1 , Banwari Agarwal 2 ,<br />

Rajiv Jalan 1 ; 1 Liver failure group, UCL Institute for Liver and Digestive<br />

Health, UCL Medical School, Royal Free Campus,, London,<br />

United Kingdom; 2 Intensive Care Unit, Royal Free Hospital, Royal<br />

Free Hampstead NHS Trust, University College London, London,<br />

United Kingdom<br />

Introduction: In patients with ALF fulfilling poor prognostic criteria,<br />

mortality rates without liver transplantation (LT) are high. In<br />

a recent study (1), unexpectedly, even the mortality of patients<br />

with ALF who recovered spontaneously was reported to be<br />

high. Due to the multicenter nature of the study, long-term follow<br />

up data for many patients was not defined. In this single<br />

center study, we aimed to confirm or refute this observation in<br />

ALF patients admitted to a single intensive care unit. Methods:<br />

Consecutive ALF patients admitted between 1990-2014 were<br />

considered and the 90-day survivors (Early Survivors) included<br />

to evaluate the 3-year outcome. According to ALF etiology and<br />

management, 4-cohorts were identified: Paracetamol Overdose<br />

(POD) Transplanted, POD Not-Transplanted, Non-POD<br />

Transplanted and Non-POD Not-Transplanted. Chi Square or<br />

Fisher test were used to compare outcome between cohorts<br />

(p< .05) and Kaplan Meier analysis to estimate their cumulative<br />

survival. Predictors of 3-year mortality were modeled<br />

with Cox Regression analyses. Causes of mortality were also<br />

identified. Results: 200 ALF patients were admitted to intensive<br />

care unit, and 124 were Early Survivors. 112/124 reached<br />

the 3-year follow up and were included in this study (Female<br />

77, age 35,4 + 11.7, POD 58). 13/112 POD Transplanted<br />

(11.6%), 45/112 POD Not-Transplanted (40.2%), 30/112<br />

Non-POD Transplanted (26.8%) and 24/112 Non-POD<br />

Not-Transplanted (21.4%) were identified (p


1072A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1771<br />

Cathepsin L-deficiency enhances liver regeneration after<br />

partial hepatectomy<br />

Toshifumi Sato, Shunhei Yamashina, Kousuke Izumi, Hirofumi<br />

Fukushima, Yoshihiro Inami, Tomonori Aoyama, Akira Uchiyama,<br />

Kazuyoshi Kon, Kenichi Ikejima, Sumio Watanabe; Department of<br />

Gastroenterology, Juntendo University School of Medicine, Tokyo,<br />

Japan<br />

Background Cathepsin L (CTSL), known as a highly potent<br />

endoprotease of the cysteine cathepsin family, plays a pivotal<br />

role on lysosomal protein degradation. It was reported that<br />

the activity of collagenolytic cathepsins including CTSL was<br />

enhanced in the CCl 4<br />

-induced fibrotic liver. On the other hand,<br />

previous investigations revealed lysosomal proteolysis is suppressed<br />

accompanied by decreased biosynthesis of cathepsins<br />

in liver after partial hepatectomy. The role of decreased CTSL<br />

level in liver regeneration has been still unclear. We investigate<br />

to clarify if CTSL-deficiency affects liver regeneration after partial<br />

hepatectomy. Methods 70% of partial hepatectomy (PH)<br />

was performed in male CTSL-deficient (Ctsl-/-) mice and wildtype<br />

littermates (Ctsl +/+). Mice were sacrificed and wet weight<br />

of the whole remaining liver was measured after 2 and 5days.<br />

Restituted liver mass was expressed as percentage of the ratio<br />

of remaining liver divided by calculated-total pre-hepatectomy<br />

liver weight. Bromodeoxyuridine (BrdU)-immunostaining of liver<br />

sections was performed. Expression of cyclin D1 in the liver<br />

was evaluated by western blot analysis. Results Restituted liver<br />

mass at 5day after PH in Ctsl-/- mice was higher than wild-type<br />

mice significantly (p=0.008). BrdU incorporation into nucleai<br />

was observed in 5.16±0.90% of hepatocytes 48 hours after<br />

PH in wild-type mice; however, the rate of BrdU incorporation<br />

was significantly enhanced to 8.82±0.35% in Ctsl-/- mice. Protein<br />

level of CyclinD1 was increased in wild-type mice 48 hours<br />

after PH. Cathepsin-deficiency enhanced increase in cyclinD1<br />

expression after PH to about 3-fold of wild-type. Conclusion<br />

CTSL-deficiency accelerates regenerating response in the liver<br />

after PH. These findings suggested that the decreasing CTSL<br />

plays an important role on liver regeneration. Therefore, an<br />

increase in CTSL activity observed in CCl 4<br />

-induced fibrotic liver<br />

may attenuate liver regeneration. Taken together, CTSL might<br />

be a new therapeutic target on disorder of liver regeneration<br />

in hepatic cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Toshifumi Sato, Shunhei Yamashina,<br />

Kousuke Izumi, Hirofumi Fukushima, Yoshihiro Inami, Tomonori Aoyama,<br />

Akira Uchiyama, Kazuyoshi Kon, Kenichi Ikejima, Sumio Watanabe<br />

1772<br />

The Receptor Interacting Protein Kinase 3 is a critical<br />

early warning signs for onset process of acute-onchronic<br />

hepatitis B liver failure<br />

Yubao Zheng, Yurong Gu, Ke Wang, Min Zhang, Xiaohui Huang,<br />

Ziyu Lin, Liang Peng, Zhiliang Gao; Department of Infectious Diseases,<br />

The Third Affiliated Hospital, Sun Yat-sen University, Guangzhou,<br />

China<br />

Background and aims: Hepatitis B virus is a principal cause of<br />

Acute-on-chronic hepatitis B liver failure (ACHBLF) in the China.<br />

The pathophysiology of ACHBLF is incompletely understood.<br />

Recent findings suggest that the receptor interacting protein<br />

kinase 3 (RIP3) acts as a switch from apoptosis to necrosis.<br />

Thus, the aim of the current investigation was to determine<br />

if RIP3 is involved in onset process of ACHBLF or Hepatitis B<br />

virus-induced hepatocytel death. Methods: The expression of<br />

RIP3 and Tumor Necrosis Factor-α(TNF-α) were retrospectively<br />

investigated in liver tissues of 46 patients with ACHBLF and 78<br />

patients with chronic hepatitis B(CHB) at the China South Hepatology<br />

Center. Meanwhile, in vitro the apoptotic and necrotic<br />

levels were investigated in the HepG2.2.15 cell, the genetically<br />

modified RIP3-expressing HepG2.2.15 cells co-cultured with<br />

Natural killer(NK) cells or not. Result: Our results showed that<br />

the levels of RIP3 expressing in Liver tissues were significantly<br />

higher in ACHBLF than in CHB, as shown by immunohistochemical(IHC)<br />

and western blots. The TNF-α expression was similar<br />

to RIP3 that the expression levels of TNF-α increased more<br />

significantly in ACHBLF than in CHB. The rate of apoptosis in<br />

the genetically RIP3-over expressing HepG2.2.15 cells were<br />

significantly higher in HepG2.2.15 cells and the genetically<br />

RNA silencing RIP3 HepG2.2.15 cells co-cultured with NK<br />

cells in vitro. Result of cell experiments show that RIP3 protein<br />

effectively regulates the apoptosis and necrosis of HepG2.2.15<br />

co-cultured NK cells. Conclusion: Our study suggest that RIP3<br />

protein effectively regulates apoptotic of HepG2.2.15 co-cultured<br />

NK cells. And RIP3 is a critical early warning signs for<br />

onset process of ACHBLF. RIP3 may be a novel therapeutic<br />

target in the treatment of ACHBLF in the future.<br />

Figure:RIP3 expressing in Liver tissues and the rate of apoptosis in<br />

HepG2.2.15 cells co-cultured NK cells.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yubao Zheng, Yurong Gu, Ke<br />

Wang, Min Zhang, Xiaohui Huang, Ziyu Lin, Liang Peng, Zhiliang Gao<br />

1773<br />

Nitric Oxide induced hepatic stellate cells apoptosis<br />

through autophagy inhibition in acute liver failure<br />

Yingli He 3 , Li Jin 1 , Jiuping Wang 2 , Xiaogang Zhang 5 , Jing Wang 1 ,<br />

Jinfeng Liu 3 , Yuan Yang 3 , Taotao Yan 1 , Tianyan Chen 3 , Furong<br />

Cao 1 , Xiaoni Kou 4 , Yingren Zhao 3 ; 1 School of Medicine, Xi’an<br />

Jiaotong University, Xi’an, China; 2 Department of Infectious Diseases,<br />

Xijing Hospital, Fourth Military Medical University, Xi’an,<br />

China; 3 Department of Infectious Diseases, First Affiliated Hospital,<br />

School of Medicine, Xi’an Jiaotong University, Xi’an, China;<br />

4 Department of Hepatology, First Affiliated Hospital, Shaanxi<br />

University of Chinese Medicine, Xianyang, China; 5 Department<br />

of Hepatobiliary Surgery, First Affiliated Hospital, School of Medicine,<br />

Xi’an Jiaotong University, Xi’an, China<br />

Background & Aim Previously we have shown acute liver failure<br />

(ALF) accompanied by fibrosis, a process modulated by<br />

autophagy induction and activation of hepatic stellate cells<br />

(HSCs) and eventually favor for prognosis of ALF. Nitric oxide<br />

(NO), a multipotent mediator, has been shown to promote<br />

apoptosis in HSCs. However, the role NO played in ALF and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1073A<br />

underlying mechanisms remain unclear. Methods ALF patients<br />

were recruited to investigate the correlation between plasma<br />

NO level and clinical feature. Plasma nitrite was measured<br />

using the Griess reaction. The expression of nitric oxide synthases<br />

(iNOS, eNOS and nNOS), HSCs activation (alpha-<br />

SMA), and autophagic activity (LC3) in liver were investigated<br />

using immunohistochemistry. HSCs were treated with NO<br />

donor, and then autophagy flux was investigated by immunoblotting<br />

and confocal microscopy. Apoptosis was quantified<br />

by characteristic DAPI stain nuclei. Results Firstly, plasma NO<br />

level is significantly increased in patients with ALF as compared<br />

with cirrhosis (53.60±19.74 μmol/L vs 19.40±9.03<br />

μmol/L, t=10.57, p


1074A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

The following authors have nothing to disclose: Jody A. Rule, Linda S. Hynan,<br />

Nahid Attar, William J. Korzun<br />

Disclosures:<br />

William Bernal - Advisory Committees or Review Panels: Ocera Therapeutics ;<br />

Consulting: Vital Therapies Inc<br />

Julia Wendon - Consulting: Pulsion, Excalenz<br />

The following authors have nothing to disclose: Georg Auzinger, Robert Loveridge,<br />

Sameer Patel, Christopher Willars, Krishna Menon, Hector Vilca-Melendez<br />

1776<br />

Serial Procalcitonin Measurements in Acute Liver Failure<br />

Patients with Bacterial Infections<br />

Jody A. Rule 1 , Linda S. Hynan 1 , Nahid Attar 1 , William J. Korzun<br />

2 , William M. Lee 1 ; 1 Internal Medicine, University of Texas<br />

Southwestern Medical Center, Dallas, TX; 2 Allied Health, Virginia<br />

Commonwealth University, Richmond, VA<br />

Background: Multi-organ failure in ALF patients resembles<br />

severe sepsis and septic shock, making identification of bacterial<br />

infection can be difficult. Procalcitonin (PCT) has proven<br />

useful in sepsis in general to identify and predict response<br />

to infection. We hypothesized that serial PCT values in ALF<br />

patients might be predictive of outcome. Methods: Among ALF<br />

Study Group 47 patients with bacterial infection (positive culture<br />

on Day 1), we compared serial PCT levels in sera available<br />

at least for Days 1-4, using the ADVIA Centaur BRAHMS PCT<br />

assay, a sandwich chemi-luminescent immunoassay, comparing<br />

change in levels to outcome as spontaneous survival (SS)<br />

vs. death/transplantation (non-SS). Results: Elevated PCT levels<br />

were observed in all patients (median 2.31 ng/mL), falling rapidly<br />

in the SS group from median of 2.9 to 1.07 ng/mL by Day<br />

4, compared to the non-SS group that actually increased from<br />

1.42 to 2.04 ng/mL by Day 4. Both groups had an overall<br />

decrease by Day 7 of 70.3% (SS) and 49.3% (non-SS). Conclusions:<br />

High PCT levels are characteristic of most ALF subjects<br />

falling to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1075A<br />

1778<br />

Early elevation in serum IFN-λ3 is responsible for progressive<br />

liver failure caused by acute infection of Hepatitis<br />

E virus genotype 4<br />

Jong-Hon Kang 1 , Kazumoto Murata 2 , Yoshiyasu Karino 3 , Takeshi<br />

Matsui 1 , Itaru Ozeki 3 , Kazuaki Takahashi 4 , Masahiro Arai 4 ,<br />

Masashi Mizokami 2 ; 1 Center for Gastroenterology, Teine Keijinkai<br />

Hospital, Sapporo, Japan; 2 The Research Center for Hepatitis and<br />

Immunology, National Center for Global Health and Medicine,<br />

Tokyo, Japan; 3 Department of Hepatology, Sapporo Kosei General<br />

Hospital, Sapporo, Japan; 4 Department of Medical Sciences,<br />

Toshiba General Hospital, Tokyo, Japan<br />

Background / Aims: Acute sporadic infection of hepatitis E virus<br />

(HEV) has been emerging in non-endemic countries because of<br />

increasing clinical impacts. Hepatitis E virus (HEV) of genotype<br />

(Gt) 4 was showed to be more responsible for development of<br />

acute liver failure (ALF) compared with Gt 3 in our previous<br />

report (abstract ID; 745, AASLD 2014). However, the relationships<br />

between genotypic differences and disease severity have<br />

been obscure. Interferon (IFN)-λ3 has been recognized as one<br />

of markers for innate immune responses against viral infection,<br />

such as HCV infection. The aim of this study is to clarify the<br />

mechanism of disease progression in acute HEV of Gt 4 infection<br />

by the analysis of serum IFN-λ3 levels. Methods: Among<br />

98 patients with sporadic and autochthonous hepatitis E diagnosed<br />

in Hokkaido, Japan, a total of 43 patients in whom early<br />

sera were sampled during the day 2 to 14 after the onset were<br />

enrolled. Acute HEV infection was diagnosed upon the detection<br />

of HEV RNA by PCR or anti-HEV antibody (IgM) in sera by<br />

enzyme linked immune-sorbent assay (ELISA). HEV genotypes<br />

(Gt) were determined by comparison of a 326-nt sequence<br />

within ORF1 of HEV genome. IFN-λ3 levels were measured<br />

by chemiluminescence ELISA and HEV RNA were quantified<br />

by real time PCR for the same sera obtained from the patients.<br />

Acute liver failure (ALF) was defined to be a case with longer<br />

prothrombin time (INR>1.5). Results: Out of 43 patients with<br />

acute HEV (32 males, median age 56 years) thus collected, 15<br />

developed ALF, in which 2 presented hepatic coma. HEV Gt<br />

4 was determined in 27 patients, Gt 3 in 13, co-infection with<br />

Gt 3+4 in 1, and not determined in 2. IFN-λ3 level was 11.5<br />

(1.5-120.4) pg/ml at Day 4.5 (1-13) in median. In Gt 4 cases,<br />

median peak levels in AST and ALT were higher than those with<br />

Gt 3, respectively, and, as previously reported, Gt 4 infection<br />

was correlated with development of ALF in comparison with Gt<br />

3 (14/15 vs. 1/15, p=0.002, OR 16.2[95%CI, 1.9-140.1]).<br />

Serum IFN-λ3 levels were higher in Gt 4 cases than those in Gt<br />

3 cases (15.9 vs. 5.0 pg/ml, p=0.001), despite of no difference<br />

in sampling timing of sera between Gts. Among 27 cases<br />

with Gt 4 infection, IFN-λ3 levels were higher in ALF cases than<br />

in self-limited (26.3 vs. 12.8 pg/ml, p=0.046), and IFN-λ3 levels<br />

revealed correlation with HEV RNA titrations in sera in Gt 4<br />

(r=0.701, p=0.008). Conclusion: Our study may implicate the<br />

pathological roles of IFN-λ3 in acute hepatitis E especially ALF<br />

due to HEV Gt 4 infection, and serum IFN-λ3 would be a useful<br />

predictive marker for disease-severities.<br />

Disclosures:<br />

Yoshiyasu Karino - Speaking and Teaching: BMS KK<br />

The following authors have nothing to disclose: Jong-Hon Kang, Kazumoto<br />

Murata, Takeshi Matsui, Itaru Ozeki, Kazuaki Takahashi, Masahiro Arai,<br />

Masashi Mizokami<br />

1779<br />

Expression of Krüppel-like factor 6 parallels induction of<br />

autophagy in acute liver injury<br />

Svenja Sydor 1 , Jan Best 1 , Paul P. Manka 1 , Sami Jafoui 1 , Martin<br />

Schlattjan 1 , Francisco Javier Cubero 2 , Thomas Schreiter 1 , Diana<br />

Vetter 3 , Andreas Paul 4 , Scott L. Friedman 5 , Guido Gerken 1 , Ali<br />

Canbay 1 , Lars Bechmann 1 ; 1 Department of Gastroenterology and<br />

Hepatology, University Hospital Essen, Essen, Germany; 2 Department<br />

of Medicine III, Gastroenterology, Metabolic Disease and<br />

Intensive Care, University Hospital RWTH Aachen, Aachen, Germany;<br />

3 Department of Surgery, University Hospital Zurich, Zurich,<br />

Switzerland; 4 Department of General- and Transplant -Surgery,<br />

University Hospital Essen, Essen, Germany; 5 Division of Liver Diseases,<br />

Mount Sinai School of Medicine, New York, NY<br />

Krüppel-like factor 6 (KLF6) is a ubiquitously expressed, multifunctional<br />

transcription factor and tumor suppressor gene regulating<br />

hepatocyte glucose and lipid homeostasis and KLF6<br />

down-regulation is associated with proliferation in hepatocellular<br />

cancer. Autophagy regulates turnover of long-lived or<br />

damaged organelles and proteins, which affects liver regeneration<br />

by maintaining intracellular energy production. With<br />

this study, we characterized the role of KLF6 in liver injury<br />

and regeneration and found significant correlations with the<br />

induction of autophagy. KLF6 expression was quantified in<br />

liver samples from 10 patients with acute liver failure (ALF)<br />

and 10 controls. Human liver preparations were treated with<br />

acetaminophen (APAP) up to 30 hours in an established ex<br />

vivo perfusion model. Furthermore, C57B/6 mice were sacrificed<br />

8h after APAP injection (6 mice/group) and HepG2<br />

cells were incubated with APAP for 24h. We characterized<br />

the role of KLF6 in liver regeneration in vivo in a model of partial<br />

hepatectomy (PHx) in wildtype (wt) or hepatocyte-specific<br />

KLF6 knockout mice (dKLF6) (6 mice/group). We performed<br />

an Affymetrix gene array (HT-MG4306) on liver samples of<br />

dKLF6 and wt mice at 12h after PHx. KLF6 expression was significantly<br />

induced in hepatocytes from ALF patients compared<br />

to controls. APAP perfusion of liver tissue significantly induced<br />

KLF6 expression and KLF6 was significantly upregulated in<br />

APAP treated mice. The gene array revealed significant alterations<br />

in autophagy related genes in dKLF6 mice compared to<br />

wt following PHx. Hepatocyte proliferation was significantly<br />

induced at early time points after PHx in dKLF6 mice compared<br />

to wt. Western blot analyses of whole liver tissue from dKLF6<br />

mice revealed reduced levels of LC3-II 72h after PHx. dKLF6<br />

mice had decreased expression levels of autophagy related<br />

molecules Atg7 and Beclin1 before and after PHx compared to<br />

wt mice. HepG2 cells and murine liver tissue showed increased<br />

LC3-II levels following APAP treatment. Here, we observed a<br />

consistent induction of hepatic KLF6 expression in various models<br />

of acute liver injury. Results from hepatocyte-specific KLF6<br />

knockout mice implicate a significant role for KLF6 in early<br />

hepatic regeneration. Furthermore, gene array data as well as<br />

<strong>studies</strong> on mRNA and protein expression in mice, humans and<br />

cell culture reveal a novel interaction between KLF6 and autophagy<br />

in acutely injured hepatocytes. The underlying mechanisms<br />

need further characterization.<br />

Disclosures:<br />

Jan Best - Speaking and Teaching: BTG<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma


1076A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Svenja Sydor, Paul P. Manka,<br />

Sami Jafoui, Martin Schlattjan, Francisco Javier Cubero, Thomas Schreiter, Diana<br />

Vetter, Andreas Paul, Guido Gerken, Ali Canbay, Lars Bechmann<br />

1780<br />

Liver specific organ failure scoring systems, CLIF-OF and<br />

CLIF-ACLF are better than King’s College Hospital criteria<br />

in predicting mortality in patients with Acute Liver<br />

Failure (ALF) admitted to Intensive Care Unit<br />

Francesco Figorilli 2,1 , Antonella Putignano 2 , Banwari Agarwal 3 ,<br />

Rajiv Jalan 2 ; 1 Universita’ di Cagliari, Cagliari, Italy; 2 UCL Medical<br />

School, Royal Free Campus, Liver failure group, UCL Institute for<br />

Liver and Digestive Health, London, United Kingdom; 3 Intensive<br />

Care Unit, Royal Free Hospital, Royal Free Hampstead NHS Trust,<br />

University College London, London, United Kingdom<br />

Background. ALF is a critical illness and a liver transplant (LT)<br />

is often the only life-saving treatment. King’s College Hospital<br />

criteria (KCH) are currently used to identify LT potential candidates.<br />

Given high mortality and the organ shortage, a more<br />

sensitive score is needed. In this study, nine different exiting<br />

scores (MELD, UKELD, iMELD, APACHE2, APACHE3, SOFA,<br />

KCH, CLIF-OF and CLIF-ACLF) were assessed at the admission<br />

to find the best predictor of mortality. Methods. Transplanted<br />

and non-transplanted patients were analysed separately and<br />

further stratified in base of aetiology (Paracetamol overdose<br />

(POD) or other causes (Non-POD)). Primary endpoint was the<br />

3-month mortality of any cause. For each score we assessed the<br />

area under the curve (AUC), the best cut-off value, sensitivity,<br />

specificity and positive predictive value (PPV). Cox analysis<br />

was used to identify the best predictor of 3-month mortality.<br />

Results. 200 patients admitted between 1990-2014 in our<br />

centre were included in this study (98 POD, female 65%, mean<br />

age 38.3±12,8). 70/200 received a LT (19 POD, 51 Non-<br />

POD), and 22 (31.4%) died within 3 months. In the Non LT<br />

cohort (79 POD and 51 NON POD), the mortality rate was<br />

41.5%. 90 patients fulfilled poor prognostic KCH criteria but<br />

did not receive a LT (12 were not listed for psychiatric problem,<br />

29 was considered too unwell for a LT and 31 spontaneously<br />

recovered). 18/90 were listed but 12 died on the waiting list<br />

and 6 improved. In Non-LT/Non-POD group CLIF-ACLF had<br />

the best AUC (0.799) with 73.9% sensitivity and 70.8% PPV;<br />

also it was the best mortality predictor in the multivariate analysis<br />

(p=0.001; HR 1.09; 95%CI 1.04-1.16). In Non-LT/POD<br />

CLIF-OF showed the second highest AUC (0.793) after SOFA<br />

(0.799) but with a higher sensitivity (93.5% vs 77.4%). In<br />

the univariate analysis CLIF-OF cut-off (12) had the highest<br />

Hazard Ratio (38.9). In Cox regression SOFA was the only<br />

significant in Non-LT/POD group (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1077A<br />

1782<br />

Histology predicts the need for liver transplantation in<br />

patients with acute severe autoimmune hepatitis<br />

Eleonora De Martin 1,2 , Audrey Coilly 1,2 , Mylene Sebagh 1,2 , Badr<br />

Serji 1 , Teresa Maria Antonini 1,2 , Eric Ballot 2,3 , Florent Artru 1 ,<br />

Philippe Ichai 1,2 , Didier Samuel 1,3 , Jean-Charles Duclos-Vallee 1,2 ;<br />

1 Centre Hepato-Biliaire, AP-HP Hopital Paul Brousse, Villejuif,<br />

France; 2 Unit 1193, Inserm, Villejuif, France; 3 Univ Paris-Sud,<br />

Villejuif, France<br />

Background: Acute severe autoimmune hepatitis (AS-AIH) is<br />

a rare, heterogeneous disease, with a high rate of liver transplantation<br />

(LT) or death and a poor determination of prognostic<br />

factors. Aim: The aim of the present study was to describe the<br />

prognostic role of chronic hepatitis for the outcome of patients<br />

presenting with AS-AIH. Methods: Patients admitted to our center,<br />

between January 2009 and June 2014, with the diagnosis<br />

of AS-AIH were retrospectively identified. The AS-AIH diagnosis<br />

was based on: 1. IAIHG score definite or probable, 2.<br />

INR > 1.5, 3. No previous medical history of AIH, 4. Histological<br />

features of AIH. A single expert pathologist reviewed the<br />

liver biopsies. Patients with chronic hepatitis (CH) defined by<br />

a fibrosis stage 2, according to METAVIR, were compared to<br />

patients without chronic hepatitis (non-CH). Results: Fourteen<br />

patients were included, 11 (78.5%) female, median age 52<br />

[18-69] years. CH was diagnosed in 8/14 (57%) patients on<br />

liver biopsy, among them 4 were cirrhotic. The histological<br />

evaluation was performed on the explant for 3 patients, all of<br />

them without CH. Median age of CH patients was 57 [18-69]<br />

years versus (vs) 44 [18-66] of non-CH patients (p=ns). CH<br />

patients compared to non-CH patients had: lower MELD score,<br />

27 [16-32] vs 33 [24-34] (p=0.02), lower bilirubin, 237<br />

mmol/L [26-363] vs 378 [241-541] (p=0.03), lower median<br />

INR 2.77 [1.68-4.04] vs 3.75 [1.9-4.29] (p=0.12) although<br />

the difference was not statistically significant. Patients of both<br />

groups had similar IgG levels, 32.3 [22.2-61.7] vs 31.9 [9.67]<br />

(p=ns), and all but one had positive antinuclear or anti-smooth<br />

muscle antibodies. Eleven (78.5%) patients received steroids<br />

therapy (0.5 to 1 mg/kg/day): 7/8 patients with CH versus<br />

4/6 patients without. The median interval time between hospital<br />

admission and steroids therapy was 3 [0-119] days for CH<br />

patients vs 9 [2-69] days for non-CH pts (p=ns). Infections were<br />

detected in 37.5% (3/8) of CH patients vs 50% (3/6) non-CH<br />

pts (p=ns). All patients who did not receive steroids therapy<br />

underwent LT. LT was performed in 25% (n=2) of CH patients<br />

vs 83% (n=5) of non-CH pts (p=0.03). Overall survival was<br />

100%, with a median follow-up of 24 [3-69] months. Conclusions:<br />

In patients with severe acute autoimmune hepatitis histological<br />

assessment has a key prognostic role as the absence<br />

of chronic hepatitis predicts the need of liver transplantation,<br />

despite steroids therapy.<br />

Disclosures:<br />

Audrey Coilly - Speaking and Teaching: Gilead, BMS, Janssen, MSD, Roche,<br />

Novartis, Astellas<br />

Didier Samuel - Consulting: Astellas, MSD, BMS, Roche, Novartis, Gilead, LFB,<br />

Janssen-Cilag, Biotest, Abbvie<br />

The following authors have nothing to disclose: Eleonora De Martin, Mylene<br />

Sebagh, Badr Serji, Teresa Maria Antonini, Eric Ballot, Florent Artru, Philippe<br />

Ichai, Jean-Charles Duclos-Vallee<br />

1783<br />

Hemophagocytic lymphohistiocytosis (HLH) presenting<br />

as acute liver failure: three cases at a liver transplant<br />

center<br />

Amanda Schneier 1 , Ponni V. Perumalswami 2 , Sunyoung Lee 3 ;<br />

1 Medicine, Icahn School of Medicine at Mount Sinai, New York,<br />

NY; 2 Liver Diseases, Icahn School of Medicine at Mount Sinai,<br />

New York, NY; 3 Medicine, Elmhurst Hospital, Queens, NY<br />

OBJECTIVE: Hemophagocytic lymphohistiocytosis (HLH) is<br />

primarily known as a pediatric disease but has been rarely<br />

described in adults, usually in the setting of EBV infection,<br />

malignancy, rheumatologic disease, or as a side effect of<br />

immune-modulating medications. The pathophysiology of HLH<br />

has been described as an uncontrolled activation of the immune<br />

system and a cytokine storm, resulting in excessive inflammation<br />

and phagocytosis of patients’ own blood cells by histiocytes.<br />

Patients with HLH typically present with develop fever,<br />

organomegaly, lymphadenopathy, liver dysfunction, and cytopenias<br />

in the absence of infection. METHODS: This case series<br />

describes three patients at an urban, academic liver transplant<br />

center. All cases presented with acute liver failure and were ultimately<br />

diagnosed with HLH secondary to occult malignancies.<br />

The cases presented from February 2014 to February 2015.<br />

RESULTS: All three cases were female patients ranging from 44<br />

to 53 years old. All were transferred from other hospitals for<br />

workup of severe jaundice and liver dysfunction, suspected to<br />

be secondary to toxin-induced liver injury. The patients went on<br />

to develop severe fevers and refractory cytopenias. All three<br />

underwent a rapid evaluation for liver transplant and were evaluated<br />

by a multidisciplinary team including hepatology, critical<br />

care, infectious disease, and hematology. A complete work up<br />

for acute liver failure including viral serologies, autoimmune<br />

markers, and ceruloplasmin were negative for all three cases,<br />

and none had significant alcohol consumption or other drug or<br />

toxin exposure. The diagnosis of HLH was made after a mean<br />

of 10 days by bone marrow biopsy. Bone marrow pathology<br />

featured histiocytes with hemophagocytosis in all patients. Two<br />

out of three cases also had evidence of hemophagocytosis on<br />

liver pathology (one upon autopsy). The patients’ initial ferritin<br />

levels ranged from 4,170 to 48,000 ng/mL, and all eventually<br />

rose to >50,000 ng/mL. All cases were treated with HLH-directed<br />

therapy (two with etoposide, one with pentostatin) but<br />

expired during their hospitalization. All cases were found to<br />

have prior exposure to EBV but undetectable viral loads during<br />

admission, and were subsequently found to have occult malignancies:<br />

two with hepatosplenic T-cell lymphoma and one with<br />

papillary thyroid carcinoma. CONCLUSIONS: HLH can present<br />

as acute liver failure in adult patients. Given the low success<br />

rate of treatment, early diagnosis of HLH is critical. Therefore,<br />

a higher degree of suspicion should be exercised in otherwise<br />

unexplained rapidly progressing acute liver failure in patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Amanda Schneier, Ponni V.<br />

Perumalswami, Sunyoung Lee<br />

1784<br />

Single Center Experience: Hypothermia in Acute Liver<br />

Failure Introduction<br />

Edson S. Franco 2 , Alexia Makris 1 , Manohar M. Lahoti 1 , Angel<br />

Alsina 2 ; 1 Information Systems and Decision Sciences, University<br />

of South Florida, Tampa, FL; 2 Surgery, Tampa General Hospital,<br />

Tampa, FL<br />

Introduction Acute liver failure is a rare condition with a dismal<br />

prognosis in the setting of hepatic encephalopathy (HE).


1078A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Hypothermia (HT) has been shown to prevent cerebral edema<br />

and intracranial hypertension associated HE. In this study, we<br />

report the outcomes of ALF patients treated with hypothermia<br />

with and without transplantation. Methods This is a single<br />

center retrospective study from 2002 to 2014. We reviewed<br />

all ALF patients treated with HT for HE with and without liver<br />

transplantation (Oltx). Demographic data was collected. Primary<br />

and secondary outcomes include survival and complications<br />

respectively. Survival was calculated using Kaplan-Meier<br />

method. Bootstrap analysis was used to perform an additional<br />

internal validation. Results 44 patients were treated for ALF<br />

during the study period. The cohort was stratified as follows:<br />

Group 1 (HT/Oltx), Group 2 (HT/No Oltx), Group 3 (Medical<br />

Management/Oltx). The indication for HT was grade III/IV<br />

HE and instituted at 32°-34°. Twenty patients (46.5%) were<br />

bridged to transplantation. Group 2 patients were evaluated<br />

for Oltx except for one and 3 were listed. Survival was poor<br />

with mean 20 days. One died while waiting and 2 were<br />

removed. Four patients (17%) improved. Overall survival was<br />

50% (mean 362 days). One year-survival was 75% in Group<br />

1 versus 58% for Group 3 when patients received transplant.<br />

At the 2.5 years post Oltx, 50% (6 out of 12) in Group and<br />

62.5% (5 out of 8) in Group 1 were alive. Mean survival was<br />

584 days (SD 106) for Group 1, 294 (SD 49) for Group 3, 20<br />

(SD 3.4) for Group 2. Bootstrapping showed a difference in<br />

survival between Group 1 and Group 3 (Wilcoxon, Log-Rank,<br />

Likelihood Ratio p-Value 6 months,<br />

n=75); and G3, past HCV infection (ie, positive anti-HCV and<br />

negative HCV RNA PCR, n=15). RESULTS. In G1without HCV<br />

infection, 20 of 20 (100%) urine specimens tested negative for<br />

HCV-Ags using our EIA assay. Similarly, in G3 with past or<br />

resolved HCV infection, all 15 of 15 (100%) urine specimens<br />

tested negative for HCV-Ags, confirming 100% specificity of<br />

our HCV-Ags EIA when used for testing urine samples. Patients<br />

in G2 had chronic HCV genotype 1-6 infections, serum HCV<br />

RNA levels ranged from 62,000 to 9,960,000 IU/mL quantified<br />

using PCR, and serum creatinine level ranging from 0.5<br />

to 1.7 mg/dL. In 75 of 75 of G2 patients, all urine specimens<br />

tested positive for HCV-Ags, indicating our HCV-Ags assay had<br />

a sensitivity of 100% for testing urine specimens. The mean<br />

coefficient of variation (CVs) was 9.8% and 6.4%, respectively,<br />

for intra-assay precision and inter-assay reproducibility.<br />

Using a group of urine specimens from patients with very low<br />

serum HCV RNA PCR ranging from 26 to 148 IU/mL, the<br />

lowest detection limit of our HCV-Ags EIA for urine specimens<br />

was equivalent to serum HCV RNA level of 26 IU/mL. Optical<br />

density measurements for HCV-Ags from the urine specimens<br />

with chronic HCV infection correlated significantly with the<br />

corresponding serum HCV RNA level quantified using PCR (y<br />

= 0.0702 x + 0.141, R 2 = 0.8212, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1079A<br />

with positive antibody tests who did not have RNA testing and<br />

(c) 38,061 additional cases if the unscreened population in<br />

VA care was screened. Results for each step of the Cascade of<br />

HCV Care for 2013 and 2014 appear in the Table. Conclusion<br />

VA has diagnosed the majority of patients with HCV and linked<br />

them to HCV care. While antiviral treatment and SVR rates<br />

have improved in 2014 compared to 2013, the largest gap in<br />

care remains the initiation of antiviral treatment.<br />

VA Cascade of HCV Care<br />

HCV infection who were referred to care attended their first<br />

appointment and this was similar among non-Hispanic blacks<br />

and whites. Our findings suggest birth cohort testing is likely<br />

to identify a substantial number of persons living with HCV<br />

infection, although barriers persist early in the care continuum,<br />

particularly for non-Hispanic blacks in whom HCV infection<br />

was most prevalent. Additional efforts are needed to ensure<br />

confirmatory testing and linkage to care for all persons with an<br />

anti-HCV+ test.<br />

Disclosures:<br />

The following authors have nothing to disclose: Claudia Vellozzi, Rajiv Patel, Ben<br />

Schoenbachler, Susan Hariri<br />

Disclosures:<br />

The following authors have nothing to disclose: Lisa I. Backus, Pamela S. Belperio,<br />

Timothy P. Loomis, Larry A. Mole<br />

1787<br />

Hepatitis C Birth-Cohort Testing and Linkage to Care,<br />

Selected U.S. Sites, 2012-2014<br />

Claudia Vellozzi 1 , Rajiv Patel 2 , Ben Schoenbachler 2 , Susan<br />

Hariri 1 ; 1 CDC, Atlanta, GA; 2 ORISE, Oak Ridge, TN<br />

Objective To describe hepatitis C virus (HCV) infection prevalence<br />

and examine the provision of HCV testing and care by<br />

race/ethnicity among persons born during 1945-1965 seen<br />

at selected U.S. clinic settings primarily serving under/uninsured<br />

persons. Methods Data were drawn from CDC’s Hepatitis<br />

Testing and Linkage to Care project conducted at 105<br />

settings in 10 states from 2012-2014 to assess implementation<br />

of one-time HCV testing recommendation for persons born<br />

during 1945-1965. We examined anti-HCV positivity overall<br />

and by age, race/ethnicity, sex, and health insurance. Among<br />

persons who were anti-HCV+, testing and care was evaluated<br />

using the following indicators: HCV-RNA testing, chronic infection<br />

as determined by HCV-RNA positivity, referral to medical<br />

care, and attending first medical appointment. We further<br />

examined differences in HCV testing and care among non-Hispanic<br />

blacks compared to non-Hispanic whites. Results Among<br />

24,966 persons tested, 11.6% (2,900) were anti-HCV+;<br />

positivity was 17.1% in males (2,073/12,827), 13.9% in<br />

non-Hispanic blacks (1,701/12,202), and 16.0% in persons<br />

with public health insurance (1,868/11,650). Among all anti-<br />

HCV+ persons, 73% (2,108) received an HCV RNA test, 71%<br />

(1,497) were currently infected, 82% (1,223) were referred<br />

to care, and 77% (938) attended a first medical appointment.<br />

Among anti-HCV+ persons, the proportion receiving an RNA<br />

test was significantly higher among non-Hispanic whites compared<br />

to non-Hispanic blacks (87.6% vs. 63.2%; p < 0.0001).<br />

Conversely, a significantly larger proportion of non-Hispanic<br />

blacks were chronically infected compared to non-Hispanic<br />

whites (76.9% vs. 61.2%; p< 0.0001). There was no difference<br />

between non-Hispanic whites vs non-Hispanic blacks<br />

in attending a first medical appointment (75.0% vs. 79.2%;<br />

p=0.1954). Conclusion We found high (>11%) anti-HCV positivity<br />

among a population drawn from settings primarily serving<br />

under/uninsured persons. Overall, nearly 30% of individuals<br />

in this population did not receive an HCV-RNA test, and testing<br />

was disproportionately low in non-Hispanic blacks compared<br />

to non-Hispanic whites. Nearly 80% of all persons with current<br />

1788<br />

Treatment With Statins Reduces Liver Cancer Risk In<br />

Patients With Chronic Hepatitis C<br />

Anders H. Nyberg 1 , Ekaterina Sadikova 1 , Jiaxiao Shi 1 , T. Craig<br />

Cheetham 1 , Kevin Chiang 1 , Zobair M. Younossi 2 , Lisa M.<br />

Nyberg 1 ; 1 Kaiser Permanente, San Diego, CA; 2 Department of<br />

Medicine, Inova Faifax Medical Campus, Fairfax, VA<br />

Background: Statin use has been associated with a decreased<br />

risk of developing hepatocellular carcinoma (HCC) in patients<br />

with cardiovascular risk factors (1-2). An analysis of Taiwanese<br />

registry data by Tsan et al showed a decreased risk of<br />

developing HCC in patients with hepatitis C who were taking<br />

statins. However, there is a paucity of data in patients with<br />

liver disease. The aim of this study was to evaluate the effect<br />

of statin use on rates of HCC in our population of hepatitis C<br />

patients at Kaiser Permanente Southern California (KPSC), a<br />

community based health care system. Methods: A sample of<br />

35,712 patients >18 years of age with a chronic HCV diagnosis<br />

by ICD -9 code or positive HCV RNA test between 2002<br />

and 2012 was identified in the KPSC research database. The<br />

KPSC National Cancer Institute Surveillance, Epidemiology,<br />

and End Results (KPSC-NCI SEER) affiliated cancer registry was<br />

used to identify new HCC cases between 2008 and 2012.<br />

Statin use was defined as having 2 or more filled prescriptions<br />

prior to the outcome or the end of the follow-up. Clinical<br />

and sociodemographic patient characteristics were measured<br />

prior and during the study period. Crude and adjusted odds<br />

ratios of liver cancer associated with statin use were assessed<br />

using logistic regression modeling. Results: Statin use and liver<br />

cancer were independently associated with older age, race,<br />

higher prevalence of diabetes, liver cirrhosis, smoking, BMI<br />

and more total healthcare utilization. Crude and adjusted liver<br />

cancer logistic regression models are summarized below. Statin<br />

use prevailed as a strong protective effect after adjustment<br />

of clinical and sociodemographic characteristics.<br />

Disclosures:<br />

T. Craig Cheetham - Grant/Research Support: Gilead, BMS<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

Lisa M. Nyberg - Grant/Research Support: Merck, Gilead, Abbvie


1080A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Anders H. Nyberg, Ekaterina<br />

Sadikova, Jiaxiao Shi, Kevin Chiang<br />

1789<br />

Cause of death in HCV patients who achieved sustained<br />

virologic response: Chronic Hepatitis Cohort Study<br />

(CHeCS), 2006-2012<br />

Fujie Xu 1 , Jian Xing 1 , Anne C. Moorman 1 , Loralee B. Rupp 2 , Stuart<br />

C. Gordon 2 , Mei Lu 2 , Eyasu H. Teshale 1 , Philip R. Spradling 1 ,<br />

Joseph A. Boscarino 3 , Connie M. Trinacty 4 , Mark A. Schmidt 5 ,<br />

Scott D. Holmberg 1 ; 1 Division of Viral Hepatitis, CDC, Atlanta,<br />

GA; 2 Henry Ford Health System, Detroit, MI; 3 Geisinger Health<br />

System, Danville, PA; 4 Kaiser Permanente Hawaii, Honolulu, HI;<br />

5 Kaiser Permanente Northwest, Portland, OR<br />

Background: Successful treatment of chronic HCV infection is<br />

associated with reduced morbidity and mortality. The objective<br />

of this study was to describe the stage of liver disease at the<br />

time of sustained virologic response (SVR) and the timing and<br />

cause of death in post-SVR patients. Methods: We analyzed<br />

data from the Chronic Hepatitis Cohort Study (CHeCS), an<br />

ongoing ‘dynamic’ observational study of patients with chronic<br />

HCV conducted at four integrated health care systems located<br />

in the United States (Henry Ford Health System, Geisinger<br />

Clinic, Kaiser Permanente-Northwest, and Kaiser Permanente-Hawaii).<br />

The stage of liver disease was approximated by<br />

FIB-4 score within 3 months of the SVR date (at 12 weeks post<br />

treatment). Deaths that occurred in 2006-2012 were ascertained<br />

by matching cohort patient records to the most recent<br />

National Death Index, Social Security Death Index, or electronic<br />

state death registries. We examined the causes of death<br />

listed on death certificates in patients whose HCV treatment<br />

had resulted in SVR. Results: Of 15,158 HCV patients enrolled<br />

in CHeCS, HCV treatment resulted in SVR in 1,492 (9.8%)<br />

patients between 2000 and 2012. Of these 1,492 patients,<br />

785 (52.6%) had one or more FIB-4 scores measured within<br />

3 months of the SVR date: the median FIB-4 score was 1.75,<br />

and 24.8% had a FIB-4 score >3.25. During 2006-2012,<br />

2,314 (15.3%) deaths occurred in HCV cohort patients, and<br />

of these, 71 (3.1%) were in patients known to have achieved<br />

SVR (4.8% (71/1,492) of SVR patients died versus 16.4%<br />

(2243/13,666) of non-SVR patients, p3.25 prior to SVR; the mean time<br />

from SVR date to death was 4.9 years (95% CI, 3.5-6.2),<br />

ranging from 1.4 years to 9.3 years. Conclusion: In a large<br />

HCV cohort in the United States, only 10% had achieved SVR<br />

by the end of 2012. While mortality among patients who had<br />

achieved SVR was lower, premature death and liver cancer<br />

can still occur and may be related to the late stage of liver<br />

disease at the time of treatment and SVR.<br />

Disclosures:<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

The following authors have nothing to disclose: Fujie Xu, Jian Xing, Anne C.<br />

Moorman, Loralee B. Rupp, Mei Lu, Eyasu H. Teshale, Philip R. Spradling, Joseph<br />

A. Boscarino, Connie M. Trinacty, Mark A. Schmidt, Scott D. Holmberg<br />

1790<br />

Development of End Stage Liver Disease, Hepatocellular<br />

Carcinoma and Liver-Related Death According to<br />

Biopsy-Confirmed Ishak Fibrosis Stage in the Hepatitis C<br />

Alaska Cohort Study (AK-HepC)<br />

Dana J. Bruden 1 , Lisa J. Townshend-Bulson 2 , James E. Gove 2 , Julia<br />

N. Plotnik 2 , Chriss E. Homan 2 , Annette Hewitt 2 , Michael Bruce 1 ,<br />

Prabhu P. Gounder 1 , Youssef Barbour 2 , Brenna Simons-Petrusa 2 ,<br />

Brian J. McMahon 2,1 ; 1 Arctic Investigations Program, Division of<br />

Preparedness and Emerging Infections, Centers for Disease Control<br />

and Prevention, Anchorage, AK; 2 Liver Disease and Hepatitis<br />

Program, Alaska Native Tribal Health Consortium, Anchorage, AK<br />

Background Most <strong>studies</strong> on liver biopsy results among hepatitis<br />

C virus (HCV) infected persons are of a cross-sectional study<br />

design. Long-term prospective <strong>studies</strong> of the outcomes associated<br />

with HCV are rare and critical for assessing the potential<br />

impact of HCV treatment on sequelae. We used liver biopsy<br />

as a study start point and looked at development of end stage<br />

liver disease (ESLD), hepatocellular carcinoma (HCC) and liver-related<br />

death (LRD) according to the Ishak fibrosis score.<br />

Methods Since 1994, all American Indian/Alaska Native (AI/<br />

AN) persons positive for anti-HCV have been invited to enroll in<br />

a long-term study of HCV outcomes. By 2012, 1,325 persons<br />

were enrolled, of whom 1,080 were HCV RNA-positive, and<br />

of these, 412 had a liver biopsy performed. In persons with<br />

multiple liver biopsies, the most recent was used. The Ishak<br />

fibrosis score was used to categorize persons into mild/moderate<br />

(Ishak = 0, 1, 2), severe (Ishak = 3, 4) or cirrhosis (Ishak<br />

= 5, 6). The date of liver biopsy was the starting point in a<br />

survival analysis examining time until development of ESLD,<br />

HCC and LRD. We report survival probabilities at 5-years using<br />

Kaplan-Meier estimates. Results Of 412 persons undergoing<br />

liver biopsy, 68% (n = 282) had mild/moderate fibrosis, 21%<br />

(n = 87) had severe fibrosis and 10% (n = 43) had cirrhosis.<br />

The average length of following liver biopsy was 7.7 years<br />

with a total of 3,162 years of follow-up. The average age<br />

at the time of biopsy was 44.3 years and 51% (n = 211)<br />

were female. Within 5 years of biopsy, 4.6% (95% confidence<br />

interval (CI): 2.5, 8.5) of persons with mild/moderate fibrosis<br />

had developed ESLD compared to 16.4% (CI: 9.6, 27.2) and<br />

45.1% (CI: 29.7, 63.9) of persons with severe fibrosis and<br />

cirrhosis, respectively (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1081A<br />

1791<br />

Ability of EMR-based fibrosis scores to identify HCV-infected<br />

patients with cirrhosis<br />

Mohammad Qasim Khan 3 , Vijay Anand 3 , Ammar Hassan 3 , Alya<br />

Assan 3 , Amnon Sonnenberg 2 , Claus J. Fimmel 1 ; 1 Gastroenterology,<br />

Northshore University Health System, Evanston, IL; 2 Gastroenterology,<br />

Oregon Health Sciences University, Portland, OR;<br />

3 Internal Medicine, NorthShore University Health System, Evanston,<br />

IL<br />

Background: With the advent of fixed-dose treatment regimens<br />

for chronic HCV (CHC) infection and the updated, population-based<br />

screening guidelines, clinicians are faced with the<br />

challenge of assessing the degree of liver fibrosis in large<br />

patient populations. We asked whether it would be possible<br />

to predict cirrhosis in HCV-infected patients using fibrosis algorithms<br />

that are entirely based on simple laboratory and demographic<br />

features extractable from the electronic medical record<br />

system (EMR). Objective: To compare a set of seven EMR-derived<br />

fibrosis scores with liver imaging <strong>studies</strong> in a cohort of<br />

CHC patients. Methods: Patients with CHC (n=869) were identified<br />

by searching the EPIC-derived patient data warehouse. A<br />

total of 566 patients had undergone a liver imaging study, and<br />

had no confounding medical conditions affecting the fibrosis<br />

scores. Cirrhosis was defined as the presence of a nodular<br />

liver or portal collaterals on ultrasound, CT, or MRI imaging.<br />

Demographic and laboratory data were extracted from the<br />

EMR, matched with the date of the imaging <strong>studies</strong>, and used<br />

to calculate the following previously established fibrosis scores:<br />

APRI, Fib4, Fibrosis Index, Forns, GUCI, Lok Index, and Vira-<br />

HepC. Areas under the receiver operating curves (AUROC),<br />

optimum cut-offs, positive and negative predictive values were<br />

calculated for each score. Results: The seven algorithms performed<br />

similarly well in predicting cirrhosis (Table). No single<br />

test was superior as all of their AUROC confidence intervals<br />

overlapped. Sensitivites ranged from 0.65 to 1.00, specificities<br />

from 0.67 to 0.90, positive predictive values from 0.33<br />

to 0.38, and negative predictive values from 0.93 to 1.00.<br />

Summary: EMR-based algorithms performed well in ruling out<br />

radiologic cirrhosis, whereas their ability to predict the presence<br />

of cirrhosis was modest. Conclusions: Simple, EMR-based<br />

scoring systems can be used to rule out advanced cirrhosis in<br />

CHC patients with high reliability. They are less accurate in<br />

identifying patients with advanced cirrhosis, suggesting that<br />

positive scores would need to be confirmed by secondary testing.<br />

Table<br />

1792<br />

Characteristics of Patients Tested for Hepatitis C and<br />

Intervention Costs in the BEST-C Study<br />

Joanne E. Brady 1 , Danielle Liffmann 1 , Anthony K. Yartel 2 , Natalie<br />

B. Kil 3 , Alex D. Federman 3 , Cynthia E. Jordan 4 , Omar I. Massoud 4 ,<br />

David R. Nerenz 5 , Kimberly Ann Brown 5 , Bryce Smith 6 , Claudia<br />

Vellozzi 2 , David B. Rein 1 ; 1 Public Health, NORC at the University<br />

of Chicago, Bethesda, MD; 2 Division of Viral Hepatitis, U.S. Centers<br />

for Disease Control and Prevention, Atlanta, GA; 3 Division<br />

of General Internal Medicine, Mount Sinai School of Medicine,<br />

New York, NY; 4 Department of Medicine, University of Alabama<br />

at Birmingham, Birmingham, AL; 5 Department of Medicine, Henry<br />

Ford Hospital, Detroit, MI; 6 Division of Diabetes Translation, U.S.<br />

Centers for Disease Control and Prevention, Atlanta, GA<br />

Given that 80% of Hepatitis C virus (HCV)-infected Americans<br />

were born during the years 1945-1965, the Centers for Disease<br />

Control and Prevention (CDC) and the U.S Preventive<br />

Services Task Force recommended a one-time HCV antibody<br />

test for adults born in the 1945-1965 birth cohort (BC). CDC’s<br />

Birth-cohort Evaluation to Advance Screening and Testing for<br />

Hepatitis C was designed to assess the impact of testing interventions<br />

on the probability of HCV testing in primary care<br />

(PC) among BC patients as compared to usual care and the<br />

incremental costs per person tested and per case identified<br />

at each site. From December 2012-March 2014, 3 health<br />

systems implemented independent testing interventions using<br />

randomized designs to compare intervention testing and identification<br />

rates to usual care. Site 1 mailed paid lab test orders<br />

and repeated reminders to a randomly selected list of active<br />

patients compared to a second list who received no mailings.<br />

Site 2 created an electronic health record best practice alert<br />

(BPA) implemented or not implemented based on cluster randomized<br />

design. Site 3 directly solicited patients following a<br />

scheduled PC visit and used a cluster randomized crossover<br />

design. Multilevel multivariable regression was used to estimate<br />

the risk ratio for HCV testing; activity-based costing was used<br />

to estimate costs. HCV testing was significantly more common<br />

for all interventions compared to controls; adjusted risk ratio<br />

(aRR) 19.2, (95% CI, 9.7–38.2), 13.2 (95% CI, 3.6–48.6),<br />

and 32.9 (95% CI 19.3–56.1) for sites 1, 2, and 3, respectively.<br />

The BPA intervention had the lowest incremental cost<br />

per person tested ($25 with fixed startup costs, $3 without<br />

startup costs). The incremental cost per new case identified<br />

under usual care ranged from $3,771-$6207 across sites. All<br />

interventions increased HCV testing among the BC compared<br />

to usual care, but also increased the costs. The cost per case<br />

identified excluding startup costs was lowest for the BPA intervention<br />

($1,691), suggesting that integrating BC testing into<br />

usual care is likely to be more cost-effective than instituting an<br />

intervention in addition to usual care, e.g., repeated-mailings<br />

and patient-solicitation.<br />

n=number of patients, AUROC=area under the receiver operating<br />

curve, CI=confidence interval<br />

Disclosures:<br />

The following authors have nothing to disclose: Mohammad Qasim Khan, Vijay<br />

Anand, Ammar Hassan, Alya Assan, Amnon Sonnenberg, Claus J. Fimmel<br />

Costs measured in dollars


1082A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Kimberly Ann Brown - Advisory Committees or Review Panels: CLDF, Merck,<br />

BMS, ABBVIE, Gilead, Gilead, Janssen, Salix; Consulting: Blue Cross Transplant<br />

Centers; Grant/Research Support: CLDF, Gilead, Exalenz, CDC, BMS, Hyperion,<br />

Merck; Speaking and Teaching: ABBVIE, Gilead, CLDF<br />

David B. Rein - Grant/Research Support: Gilead Sciences, Inc.<br />

The following authors have nothing to disclose: Joanne E. Brady, Danielle Liffmann,<br />

Anthony K. Yartel, Natalie B. Kil, Alex D. Federman, Cynthia E. Jordan,<br />

Omar I. Massoud, David R. Nerenz, Bryce Smith, Claudia Vellozzi<br />

1793<br />

Risk of hepatitis C virus infection in people who use<br />

drugs who report male-to-male sex<br />

Chloe Le Marchand 1 , Ali Mirzazadeh 1 , Stephen Shiboski 1 , Thomas<br />

M. Rice 1 , Meghan D. Morris 1 , Judith A. Hahn 1 , Kimberly Page 2,1 ;<br />

1 University of California, San Francisco, Berkeley, CA; 2 University<br />

of New Mexico, Albuquerque, NM<br />

Introduction Most new hepatitis C virus (HCV) infections occur<br />

among people who use drugs (PWID) via shared injecting<br />

equipment. Rising HCV infection rates in non-injecting, HIV-infected<br />

men who have sex with men (MSM) are linked to highrisk<br />

sexual activity. Studies examining the incidence of HCV in<br />

MSM have mixed results; several show an increased susceptibility<br />

to HCV principally in HIV-infected patients. Incidence<br />

rates range from 18.0/1000 person-years (py) in HIV-infected<br />

MSM 9.0/1000 py in MSM without HIV. In this study, we<br />

estimate HCV incidence in male PWID in association with<br />

MSM behavior, controlling for injecting exposures. Methods<br />

We used data from the INC3 collaboration, which pooled biological<br />

and behavioral data from 9 cohorts of PWID in North<br />

America, Europe and Australia, including 1,921 males at risk<br />

for HCV. We included all participants in INC3 with a history of<br />

drug use and a negative anti-HCV test at enrollment. We conducted<br />

contingency table analyses with Chi-square analysis for<br />

descriptive variables by MSM status. Incident infections were<br />

defined as a positive infection (anti-HCV positive or HCV RNA<br />

positive) occurring during follow-up. We used the log-rank test<br />

to compare survival curves by sexual and injecting risk behaviors.<br />

We estimated incidence by MSM status as compared to<br />

non-MSM males and conducted Cox regression analyses to<br />

compute hazard ratios controlling for potential confounders,<br />

including: HIV status, recent injecting, syringe and cooker sharing,<br />

injection frequency, number of injecting partners, access<br />

to clean needles, year entered the study and clustered by site.<br />

Results In 1,064 male participants, followed for 1,921 py,<br />

342 new HCV infections were observed for an overall incidence<br />

of 16.1/100 py (95% CI 14.4, 18.0). No differences<br />

were seen in estimated HCV incidence in MSM participants<br />

(16.3/100 py (95% CI 12.5, 21.2)) compared to non-MSM<br />

males (16.0/100 py (95% CI 14.1, 18.1)). We assessed HCV<br />

incidence in HIV-infected compared to non-infected groups and<br />

found no differences. MSM were different from non-MSM at<br />

baseline with respect to injecting risk: 61% vs. 47% reported<br />

syringe or equipment sharing, respectively (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1083A<br />

1795<br />

Distribution of pre-existing NS5A/NS5B resistance associated<br />

variants in genotype 1b patients with hepatitis C<br />

virus and response to direct acting antivirals<br />

Takeshi Watabe, Masaaki Korenaga, Masaya Sugiyama, Erina<br />

Kumagai, Misuzu Ueyama, Yoshihiko Aoki, Yoko Yamagiwa,<br />

Keiko Korenaga, Masatoshi Imamura, Kazumoto Murata, Naohiko<br />

Masaki, Tatsuya Kanto, Masashi Mizokami; The research center<br />

for hepatitis and Immunology, National Center of Global Health<br />

and Medicine (NCGM) at Kohnodai, Ichikawa, Japan<br />

Background and Aim: Although interferon-free antiviral treatment<br />

is improved sustained virological response (SVR) rate of<br />

hepatitis C virus (HCV), pre-existing NS5A/NS5B resistance<br />

associated variants (RAVs) may affect the efficiency of some<br />

HCV regimens. The aim of study is to investigate whether the<br />

NS5A/NS5B RAVs interact with clinical characteristics and<br />

response to 12 weeks of treatment with the nucleotide polymerase<br />

inhibitor sofosbuvir (SOF) combined with the NS5A<br />

inhibitor ledipasvir (LDV) for genotype (GT)-1b infection. Methods:<br />

Three hundred sixteen Japanese patients with HCV GT-1b<br />

(mean Age [ys]:67, male [%]:32, treatment naïve [%]:53 cirrhosis<br />

[%]:30, and hepatocellular carcinoma (HCC) [%]:10)<br />

were examined NS5A/NS5B RAVs by direct sequence. NS5A<br />

RAVs included L31M/I/V/F and Y93H, and NS5B RAVs<br />

included S282T and C316N. The frequency of the RAVs was<br />

examined the association with clinical characteristics, such as<br />

age, sex, naïve/experienced, HCVRNA, fibrosis, and hepatocellular<br />

carcinoma (HCC). Fibrosis was evaluated by Fibroscan<br />

and Mac-2 binding protein glycogen isomer. Among<br />

these patients, 59 patients (23%) who received LDV (90 mg)/<br />

SOF (400 mg) for 12 weeks were analyzed the association<br />

of pre-treatment NS5A/NS5B RAVs with treatment outcome.<br />

Results: Distribution of NS5A RAVs existed 5% for L31 and<br />

23% for Y93. Double mutations in NS5A were detected only<br />

5 patients (1.5%). Distribution of NS5B RAVs existed 48% for<br />

C316; however there were no detection for S282 mutation. No<br />

association of these RAVs distribution with clinical characteristics.<br />

Fifty nine patients with receiving SOF/LDV were detected<br />

23% for NS5A RAVs, 48% for C316N, and 11% for NS5A<br />

RAVs with C316N and All 59 patients achieved SVR (100%).<br />

Conclusions: Although pre-treatment NS5A and NS5B RAVs in<br />

Japanese patients with infected HCV GT-1b exist about 25%<br />

and 50% respectively, presence of single RAVs, and NS5A<br />

RAVs with C316N in NS5B are less association with clinical<br />

characteristics, and treatment outcome of SOF/LDV.<br />

# All 59 patinets trerated with SOF/LDV for 12 weeks and<br />

achieved SVR<br />

Disclosures:<br />

The following authors have nothing to disclose: Takeshi Watabe, Masaaki<br />

Korenaga, Masaya Sugiyama, Erina Kumagai, Misuzu Ueyama, Yoshihiko Aoki,<br />

Yoko Yamagiwa, Keiko Korenaga, Masatoshi Imamura, Kazumoto Murata, Naohiko<br />

Masaki, Tatsuya Kanto, Masashi Mizokami<br />

1796<br />

Prevalence of viral hepatitis and liver fibrosis in streetbased<br />

outreach: a French prospective multicentre study<br />

Juliette Foucher 13 , Jean Harbonnier 1 , Leon Gomberoff 2 , Sylvie<br />

Balteau 3 , Fadi Meroueh 4 , Beatrice Stambul 5 , Rolland Le Hello 6 ,<br />

Brigitte Reiller 7 , Xavier Richen 8 , Elisabeth Avril 9 , Jean-Pierre<br />

Daulouede 10 , Anne Borgne 11 , Veronique Latour 12 , Jean-Baptiste<br />

Hiriart 13 , Victor de Ledinghen 13 ; 1 CSAPA Boris Vian, Lille, France;<br />

2 CSAPA Ego Aurore, Paris, France; 3 CSAPA Les Wads, Metz,<br />

France; 4 CSAPA Arc en Ciel, Montpellier, France; 5 CSAPA Villa<br />

Floreal, Aix en Provence, France; 6 CSAPA L’envol, Rennes, France;<br />

7 CEID CSAPA Planterose, Bordeaux, France; 8 CSAPA du griffon<br />

Aria, Lyon, France; 9 CSAPA Gaia, Paris, France; 10 CSAPA Bizia<br />

Medecin du Monde, Bayonne, France; 11 service d’addictologie<br />

Hopital Muret, Sevran, France; 12 CSAPA La case Medecins du<br />

monde, Bordeaux, France; 13 hepatology unit, Bordeaux, France<br />

Introduction French Association for the Study of the Liver diseases<br />

guidelines recommends treatment of drug users whatever<br />

liver fibrosis stage. The aim of this prospective multicentre<br />

study was to evaluate viral hepatitis in French street-based outreaches<br />

and their severity. Patients and methods From January<br />

2012 to March 2013, all consecutive subjects with intravenous<br />

drug use (PWID) without any HCV management and consulting<br />

in street-based outreach in France were included. After face-toface<br />

questionnaire by outreach worker, liver stiffness was measured<br />

by FibroScan (FS). At the end of FS, a blood sample for<br />

hepatitis B and C serology was proposed as a meeting with a<br />

practitioner. During face-to-face interview, socio-demographic<br />

parameters and way of living were recorded. During the medical<br />

meeting, the result of serology and FS were explained. If<br />

case of HCV infection, users could meet a hepatologist and<br />

treatment could be proposed. Results In 12 French street-based<br />

outreaches, 860 drug-users were included (663 men, mean<br />

age 37 years, BMI 23 kg/m 2 , injected drug used history<br />

n=447, currently opioid substitution 62%, homeless 16,%,<br />

unemployed 72%, past history of incarceration 43%). 813<br />

subject accepted FS, with correct measurement in 94.5% of<br />

cases. Median liver stiffness was 5.1 kPa (2-75 kPa). 16.4% of<br />

users had significant fibrosis (FS>7.6 kPa) and 4.8% cirrhosis<br />

(FS>13 kPa). 77% of drug users with interpretable FS had a<br />

blood sample. HCV prevalence was 34%. HIV and HBV prevalence<br />

were 2% and 1.2%, respectively. Factors associated with<br />

HCV infection were age, history of IVDU, history of incarceration<br />

and no insurance. In multivariate analysis, factors associated<br />

with HCV were history of IVUD (OR 19.38, [9.9-37.8]<br />

p


1084A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Jean Harbonnier, Leon Gomberoff,<br />

Sylvie Balteau, Fadi Meroueh, Beatrice Stambul, Rolland Le Hello, Brigitte<br />

Reiller, Xavier Richen, Elisabeth Avril, Jean-Pierre Daulouede, Anne Borgne,<br />

Veronique Latour, Jean-Baptiste Hiriart<br />

1797<br />

Hepatitis C RNA assay differences in results around 6<br />

million IU/mL: Potential clinical implications for shortened<br />

Ledipasvir/Sofosbuvir therapy<br />

Gavin A. Cloherty 1 , Christoph Sarrazin 2 , Vera Holzmayer 3 , Jordan<br />

J. Feld 4 , Benjamin Maasoumy 5 , Johannes Vermehren 2 , Stephane<br />

Chevaliez 6 , Heiner Wedemeyer 5 , Jean-Michel Pawlotsky 6 , George<br />

Dawson 3 ; 1 R&D, Abbott Molecular, Des Plaines, IL; 2 Goeth Universtiy,<br />

Frankfurt, Germany; 3 R&D, Abbott Diagnostics, Abbott Park,<br />

IL; 4 Universtiy of Toronto, Toronto, ON, Canada; 5 Hannover Medical<br />

School, Hannover, Germany; 6 Hopital Henri Mondor, Paris,<br />

France<br />

Background/Aims:FDA approval of Ledipasvir/Sofosbuvir for<br />

the treatment of hepatitis C (HCV) includes the truncation of<br />

therapy from 12 to 8 weeks in treatment naïve, non-cirrhotic<br />

patients with HCV RNA levels 6 million and ART 3 on liver biopsy<br />

or APRI >1.5), ELSD (presence of ascites, esophageal varices,<br />

hepatic encephalopathy, or coagulopathy), HCC, and LRD<br />

(including liver transplantation) by using a Cox proportional<br />

hazards model. The model for advanced fibrosis and ESLD was<br />

adjusted for HCV genotype (1, 2, and 3), age at cohort entry,<br />

heavy alcohol use (average >50grams/day), obesity (body<br />

mass index [BMI] >30), and type II diabetes (DM2). The model<br />

for HCC risk was adjusted for HCV genotype, age at cohort<br />

entry, and sex. The model for LRD risk was adjusted for HCV<br />

genotype, age at cohort entry, and heavy alcohol use. Of the<br />

1068 participants, 66%, 19%, and 14% were infected with<br />

HCV genotype 1, 2, and 3, respectively. Compared with genotype<br />

1, genotype 3 was significantly associated with advanced<br />

fibrosis, ESLD, HCC and LRD (Table). Other age-adjusted significant<br />

risk factors for developing advanced fibrosis were heavy<br />

alcohol use (aHR: 2.0; CI: 1.5–2.7), obesity (aHR: 1.3; CI:<br />

1.0–1.7; p = 0.04), and DM2 (aHR: 1.6; CI: 1.1–2.2). Significant<br />

risk factors for developing ESLD were heavy alcohol use<br />

(aHR: 2.2; CI: 1.6–3.2), obesity (aHR: 1.4; CI: 1.0–1.9; p =<br />

0.03), and DM2 (aHR: 1.5; CI: 1.0–2.3; p = 0.03). Another<br />

significant risk factor for developing HCC was male sex (aHR:<br />

3.6; CI: 1.6–8.2) and for developing LRD was heavy alcohol<br />

use (aHR: 2.9; CI: 1.8–4.6). At 15 years of follow-up, twothirds<br />

(68%) with genotype 3 were predicted to have developed<br />

advanced fibrosis and half (48%) ESLD. Conclusions:


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1085A<br />

Because AI/AN persons infected with HCV genotype 3 are at<br />

high risk for developing advanced fibrosis, ESLD, HCC, and<br />

LRD, they should be prioritized for earlier antiviral treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Anne C. Moorman, Yueren Zhou,<br />

Loralee B. Rupp, Joseph A. Boscarino, Connie M. Trinacty, Scott D. Holmberg<br />

Disclosures:<br />

Lisa J. Townshend-Bulson - Grant/Research Support: Gilead<br />

James E. Gove - Grant/Research Support: Gilead Sciences<br />

Annette Hewitt - Grant/Research Support: gilead<br />

Julia N. Plotnik - Grant/Research Support: Gilead<br />

Chriss E. Homan - Grant/Research Support: Gilead<br />

Hannah G. Espera - Grant/Research Support: Gilead<br />

Youssef Barbour - Grant/Research Support: Gilead<br />

The following authors have nothing to disclose: Brian J. McMahon, Dana J.<br />

Bruden, Stephen Livingston, Susan Negus, Mary Snowball, Michael Bruce, Philip<br />

R. Spradling, Prabhu P. Gounder<br />

1799<br />

The Utility and Limitations of Electronic Medical Records:<br />

Predictive Values of ICD9 Codes indicating Chronic HCV<br />

Infection<br />

Anne C. Moorman 1 , Yueren Zhou 2 , Loralee B. Rupp 2 , Joseph A.<br />

Boscarino 3 , Connie M. Trinacty 4 , Scott D. Holmberg 1 ; 1 Division of<br />

Viral Hepatitis, C DC, Atlanta, GA; 2 Henry Ford Health Systems,<br />

Detroit, MI; 3 Geaisinger Health System, Danville, PA; 4 Kaiser Permanente-Hawaii,<br />

Honolulu, HI<br />

Objective. With increasing analysis of electronic medical<br />

records (EMR) in the United States, the predictive values of<br />

the International Classification of Diseases, 9th revision (ICD-<br />

9) coding system for chronic hepatitis C virus (HCV) infection<br />

warrant evaluation. Methods. Patients in the Chronic Hepatitis<br />

Cohort Study (CCHeCS) with chronic HCV infection, confirmed<br />

by laboratory data and abstractor review (a ‘gold standard’),<br />

were compared to “chronically infected cases” identified by<br />

electronic medical records (EMRs) ICD9 codes of adult patients<br />

who accessed services from 2006 to 2012 at four integrated<br />

healthcare systems in the United States. Results. Of over 1.6<br />

million adult patients at the 4 large medical systems in CHeCS,<br />

11,354 were confirmed as having chronic HCV infection.<br />

Having an EMR with one ICD-9 code for HCV yielded good<br />

sensitivity and positive predictive value (PPV) for true HCV<br />

infection, but poor specificity (Table), ie.e many (n=724) who<br />

were listed as having HCV infection but were not infected.<br />

Use of two hepatitis C-specific ICD-9 codes (6 or more months<br />

apart) improved the specificity with some loss of sensitivity,<br />

and the negative predictive value (NPV, non-cases accurately<br />

identified as non-cases) was poor (Table) . Because test codes<br />

were not standardized between laboratories, and viral load<br />

results were in text fields, distinguishing true from questionable<br />

or non-cases of HCV infection required extensive visual inspection<br />

and data processing. Discussion. EMRs, especially ICD9<br />

(or ICD10) codes, are increasingly used and analyzed in the<br />

United States, but these have imperfect sensitivity, specificity,<br />

and positive and negative predictive values for hepatitis C, and<br />

require considerable human review and judgment.<br />

1800<br />

Liver fibrosis stage and comorbidities in persons with<br />

diagnosed hepatitis C infection, Chronic Hepatitis Cohort<br />

Study (CHeCS)<br />

Fujie Xu 1 , Jian Xing 1 , Anne C. Moorman 1 , Stuart C. Gordon 2 ,<br />

Loralee B. Rupp 2 , Mei Lu 2 , Philip R. Spradling 1 , Eyasu H. Teshale 1 ,<br />

Joseph A. Boscarino 3 , Mark A. Schmidt 4 , Connie M. Trinacty 5 ,<br />

Scott D. Holmberg 1 ; 1 Division of Viral Hepatitis, Centers for Disease<br />

Control and Prevention, Atlanta, GA; 2 Henry Ford Health System,<br />

Detroit, MI; 3 Geisinger Health System, Danville, PA; 4 Kaiser<br />

Permanente- Northwest, Portland, OR; 5 Kaiser Permanente-Hawaii,<br />

Honolulu, HI<br />

Background and Aims: The Chronic Hepatitis Cohort Study<br />

(CHeCS), a dynamic, prospective, observational cohort study,<br />

was created to study the natural history of chronic viral hepatitis<br />

in the United States. This report described HCV cohort patients’<br />

demographic and clinical characteristics, comorbidities, treatment<br />

and treatment outcomes of cohort patients at the end of<br />

2012. Methods: CHeCS patients were drawn from about 2.7<br />

million persons aged 18 years or older who had a clinical service<br />

visit from January 2006-December 2012 at 4 integrated<br />

healthcare systems in Detroit, Michigan; Danville, Pennsylvania;<br />

Portland, Oregon; and Honolulu, Hawaii. We analyzed<br />

clinical data obtained from electronic medical records and<br />

manual chart review for HCV-infected patients who were alive<br />

but never had a liver transplant as of December 2012, including<br />

patients who achieved SVR with altered natural history.<br />

Patient’s current fibrosis stage was based on latest FIB4 scores<br />

derived from blood tests or biopsy results, if available. Laboratory<br />

tests and ICD-9 codes were used to identify patients with<br />

co-infections and comorbidities. Results: Of 15,940 total HCV<br />

cohort patients 2006-12, 11,294 were alive of Dec 2012;<br />

among these patients mean age at the end of observation was<br />

54 years; 58% were male, 55% white and 22% black. Only<br />

1,355 (9%) of the cohort had achieved SVR, which occurred<br />

a median of 4.8 years previously. Of 9,975 patients without<br />

SVR, 2,553 (25.6%) had ever been treated during a median<br />

followup of 5.3 years. Viral load data was available for 7,189<br />

patients; of these, 14% had 6 million IU/ml or higher (most<br />

recent test among those untreated and the latest before initial<br />

treatment among those treated). Impaired kidney function with<br />

eGFR


1086A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

The following authors have nothing to disclose: Fujie Xu, Jian Xing, Anne C.<br />

Moorman, Loralee B. Rupp, Mei Lu, Philip R. Spradling, Eyasu H. Teshale, Joseph<br />

A. Boscarino, Mark A. Schmidt, Connie M. Trinacty, Scott D. Holmberg<br />

1801<br />

From Care to a Cure: Improving the Hepatitis C Care<br />

Cascade through Patient Navigation in the Check Hep C<br />

Program in New York City<br />

Mary Ford, Nirah Johnson, Payal Desai, Eric J. Rude, Fabienne<br />

Laraque; New York City Department of Health and Mental<br />

Hygiene, New York, NY<br />

Background: Of the estimated 146,500 persons with hepatitis<br />

C (HCV) infection in New York City (NYC), only about 10% are<br />

thought to have ever received treatment. New, highly effective<br />

HCV treatments provide an impetus for increased public health<br />

efforts to support HCV-infected persons along the care cascade<br />

from diagnosis to cure. Check Hep C, a program of the NYC<br />

Department of Health and Mental Hygiene (DOHMH), provided<br />

HCV screening, navigation, and tele-medicine in Year 1.<br />

Over 4,000 persons were tested, and 512 were RNA positive.<br />

Of those, 435 (85%) were linked to care, 14 initiated treatment<br />

and 3 had a sustained virological response (SVR). The small<br />

number of patients treated and cured highlighted the need<br />

for intensified patient navigation services through treatment.<br />

Methods: In Year 2, navigation services were provided at four<br />

sites following a standardized protocol. Services included a<br />

comprehensive social and health assessment, appointment<br />

assistance, support for medical evaluation and treatment, and<br />

pharmacy coordination. Patient navigators worked to enroll a<br />

target of 400 participants and form strong relationships with<br />

the multidisciplinary team involved in each patient’s care. Program<br />

data were entered in a standard database by the patient<br />

navigators and sent to DOHMH. Descriptive analysis was conducted.<br />

Results: Between March 2014 and January 2015, 388<br />

(97% of the target) participants were enrolled in Check Hep C.<br />

The median age of participants was 52 years, and 236 (61%)<br />

were born between 1945-1965. Of 376 participants with<br />

race/ethnicity data, 242 (64%) were Hispanic and 103 (27%)<br />

were Black, non-Hispanic. Of 364 participants with risk history<br />

data, 182 (50%) had current chemical dependence, 108<br />

(30%) had a history of intravenous drug use, and 91 (25%)<br />

were homeless. As of April 2015, 308 (79%) participants<br />

received HCV medical care, and 301 (78%) completed an<br />

HCV medical evaluation, of which 232 (77%) were treatment<br />

candidates. Of the treatment candidates, 116 (50%) initiated<br />

treatment, and 86 (74%) of those completed treatment. Thus<br />

far, 70 participants (81% of those who completed treatment)<br />

achieved SVR. Conclusion: The Check Hep C program successfully<br />

assisted high-need participants through HCV care and<br />

treatment, including those with active chemical dependence<br />

and homelessness, helping a much higher number of patients<br />

in achieving SVR in Year 2. Programs such as this community-based<br />

patient navigation intervention can help improve the<br />

HCV care cascade, particularly for high-need persons. Sustainable<br />

sources of funding, including insurance reimbursement,<br />

need to be provided to support these critical services.<br />

Disclosures:<br />

Eric J. Rude - Grant/Research Support: Vertex, Merck, Bristol Myers Squibb,<br />

Orasure, Janssen, Gilead, Kadmon, Boehringer-Ingelheim, Abbott, Genentech<br />

The following authors have nothing to disclose: Mary Ford, Nirah Johnson, Payal<br />

Desai, Fabienne Laraque<br />

1802<br />

Treatment prioritization according to the EASL HCV CPG<br />

2015: a real-life evaluation on the PITER (Piattaforma<br />

Italiana per lo studio della Terapia delle Epatiti viRali)<br />

cohort<br />

Loreta A. Kondili 1 , Stefano Rosato 2 , Maria Giovanna Quaranta 1 ,<br />

Liliana E. Weimer 1 , Loredana Falzano 1 , Alessandra Mallano 1 ,<br />

Maria Elena Tosti 2 , Maurizio Massella 1 , Maurizia R. Brunetto 3 ,<br />

Anna Linda Zignego 4 , Mario Rizzetto 5 , Alfredo Di Leo 6 , Giovanni<br />

Raimondo 7 , Carlo Ferrari 8 , Gloria Taliani 9 , Pierluigi Blanc 28 , Antonio<br />

Gasbarrini 10 , Luchino Chessa 11 , Elke M. Erne 12 , Giovanna Fattovich<br />

13 , Pietro Andreone 14 , Maria Vinci 15 , Francesco P. Russo 16 ,<br />

Erica Villa 17 , Giovanni B. Gaeta 18 , Teresa A. Santantonio 19 , Guglielmo<br />

Borgia 20 , Gabriella Verucchi 21 , Carmine Coppola 22 , Marcello<br />

Persico 23 , Liliana Chemello 29 , Alfredo Alberti 16 , Vincenzo De<br />

Maria 30 , Massimo Puoti 31 , Raffaele Bruno 24 , Paolo Caraceni 14 ,<br />

Massimo Andreoni 25 , Marco Marzioni 26 , Stefano Vella 1 , Antonio<br />

Craxi 27 ; 1 Therapeutic Research and Medicines Evaluation,<br />

Istituto Superiore di Sanità, Rome, Italy; 2 Istituto Superiore di Sanità,<br />

Rome, Italy; 3 Azienda Ospedaliero-Universitaria Pisana, Pisa,<br />

Italy; 4 University of Florence, Florence, Italy; 5 University of Torino,<br />

Torino, Italy; 6 University of Bari, Bari, Italy; 7 University Hospital of<br />

Messina, Messina, Italy; 8 Azienda Ospedaliero- Universitaria di<br />

Parma, Parma, Italy; 9 Sapienza University of Rome, Rome, Italy;<br />

10 Policlinico Universitario Agostino Gemelli, Rome, Italy; 11 University<br />

of Cagliari, Cagliari, Italy; 12 Azienda Ospedaliera Padova,<br />

Padova, Italy; 13 University of Verona, Verona, Italy; 14 University<br />

of Bologna, Bologna, Italy; 15 Niguarda Hospital, Milan, Italy;<br />

16 Azienda Ospedaliero-Universitario Padova, Padova, Italy; 17 University<br />

of Modena and Reggio Emilia, Modena, Italy; 18 Second<br />

University of Naples, Naples, Italy; 19 University of Foggia, Foggia,<br />

Italy; 20 Federico II University of Naples, Naples, Italy; 21 Sant’Orsola<br />

Malpighi Hospital of Bologna, Bolgna, Italy; 22 Gragnano<br />

Hospital, Naples, Italy; 23 University of Salerno, Salerno, Italy;<br />

24 University of Pavia, Pavia, Italy; 25 Tor Vergata University,<br />

Roma, Italy; 26 Università Politecnica delle Marche, Ancona, Italy;<br />

27 University of Palermo, Palermo, Italy; 28 Santa Maria Annunziata<br />

Hospital, Florence, Italy; 29 University of Padua, Padua, Italy;<br />

30 Azienda Ospedaliero-Universitaria Mater Domini, Catanzaro,<br />

Italy; 31 Azienda Ospedaliera Niguarda-Cà Granda, Milano, Italy<br />

Aim. The high cost of DAAs for chronic hepatitis C has generated<br />

allocation policies mostly based on fibrosis staging as<br />

a surrogate for treatment needs. The EASL HCV CPG 2015<br />

indicates treatment allocation: treatment prioritized as first line,<br />

treatment justified as second line then deferred and not recommended<br />

treatment. We assessed the impact of this approach<br />

on a real-life cohort of 6831 consecutive patients presenting<br />

for care at 80 Italian Units, collected over the last 12 months<br />

within the PITER framework. Methods. The EASL CPG treatment<br />

prioritization algorithm considers fibrosis stage, HIV or HBV<br />

coinfection, pre and post transplant setting, extra-hepatic manifestations,<br />

i.v. drug use, hemodialysis. The independent effect<br />

of each of these factors and of major comorbidities (cardiovascular,<br />

metabolic, autoimmune/reumatological, neurological/psychiatric,<br />

neoplastic severe diseases) in patients older<br />

than 75 years, was evaluated by multivariate analysis. Results.<br />

Median age of the 6831 patients was 58 years (range 20-95<br />

yrs); 3797 (56%) were male. Matching of patients with the<br />

EASL CPG treatment indications resulted being categorized as<br />

follow: prioritized: 739 (11%) patients with F3 fibrosis; 1846<br />

(27%) patients with F4 fibrosis/Child A cirrhosis; 723 (11%)<br />

patients with Child B/C cirrhosis of whom 340 (5%) patients<br />

with HCC; 173 (3%) women of childbearing age; 280 (0.4%)<br />

drug users; 397 (6%) HIV or HBV coinfected patients; 587<br />

(9%) patients with severe extra-hepatic manifestations; 52 (1%)<br />

patients in LT waiting list; 84 (1%) patients with post LT HCV


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1087A<br />

recurrence. The following were categorized as justified: 440<br />

(6.4%) patients with F2 fibrosis; 661(10%) patients with F0-F1<br />

fibrosis fall in the deferrable category. Hence, in this large<br />

cohort, treatment should overall be prioritized in 3417 (50%)<br />

patients (HCV Gt 1: 66%; Gt 2: 10%; Gt 3: 12%; Gt 4: 7%),<br />

can be justified in 440 (6.4%) and deferred in 661 (9.6%)<br />

patients, while it is not indicated in 2313 (33.8%) patients due<br />

to severe co-morbidities or age older than 75 years. By comparison,<br />

the prioritization algorithm endorsed by AIFA (the Italian<br />

Medicines Agency), mostly based on fibrosis alone, which<br />

dictates reimbursement of DAA in Italy, would allow treatment<br />

of only 2827 (41%) of these patient cohort. Conclusion: Among<br />

patients with chronic hepatitis C, allocation of DAA according<br />

to priority rules linked to a perceived need for more prompt<br />

treatment will allow access to therapy to 40-55% of patients.<br />

There is a clear need for a validated predictive model to assess<br />

the impact of delaying treatment or not treating the remaining<br />

patients.<br />

Disclosures:<br />

Maurizia R. Brunetto - Speaking and Teaching: Roche, Gilead, Schering-Plough,<br />

Bristol-Myers Squibb, Abbott, Roche, Gilead, MSD, Novartis<br />

Mario Rizzetto - Advisory Committees or Review Panels: Merck, Janssen, BMS<br />

Alfredo Di Leo - Speaking and Teaching: Abbvie, MSD, Gilead, Roche, Italfarmaco,<br />

THD, CMD pharmalimited<br />

Giovanni Raimondo - Speaking and Teaching: BMS, Gilead, Roche, Merck,<br />

Janssen, Bayer, MSD<br />

Carlo Ferrari - Advisory Committees or Review Panels: Gilead, Roche, Abbvie,<br />

BMS, Merck, Arrowhead; Grant/Research Support: Gilead, Roche, Janssen<br />

Gloria Taliani - Speaking and Teaching: ROCHE, Merck, BMS, Gilead, Jannsen,<br />

Novartis, AbbVie<br />

Luchino Chessa - Board Membership: Abbvie; Speaking and Teaching: BMS,<br />

Jannsen<br />

Pietro Andreone - Advisory Committees or Review Panels: Janssen-Cilag, Gilead,<br />

MSD/Schering-Plough, Abbvie; Speaking and Teaching: Gilead, BMS<br />

Erica Villa - Advisory Committees or Review Panels: MSD, Abbvie, GSK, Gilead;<br />

Speaking and Teaching: Novartis<br />

Giovanni B. Gaeta - Advisory Committees or Review Panels: Janssen, Merck,<br />

Abbvie, Roche; Speaking and Teaching: BMS, Gilead<br />

Alfredo Alberti - Advisory Committees or Review Panels: Merck, roche, Gilead,<br />

Merck, roche, Gilead, Merck, roche, Gilead, Merck, roche, Gilead; Grant/<br />

Research Support: Merck, gilead, Merck, gilead, Merck, gilead, Merck, gilead;<br />

Speaking and Teaching: novartis, BMS, novartis, BMS, novartis, BMS, novartis,<br />

BMS<br />

Massimo Puoti - Advisory Committees or Review Panels: GSK, Abbott, Janssen,<br />

MSD, Roche, Gilead Sciences, Novartis, GSK, Abbott, Janssen, MSD, Roche,<br />

Gilead Sciences, Novartis, GSK, Abbott, Janssen, MSD, Roche, Gilead Sciences,<br />

Novartis, GSK, Abbott, Janssen, MSD, Roche, Gilead Sciences, Novartis; Speaking<br />

and Teaching: BMS, BMS, BMS, BMS<br />

Paolo Caraceni - Advisory Committees or Review Panels: GSK; Consulting: BMS;<br />

Grant/Research Support: Grifols; Speaking and Teaching: Baxter, Kedrion<br />

The following authors have nothing to disclose: Loreta A. Kondili, Stefano Rosato,<br />

Maria Giovanna Quaranta, Liliana E. Weimer, Loredana Falzano, Alessandra<br />

Mallano, Maria Elena Tosti, Maurizio Massella, Anna Linda Zignego, Pierluigi<br />

Blanc, Antonio Gasbarrini, Elke M. Erne, Giovanna Fattovich, Maria Vinci, Francesco<br />

P. Russo, Teresa A. Santantonio, Guglielmo Borgia, Gabriella Verucchi,<br />

Carmine Coppola, Marcello Persico, Liliana Chemello, Vincenzo De Maria, Raffaele<br />

Bruno, Massimo Andreoni, Marco Marzioni, Stefano Vella, Antonio Craxi<br />

1803<br />

Hepatitis E in haematological and rheumatological<br />

patients, a multicentre study<br />

Johann von Felden 1 , Laurent Alric 2 , Vincent Mallet 11 , Ansgar W.<br />

Lohse 1 , Paul Schnitzler 3 , Heiner Wedemeyer 4 , Christoph Hoener<br />

zu Siederdissen 4 , Maria Teresa Giordani 5 , Celia Aitken 6 , Jan<br />

Cornelissen 10 , Dominik Bettinger 7 , Robert Thimme 7 , Jean-Marie<br />

Peron 8 , Sven Pischke 1 , Robert A. de Man 9 ; 1 Gastroenterology,<br />

Hepatology, and Infectious Diseases, University Medical Centre<br />

Hamburg, Hamburg, Germany; 2 Unité de recherche clinique sur<br />

les hépatites virales, Médecine Interne-Pôle Digestif, CHU Purpan,<br />

Toulouse, France; 3 Virology, University Hospital, Heidelberg, Germany;<br />

4 Gastroenterology, Hepatology, Endocrinology, Hannover<br />

Medical School, Hannover, Germany; 5 Infectious and Tropical<br />

Diseases Unit, San Bartolo Hosptial, Vincenza, Italy; 6 Virology,<br />

NHS Greater Glasgow and Clyde, Glasgow, United Kingdom;<br />

7 Dept. Internal Medicine II, University Medical Centre Freiburg,<br />

Freiburg, Germany; 8 Gastro-entérologie et hépatologie, Médicine<br />

Interne-Pôle Digestif, Toulouse, France; 9 Hepatology, Gastroenterology,<br />

and Infectious Diseases, Erasmus Medical Center,<br />

Rotterdam, Netherlands; 10 Dept. Haematology, Erasmus Medical<br />

Center, Rotterdam, Netherlands; 11 Unité d’hépatologie, Université<br />

Paris Descartes, Paris, France<br />

Background Acute and chronic hepatitis E (HEV) in solid<br />

organ transplant recipients has been frequently described<br />

but the knowledge about HEV in patients with underlying<br />

haematological or rheumatological disease is limited. Methods<br />

We investigated retrospectively the clinical course of 47<br />

HEV infections in haematological (n=38) or rheumatological<br />

patients (n=9) within a multicentre observational study. Results<br />

Patients suffered from the following diseases: lymphoma n=17,<br />

myeloma n=3, chronic leukemia n=7, acute leukemia n=4,<br />

other haematological diseases n=7, rheumatoid arthritis, n=5,<br />

other rheumatological diseases=4. All patients received standard<br />

immunosuppressing therapy (calcineurin inhibitors: n=8,<br />

R-CHOP/R-Benda/R-ICE: n=7, cyclophosphamide n=7, other:<br />

n=28). 12/47 patients were stem cell transplant recipients<br />

(mean date of transplantation: 832 days before positive testing<br />

for HEV, range 20-4835 days). All patients tested positive for<br />

HEV RNA, and had laboratory signs of viral hepatitis (mean<br />

ALT 866 IU/ml, range 66-2849 IU/ml), while peak bilirubin<br />

was mean 7.3 mg/dl and peak creatinine was 90.2 mmol/l.<br />

All but one HEV infection had been acquired in Europe (1<br />

from Indonesia). Treatment with ribavirin (RBV) was initiated<br />

in 23/47 patients (5-22mg/kg bodyweight, duration: 1-32<br />

weeks, mean 14 weeks). Start of treatment ranged from immediately<br />

at the time of diagnosis up to 24 weeks after first detection<br />

of HEV (mean 9 weeks). 14/23 treated patients cleared<br />

the infection (61%), four patients (17%) failed to achieve viral<br />

clearance, one died (from deteriorating GvHd under treatment),<br />

two are still treated at the time of writing. 19 of the 24<br />

non-treated patients (79%) cleared the infection spontaneously,<br />

three died (two died directly after hepatitis E diagnosis from<br />

leukemia associated problems, one from GvHd), two are still<br />

under observation. RBV vs. untreated patients did not differ<br />

in peak ALT levels (RBV 836 IU/l, range 66-2775 IU/l vs. no<br />

treatment 931 IU/l, range 72-2849 IU/l), restitution of ALT<br />

after HEV infection (RBV 28 IU/l, range 11-85 IU/l (one patient<br />

with 937 IU/l) vs. no treatment 27 IU/l, range 13-76 IU/l)<br />

or in duration of HEV infection (RBV 27.1 weeks vs. no treatment<br />

27.5 weeks). ALT levels normalized in 37/47 patients<br />

(79%). Conclusion Chronic HEV infection is not restricted to<br />

solid organ transplant recipients but can also occur in immunosuppressed<br />

patients with haematological or rheumatological<br />

diseases. Ribavirin seems to be an efficient and safe treatment<br />

option. However, parameters that predict spontaneous clear-


1088A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ance need to be established in order to develop standards for<br />

initiating ribavirin therapy in immunosuppressed patients.<br />

Disclosures:<br />

Laurent Alric - Board Membership: Schering Plough, Schering Plough, Schering<br />

Plough, Schering Plough; Consulting: MSD; Speaking and Teaching: Roches,<br />

BMS, Gilead, Roches, BMS, Gilead, Roches, BMS, Gilead, Roches, BMS, Gilead,<br />

MSD, Abbvie<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

Jean-Marie Peron - Board Membership: BAYER; Consulting: BMS, GILEAD, BOS-<br />

TON SCIENTIFIC<br />

Robert A. de Man - Advisory Committees or Review Panels: Norgine; Grant/<br />

Research Support: Biotest<br />

The following authors have nothing to disclose: Johann von Felden, Vincent<br />

Mallet, Ansgar W. Lohse, Paul Schnitzler, Christoph Hoener zu Siederdissen,<br />

Maria Teresa Giordani, Celia Aitken, Jan Cornelissen, Dominik Bettinger, Robert<br />

Thimme, Sven Pischke<br />

1804<br />

Linkage to Care Outcomes for Hepatitis C Cases Identified<br />

in the Emergency Department<br />

Ricardo A. Franco, Edgar T. Overton, Ashutosh Tamhane, Jordan<br />

M. Forsythe, Joel B. Rodgers, Deanne Guthrie, Suneetha Thogaripally,<br />

Anne Zinski, Michael Saag, Michael Mugavero, James W.<br />

Galbraith; University of Alabama at Birmingham, Birmingham, AL<br />

Purpose: Screening is an essential strategy for hepatitis C detection<br />

and urban Emergency Departments (EDs) are high-yield<br />

testing sites. While a critical step in the Care Cascade, no prior<br />

efforts longitudinally evaluated linkage to care (LTC) outcomes<br />

in ED screening programs. This study reports interim results of<br />

integrated LTC interventions in this setting. Methods: Consecutive<br />

baby boomers presented for routine care and underwent<br />

opt-out HCV screening from Sept 2013 through June 2014.<br />

Educational materials were distributed by ED staff to HCV Ab<br />

positive subjects. A LTC coordinator undertook follow-up phone<br />

calls, charity care applications for the uninsured and letter mailing<br />

for contact failures. Electronic Health Record was queried<br />

for demographics, ICD-9 codes for major morbidity domains<br />

and visit logs. LTC was defined as at least one HCV clinic<br />

visit post screening. Results: Over 4300 baby boomers were<br />

screened and 473 tested positive for HCV Ab (11%). Viral<br />

load was checked in 402 subjects and 336 patients had confirmed<br />

HCV viremia (84%). Among viremic patients, the mean<br />

age was 57.3 years (SD± 4.8), 70% were males and 61%<br />

were African Americans. Forty-two percent had public insurance,<br />

14% were uninsured and 82% had multi-morbidity (≥ 2<br />

morbidity domains). Substance abuse (35%), psychiatric disease<br />

(29%) and cirrhosis (31%) were highly prevalent. A total<br />

of 161 baby boomers (48%) were hospitalized and 221 (66%)<br />

utilized multi-specialty care within campus during follow-up.<br />

Median follow-up was 433 days (IQR 354 - 500). The median<br />

time between screening and the first HCV care visit was 81<br />

days (IQR 40 - 205). Overall, 121 baby boomers (36%) did<br />

not attend any visits on campus, 94 (29%) attended non-HCV<br />

care visits only and 117 (35%) attended HCV care visits at<br />

least once during follow-up. Lack of HCV clinic attendance<br />

had independent associations with Males (aOR 1.79, 95%<br />

CI 1.0 – 3.19), Uninsured Status (aOR 5.62, 95% CI 1.57-<br />

20.11), Hospitalization (aOR 0.52, 95% CI 0.29 – 0.93) and<br />

Primary Care Attendance in the university system (aOR 0.22,<br />

95% CI 0.10 – 0.43). Conclusions: In this interim analysis,<br />

approximately 1 of every 3 HCV-infected baby boomers was<br />

linked to a HCV treatment center following ED screening. LTC<br />

interventions in this setting should be robust in order to address<br />

significant lack of access to care, frequent competing medical<br />

priorities and meet the great need created by high-yield screening.<br />

Further research is warranted for optimal LTC practices.<br />

Disclosures:<br />

Ricardo A. Franco - Advisory Committees or Review Panels: Gilead; Grant/<br />

Research Support: Gilead<br />

Michael Saag - Grant/Research Support: BMS, Gilead, Abbvie, Merck, ViiV,<br />

Janssen<br />

The following authors have nothing to disclose: Edgar T. Overton, Ashutosh<br />

Tamhane, Jordan M. Forsythe, Joel B. Rodgers, Deanne Guthrie, Suneetha Thogaripally,<br />

Anne Zinski, Michael Mugavero, James W. Galbraith<br />

1805<br />

Racial Differences in Socioeconomic Status (SES), Cirrhosis,<br />

and Treatment of Chronic Hepatitis C (CHC): Highest<br />

Treatment Rates in Caucasian Americans (CA), followed<br />

by Hispanic Americans (HA), African Americans (AFA)<br />

and Lowest in Asian Americans (AA)<br />

Philip Vutien 2,1 , Joseph K. Hoang 1 , Louis Brooks 3 , Mindie H.<br />

Nguyen 1 ; 1 Division of Gastroenterology and Hepatology, Stanford<br />

University Medical Center, Palo Alto, CA; 2 Rush University Medical<br />

Center, Chicago, IL; 3 Optuminsight, Eden Prairie, MN<br />

Background: In the USA, CHC disproportionately affects ethnic<br />

minorities with higher prevalences but their linkage to care<br />

remains understudied. Methods: We performed a retrospective<br />

study of a US adult cohort of 76,849 CHC patients identified<br />

via ICD9 query from OptumInsight’s Data Mart from 1/2009-<br />

12/2013. All patients had continuous medical and pharmacy<br />

coverage: 72,746 with commercial insurance and 4,103 with<br />

Medicaid. Results: The majority were male (61%) and older<br />

than 50 (62%). The proportions of patients with a BA/BS<br />

degree were highest in AA (28%), followed by CA (14%), HA<br />

(7%), and AFA (4%). Similarly, household income > 100,000<br />

USD were most frequently seen in AA (45%) and CA (36%)<br />

and less often in HA (20%) and AFA (15%). At presentation,<br />

HA were most likely to have cirrhosis compared to CA, AA,<br />

and AFA (35% vs. 31%, 30%, and 30%, respectively). AA<br />

were the least likely to have any psychiatric (8%) or medical<br />

comorbidity (54%) while CA were the most likely to have a<br />

psychiatric comorbidity (21%) and AFA to have a medical<br />

comorbidity (76%). Only 10% of all patients and 0.1% of<br />

Medicaid patients were prescribed anti-HCV treatment. AA<br />

had the lowest prescription rates at 7.8% overall and in each<br />

income category (Figure). In a multivariate analysis, CA was<br />

a significant predictor of anti-HCV treatment after adjustment<br />

for medical and psychiatric comorbidities, education, income,<br />

age, and cirrhosis with odds ratios of 1.52 for CA vs. AA,<br />

1.26 for CA vs. HA, and 1.18 for CA vs. BA. Conclusions: In a<br />

cohort of insured patients, ethnic minorities (HA, AFA, AA) had<br />

lower anti-HCV treatment rates than CA. Despite having fewer<br />

comorbidities and higher incomes and educational levels, AA<br />

had the lowest treatment rates. Future <strong>studies</strong> should aim to<br />

identify underlying ethnic-related barriers to anti-HCV treatment<br />

besides SES and comorbidities.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1089A<br />

of CHC. Older age and receiving treatment were significant<br />

predictors of developing cirrhosis.<br />

Disclosures:<br />

Louis Brooks - Employment: Optum<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Philip Vutien, Joseph K. Hoang<br />

1806<br />

Progression to liver cirrhosis in patients with chronic<br />

hepatitis C (CHC) in a large United States cohort: a natural<br />

history study<br />

Nghia H. Nguyen 2,1 , Joseph K. Hoang 1 , An K. Le 1 , Changqing<br />

Zhao 1,3 , Christine Y. Chang 1 , Richard H. Le 1 , Mingjuan Jin 1,4 ,<br />

Alina Kutsenko 5 , Lee Ann Yasukawa 6 , Jian Q. Zhang 7 , Susan C.<br />

Weber 6 , Mindie H. Nguyen 1 ; 1 Division of Gastroenterology and<br />

Hepatology, Stanford University Medical Center, Palo Alto, CA;<br />

2 Medicine, University of California, San Diego, San Diego, CA;<br />

3 Department of Cirrhosis,Institute of Liver Disease, Shuguang Hospital,<br />

Shanghai, China; 4 Department of Epidemiology and Biostatistics,<br />

Zhejiang University School of Public Health, Hangzhou,<br />

China; 5 Department of Medicine, Stanford University Medical<br />

Center, Palo Alto, CA; 6 Center for Clinical Informatics, Stanford<br />

University School of Medicine, Palo Alto, CA; 7 Chinese Hospital,<br />

San Francisco, CA<br />

Purpose: Most CHC <strong>studies</strong> in the U.S. are based on estimates<br />

from simulation models. Our goal was to describe CHC disease<br />

progression to cirrhosis in a real-life cohort from 2 ethnically<br />

diverse U.S. centers. Methods: ICD-9 query identified 9207<br />

consecutive CHC patients in 2005-2014. Analysis was performed<br />

from data obtained via individual chart review in 7105<br />

confirmed CHC cases. Kaplan Meier method was performed<br />

to estimate cumulative incidence of cirrhosis. Cox regression<br />

was performed to identify predictors of disease progression<br />

Results: The majority of our patients were men (62%), had<br />

genotype 1 (69%) and mean age of 55.4±10.7. Treatment<br />

rate was low (13%) in our cohort with the majority of treated<br />

patients having received interferon-based therapy (58%).<br />

Crude incidence for development of cirrhosis was 353.7 per<br />

10,000 person-years. On subgroup analyses, men (vs women)<br />

had a significantly higher overall 5-year cumulative incidence<br />

of cirrhosis (p=0.009) and also when stratified by age (


1090A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

12% self-reported daily alcohol use. Among 895 persons not<br />

treated for HCV, 181 developed ESLD and 30 developed HCC<br />

during a total follow-up of 7,896 person-years. Among 69<br />

persons who achieved an SVR, 5 developed ESLD and none<br />

developed HCC during a total follow-up of 437 person-years<br />

after completing HCV treatment. Among 98 persons who failed<br />

or discontinued HCV treatment, 11 developed HCC and 37<br />

developed ESLD during a total follow-up of 614 person-years.<br />

Compared with persons never treated for HCV, persons who<br />

failed treatment were more likely to develop ESLD (adjusted<br />

hazard ratio [a-HR]: 2.67, 95% confidence interval [CI] 1.82,<br />

3.91) and HCC (aHR: 4.22, CI: 1.88, 8.70). Persons who<br />

achieved SVR were less likely to develop ESLD (a-HR: 0.29, CI:<br />

0.09, 0.93). Conclusions: HCV RNA positive AN/AI persons<br />

failing to achieve SVR after IFN-based therapy are at increased<br />

risk of developing ESLD and HCC as compared to those never<br />

treated and should be prioritized for treatment with direct-acting<br />

antiviral agents.<br />

Disclosures:<br />

Lisa J. Townshend-Bulson - Grant/Research Support: Gilead<br />

Chriss E. Homan - Grant/Research Support: Gilead<br />

James E. Gove - Grant/Research Support: Gilead Sciences<br />

Julia N. Plotnik - Grant/Research Support: Gilead<br />

Hannah G. Espera - Grant/Research Support: Gilead<br />

Annette Hewitt - Grant/Research Support: gilead<br />

Youssef Barbour - Grant/Research Support: Gilead<br />

The following authors have nothing to disclose: Dana J. Bruden, Brian J. McMahon,<br />

Stephen Livingston, Brenna Simons-Petrusa, Susan Negus, Mary Snowball,<br />

Prabhu P. Gounder<br />

1808<br />

Benefits of HCV eradication in compensated cirrhotic<br />

patients extend beyond liver-related complications:<br />

results from the ANRS CO12 CirVir prospective cohort.<br />

Pierre Nahon 1 , Valerie Bourcier 1 , Richard Layese 2 , Ventzislava<br />

Petrov-Sanchez 3 , Dominique Guyader 4 , Sebastien Dharancy 5 ,<br />

Hélène Fontaine 6 , Vincent Thibault 7 , Vincent Leroy 8 , Stanislas<br />

Pol 6 , Françoise Roudot-Thoraval 2 ; 1 Hôpital Jean Verdier, Bondy,<br />

France; 2 Hôpital Henri Mondor, Créteil, France; 3 ANRS, Paris,<br />

France; 4 CHU Pontchaillou, Rennes, France; 5 Hôpital Huriez, Lille,<br />

France; 6 Hôpital Cochin, Paris, France; 7 GH Pitié-Salepêtrière,<br />

Paris, France; 8 Hôpital Michallon, Grenoble, France<br />

Backround and aims: The objective was to prospectively assess<br />

the benefits of sustained viral response (SVR) on the incidence<br />

of liver-related and extra-hepatic complications in French<br />

patients with compensated HCV cirrhosis using competing<br />

risk analyses. Methods: Patients with the following inclusion<br />

criteria were enrolled in 35 French centres: a) biopsy-proven<br />

HCV cirrhosis; b) Child-Pugh A; c) absence of previous liver<br />

complications. Patients were prospectively followed-up every<br />

6 months. Sustained virological response (SVR) was defined<br />

as undetectable HCV RNA at the end of a 12-week untreated<br />

follow-up period. A given event was arbitrarily considered as<br />

occurring in a patient who achieved SVR if it was recorded at<br />

least one year after successful treatment completion. Results:<br />

A total of 1323 patients were enrolled between March 2006<br />

and June 2012 [age 55, men 63.5%, Genotype 1: 66%,<br />

2: 6 %, 3: 17%, 4: 10%, 5-6: 1%]. At end-point in January<br />

2015 (median follow-up of 51 months), the number (rates)<br />

of HCV patients with a negative viral load was 653 (51%)<br />

corresponding to SVR in 491 patients (40%) while the remaining<br />

162 HCV negative patients were still undergoing antiviral<br />

treatment (mostly based on second generation DAAs). Primary<br />

liver cancer (PLC) was diagnosed in 168 (162 hepatocellular<br />

carcinoma (HCC) and 6 cholangiocarcinomas) among which<br />

17 were developed in SVR patients (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1091A<br />

Liver histological examination revealed that 7 patients had<br />

fibrosis grade F0, 149 had F1, 87 had F2, 58 had F3, 16 had<br />

F4. Both serum levels of Fuc-Hpt and M2BP increased along<br />

the progression of liver fibrosis stage. The median of serum<br />

Fuc-Hpt levels in patients with fibrosis grade F0-1, F2, F3, and<br />

F4 were 433, 592, 1022, and 1790 U/ml respectively. The<br />

M2BP median with F0-1, F2, F3, and F4 were 1261, 2276,<br />

2740, and 4044 ng/ml. We evaluated the diagnostic ability<br />

of aspartate aminotransferase to platelet ratio index (APRI),<br />

FIB4 index, Fuc-Hpt and M2BP in liver fibrosis grade F ≥ 2, F ≥<br />

3 and F4 by ROC analysis. The AUROC of Fuc-Hpt in F ≥ 2, F<br />

≥ 3 and F4 were 0.677, 0.738 and 0.813 respectively. M2BP<br />

were 0.747, 0.734 and 0.797. Most AUROC of Fuc-Hpt and<br />

M2BP were over 0.7 in all fibrotic stages, suggesting that both<br />

Fuc-Hpt and M2BP could be applicable for evaluating liver<br />

fibrosis. However, these markers did not exceed the values of<br />

known fibrotic index such as APRI or FIB4 index. To establish<br />

better liver fibrosis markers, we evaluated the combination of<br />

known marker with new marker. To evaluate for fibrosis grade<br />

F ≥ 2, the AUROC of FIB4 index adding M2BP was most diagnosable;<br />

0.800. To evaluate for fibrosis grade F ≥ 3 and F4,<br />

FIB4 index adding Fuc-Hpt is most diagnosable; 0.841 and<br />

0.913. We examined to identify the factors associated with the<br />

progression of liver fibrosis grade F ≥ 2 and F ≥ 3. In multivariate<br />

analysis revealed that M2BP, FIB4 index and AFP were the<br />

independent factors associated with liver fibrosis grade F ≥ 2.<br />

And Fuc-Hpt, FIB4 index and AFP were the independent factors<br />

in F ≥ 3. Finally we evaluated the cumulative hepatocellular<br />

carcinoma (HCC) incidence rate. The follow up period after<br />

liver biopsy is 1148 ± 913 days. HCC was occurred in 19<br />

patients during follow up periods. The cumulative incidence of<br />

HCC was significantly higher in the patients with high serum<br />

levels of Fuc-Hpt (> 559 U/ml; median) or high M2BP (> 1847<br />

ng/ml; median) than in patients with low levels of them by log<br />

rank test. Conclusion: Both Fuc-Hpt and M2BP could be useful<br />

biomarkers to evaluate liver fibrosis in patients with chronic<br />

hepatitis C.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Seiichi Tawara, Tomohide<br />

Tatsumi, Sadaharu Iio, Ichizo Kobayashi, Shigeki Suemura, Atsuo Takigawa,<br />

Tadashi Kegasawa, Takahiro Suda, Teppei Yoshioka, Yoshiki Onishi, Akira<br />

Nishio, Satoshi Aono, Minoru Shigekawa, Ryotaro Sakamori, Naoki Hiramatsu,<br />

Eiji Miyoshi<br />

1810<br />

An Integrated Linkage to Care Model to Improve Patient<br />

Engagement and Retention in HCV Specialty Care<br />

Catelyn Coyle 1 , Kendra Viner 2 ; 1 National Nursing Centers Consortium,<br />

Philadelphia, PA; 2 Disease Control, Philadelphia Department<br />

of Public Health, Philadelphia, PA<br />

Purpose: To describe a successful linkage to care model piloted<br />

in a network of community health centers in Philadelphia, PA.<br />

Methods: Hepatitis C (HCV) patients are lost at each stage<br />

of the HCV continuum of care. In September 2013, National<br />

Nursing Centers Consortium implemented a linkage model<br />

to usher HCV-positive patients receiving care in one of five<br />

Philadelphia based community health centers into specialty<br />

care. Eligible patients were those testing HCV-antibody positive<br />

after October 1, 2012. The model included a linkage to care<br />

coordinator, EMR modifications, and the following services to<br />

help patients overcome barriers: escorts to appointments, transportation<br />

assistance (including tokens), appointment scheduling,<br />

coordination with sub-specialty offices, and connection to<br />

social services. EMR changes, including prompts, result summary<br />

flow sheets, and weekly reports, ensured that patients<br />

were followed through the entire care cascade. Patients with<br />

missed appointments were contacted to reschedule and identify<br />

barriers. Calls to emergency contacts and site visits were made<br />

to reach any patients lost to care. Results: During September<br />

2013 - February 2015, 4,893 patients were HCV tested. Of<br />

these, 601 (12%) had a positive HCV-antibody test, 584 (97%)<br />

received an HCV-RNA test, and 406 (69%) were identified<br />

with current HCV infection. Of the 406, 366 (90%) received<br />

their HCV-RNA positive result, 327 (81%) were referred to an<br />

HCV care provider, and 231 (57%) were seen by the referred<br />

HCV care provider. During the 18-month study period, testing<br />

and linkage to care rates ranged across the five testing sites:<br />

97%-100% for HCV-antibody-positive patients receiving HCV-<br />

RNA testing, 53%-73% for HCV-RNA positivity, 83%-100%<br />

for receipt of a positive HCV-RNA test result, 67%-89% for<br />

referral to HCV care, and 37%-72% for HCV care provider<br />

visits. Lack of insurance, transportation, weather-appropriate<br />

clothing, and being signed out of shelters were identified as<br />

the most common social barriers to care. During the 11 months<br />

before the addition of a linkage to care coordinator (October<br />

2012-August 2013) and the 11 months afterwards (September<br />

2013-July 2014), the number of HCV-infected patients receiving<br />

their positive HCV-RNA result rose by 68%, patient referrals<br />

by 49%, and patients seen for HCV care by 29%. Conclusion:<br />

Intensive linkage services can increase the number of patients<br />

who successfully navigate the HCV treatment cascade. The linkage<br />

to care coordinator is an important position that acts as a<br />

trusted intermediary for patients being linked to care.<br />

Disclosures:<br />

Catelyn Coyle - Advisory Committees or Review Panels: Gilead Sciences<br />

The following authors have nothing to disclose: Kendra Viner<br />

1811<br />

Increasing Hepatitis C Virus (HCV) Screening and Linkage<br />

to Care in a Large Integrated Health System<br />

Carla V. Rodriguez 1 , Kevin Rubenstein 1 , Haihong Hu 1 , Benjamin<br />

P. Linas 2,3 , Michael A. Horberg 1 ; 1 Mid-Atlantic Permanente<br />

Research Institute, Kaiser Permanente, Rockville, MD; 2 HIV Epidemiology<br />

and Outcomes Research Unit, Section of Infectious Diseases,<br />

Boston Medical Center, Boston, MA; 3 Epidemiology, Boston<br />

University School of Public Health, Boston, MA<br />

Objective: To describe the HCV cascade of care from screening<br />

to linkage in Kaiser Permanente, Mid-Atlantic States (KPMAS)<br />

relative to the 2013 release the U.S. Preventative Services Task<br />

Force “birth cohort” (born 1945-1965) screening recommendations.<br />

Methods: Cohort study among patients in KPMAS with<br />

≥8 months of enrollment from 1/1/2003-12/31/2014 and<br />

≥1 clinical visit. Birth cohort, IDU, MSM, sex, race, clinic location,<br />

income, elevated ALT (2 consecutive ALT >60 U/L), HIV<br />

and HBV status were characteristics of interest. We estimated<br />

the annual screening rate as the number antibody (Ab) tested<br />

per persons enrolled each year. We used survival methods to<br />

describe factors associated with time to Ab testing, stratified by<br />

location; and logistic regression to identify predictors of GI/ID<br />

referral. Among Ab+, we describe the cumulative probability<br />

and predictors of confirmatory RNA or genotype testing, GI/ID<br />

referral, and HCV treatment. Results: We observed 665,345<br />

patients over an 11 year period. The annual screening rate<br />

increased steadily from 23.6 to 70.8 per 1,000 person-years<br />

from 2004 to 2014; with the sharpest increase after 2013. A<br />

total of 123,572 (18.6%) KPMAS members were screened for<br />

HCV. Screening among the birth cohort was lower than non-co-


1092A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

hort members. However, the gap began to shrink during later<br />

years of study. Across all locations, screening was increased<br />

among drug users, HBV and HIV status, and elevated ALT.<br />

Screening was lower among men. Among Ab+, 84% were<br />

RNA tested. In an adjusted model, we found no differences<br />

in confirmatory testing by location, MSM, race, drug use, HIV<br />

or HBV status. Elevated ALT (aHR=1.3; ref=ALT


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1093A<br />

for (1) appropriate resource allocation to meet the needs of<br />

the affected patients and (2) strategic deployment of antiviral<br />

therapy to mitigate the current and future burden of HCV.<br />

Disclosures:<br />

W. Ray Kim - Advisory Committees or Review Panels: Bristol Myers Squibb,<br />

Gilead Sciences, Abbvie, Merck<br />

Nancy Reau - Advisory Committees or Review Panels: Jannsen, Merck, AbbVie,<br />

Intercept, Salix, BMS, Jannsen; Grant/Research Support: Merck, Gilead, BMS,<br />

AbbVie, Jannsen, BI<br />

Mitchell L. Shiffman - Advisory Committees or Review Panels: Merck, Gilead,<br />

Boehringer-Ingelheim, Bristol-Myers-Squibb, Abbvie, Janssen, Acchillion; Consulting:<br />

Roche/Genentech; Grant/Research Support: Merck, Gilead, Boehringer-Ingelheim,<br />

Bristol-Myers-Squibb, Abbvie, Beckman-Coulter, Achillion, Lumena,<br />

Intercept, Novartis, Gen-Probe; Speaking and Teaching: Roche/Genentech,<br />

Merck, Gilead, Abbvie, Janssen, Bayer<br />

Tarek I. Hassanein - Advisory Committees or Review Panels: AbbVie, Bristol-Myers<br />

Squibb; Grant/Research Support: AbbVie Pharmaceuticals, Obalon, Bristol-Myers<br />

Squibb, Eiasi Pharmaceuticals, Gilead Sciences, Janssen R&D, Idenix<br />

Pharmaceuticals, Ikaria Therapeutics, Merck Sharp & Dohme, NGM BioPharmaceuticals,<br />

Ocera Therapeutics, Salix Pharmaceuticals, Sundise, TaiGen Biotechnology,<br />

Takeda Pharmaceuticals, Vital Therapies, Tobria; Speaking and<br />

Teaching: Baxter, Bristol-Myers Squibb, Gilead, Salix, AbbVie<br />

The following authors have nothing to disclose: Henri Cournand<br />

1814<br />

Patients with Hepatitis C (HCV) Advanced Cirrhosis and<br />

the rs738409 GC/GG Genotype of Platatin-Like Phosphatase<br />

Domain Containing (PNPLA-3) are at Risk for<br />

Further Decompensation Despite Direct-Acting Antiviral<br />

(DAA) Treatment<br />

Winston Dunn 1 , Anusha Vittal 1 , Melissa Whitener 1 , Jie Zhao 1 ,<br />

Brian Bridges 1 , Shweta Chakraborty 1 , K. James Kallail 2 , Prashant<br />

K. Pandya 3 , Chuanghong Wu 4,1 , Ryan Taylor 1 , Jody C. Olson 1 ,<br />

Richard Gilroy 1 , Mojtaba S. Olyaee 1 , Timothy Schmitt 1 , Steven<br />

A. Weinman 1 ; 1 The University of Kansas Medical Center, Kansas<br />

City, KS; 2 Office of Research, The University of Kansas School of<br />

Medicine-Wichita, Wichita, KS; 3 Kansas City VA Medical Center,<br />

Kansas City, MO; 4 Shekou People’s hospital, Shenzhen, China<br />

HCV patients with compensated cirrhosis generally do well<br />

after successful virological cure, but approximately 1/3 of<br />

patients with decompensated cirrhosis experience further<br />

decompensation or develop hepatocellular carcinoma (HCC).<br />

The PNPLA3 genotype is correlated with fibrosis progression<br />

from HCV. We tested the hypothesis that the PNPLA3 genotype<br />

can predict recovery versus further decompensation in<br />

advanced cirrhosis after HCV is cured. We studied 27 patients<br />

with advanced HCV cirrhosis who were treated with an interferon-free<br />

DAA-based treatment regimen. We excluded patients<br />

with HCC, previous transplantation, and censored patients<br />

upon transplantation. Nine patients had bilirubin ≥2 mg/dl or<br />

INR ≥2 before DAA treatment. All remaining patients had clinical<br />

decompensation. Eight were listed for transplantation. The<br />

median follow-up was 166 days. The primary outcome was<br />

a change in Model for End Stage Liver Disease (MELD) from<br />

before DAA treatment. Figure 1 shows a comparison of before<br />

DAA treatment to the end of follow-up. Two patients, both with<br />

the GC genotype, died of sepsis and variceal bleeding, and<br />

one patient each with GC and CC genotypes underwent liver<br />

transplantation. Using a random coefficient model that considers<br />

repeated measurement of MELD and adjusting for baseline<br />

MELD and duration since treatment, patients with the CC allele<br />

dropped an average of 2.5 (95% confidence interval, 1.4 -<br />

3.6) MELD points compared to patients with the GC/GG allele<br />

(p = 0.0001). The effect remained significant after adjusting<br />

for age, gender, diabetes, and BMI (p = 0.02).This prospective<br />

cohort study showed a statistically significant and clinically<br />

meaningful difference in which patients with HCV advanced<br />

cirrhosis and an unfavorable PNPLA3 genotype are at risk<br />

for further decompensation despite virological cure. PNPLA3<br />

genotyping may be clinically useful to better target patients for<br />

liver transplant evaluation.<br />

Disclosures:<br />

Prashant K. Pandya - Grant/Research Support: Gilead, Merck; Speaking and<br />

Teaching: Genentech, AbbVie, Vertex<br />

Jody C. Olson - Advisory Committees or Review Panels: Baxter<br />

Richard Gilroy - Speaking and Teaching: Salix, NPS, Gilead, AbbVie, Novartis<br />

Steven A. Weinman - Consulting: Cardax, Inc.<br />

The following authors have nothing to disclose: Winston Dunn, Anusha Vittal,<br />

Melissa Whitener, Jie Zhao, Brian Bridges, Shweta Chakraborty, K. James<br />

Kallail, Chuanghong Wu, Ryan Taylor, Mojtaba S. Olyaee, Timothy Schmitt<br />

1815<br />

Natural History of Decompensated Cirrhosis after<br />

Direct-Acting Antivirals (DAA) Treatment of Hepatitis C<br />

(HCV)<br />

Winston Dunn 1 , Melissa Whitener 1 , Shweta Chakraborty 1 , Anusha<br />

Vittal 1 , Prashant K. Pandya 2 , K. James Kallail 3 , Brian Bridges 1 ,<br />

Chuanghong Wu 4 , Ryan Taylor 1 , Jody C. Olson 1 , Richard Gilroy 1 ,<br />

Timothy Schmitt 1 , Mojtaba S. Olyaee 1 , Steven A. Weinman 1 ; 1 The<br />

University of Kansas Medical Center, Kansas City, KS; 2 Kansas<br />

City VA Medical Center, Kansas City, MO; 3 Office of Research,<br />

The University of Kansas School of Medicine-Wichita, Wichita, KS;<br />

4 Shekou People’s hospital, Shekou, China<br />

With the recent approval of newer DAA, more than 95% of<br />

patients with HCV infections can achieve a Sustained Virological<br />

Response (SVR). Although patients with compensated<br />

cirrhosis generally do well after viral clearance, the outcome<br />

for decompensated cirrhosis after SVR is not well defined. This<br />

study reports the outcomes of a cohort of decompensated HCV<br />

cirrhosis patients undergoing interferon free DAA-treatment in<br />

an academic transplant center. We included 96 patients in<br />

total all of whom had one or more indications of advanced disease<br />

such as hepatocellular carcinoma (HCC, n=21), clinical<br />

decompensation defined as hepatic encephalopathy, ascites,<br />

or variceal bleeding (n=81) or biochemical decompensation<br />

(bilirubin ≥2 mg/dl or INR ≥1.7, n=47). The median follow-up<br />

was 181 days. Six patients received liver transplantation and<br />

3 died. Three patients had to discontinue treatment due to<br />

adverse events. SVR occurred in 33/45 eligible (adequate


1094A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

follow-up time) patients. Post treatment MELD score changes<br />

showed considerable heterogeneity with SVR patients showing<br />

an average MELD decrease of 2.0 (95% confidence interval<br />

0.9 - 3.1, p=0.0003) MELD points compared to patients who<br />

failed SVR. Even within the SVR group, 33% of the patients<br />

had post treatment increases in MELD score while 63% had a<br />

decrease. Four patients had refractory ascites requiring paracentesis<br />

prior to treatment and none required paracentesis after<br />

SVR. One patient developed refractory ascites after treatment.<br />

Four patients had hepatic encephalopathy (PSE) related hospital<br />

admission prior to treatment but only 1 had further PSE<br />

related admissions after treatment. Three new patients developed<br />

PSE related admission for the first time after treatment,<br />

and one patient developed HCC after SVR. In summary, 3/4 of<br />

decompensated HCV cirrhotics were able to achieve SVR with<br />

approximately 2/3 of the SVR patients achieving decreased<br />

MELD score. Most patients with refractory symptoms had clinical<br />

benefits such as reduced paracentesis and encephalopathy<br />

related admissions. Improvement is not universal and patients<br />

should be carefully monitored. DAA treatment for advanced<br />

cirrhosis is not a substitution for transplant. For those patients<br />

who did not achieve SVR, deterioration occurred in half. HCV<br />

treatment is thus indicated, even in the most advanced cases<br />

and a significant proportion of these patients may improve to<br />

the point of not requiring liver transplantation.<br />

Disclosures:<br />

Prashant K. Pandya - Grant/Research Support: Gilead, Merck; Speaking and<br />

Teaching: Genentech, AbbVie, Vertex<br />

Jody C. Olson - Advisory Committees or Review Panels: Baxter<br />

Richard Gilroy - Speaking and Teaching: Salix, NPS, Gilead, AbbVie, Novartis<br />

Steven A. Weinman - Consulting: Cardax, Inc.<br />

The following authors have nothing to disclose: Winston Dunn, Melissa Whitener,<br />

Shweta Chakraborty, Anusha Vittal, K. James Kallail, Brian Bridges, Chuanghong<br />

Wu, Ryan Taylor, Timothy Schmitt, Mojtaba S. Olyaee<br />

1816<br />

Low HCV Treatment for Chronic Hepatitis C (CHC)<br />

Patients in Pre-Protease Inhibitor (PI) and Post-PI/Pre-Direct<br />

Acting Antiviral (DAA) Eras Compared to Post-DAA<br />

Era, Especially in African Americans: Analysis of A<br />

Large Real-World Cohort of 7105 Patients<br />

Peter T. Nguyen 2,1 , Nghia H. Nguyen 3,1 , Joseph K. Hoang 1 ,<br />

Changqing Zhao 1,4 , An K. Le 1 , Christine Y. Chang 1 , Richard H.<br />

Le 1 , Michael D. Nguyen 1 , Mingjuan Jin 5,1 , Susan C. Weber 6 , Lee<br />

Ann Yasukawa 6 , Mindie H. Nguyen 1 ; 1 Division of Gastroenterology<br />

and Hepatology, Stanford University Medical Center, Palo<br />

Alto, CA; 2 University of Texas Medical Branch, Galveston, TX;<br />

3 Department of Medicine, University of California, San Diego, San<br />

Diego, CA; 4 Department of Cirrhosis, Institute of Liver Disease,<br />

Shuguang Hospital, Shanghai, China; 5 Department of Epidemiology<br />

and Biostatistics, School of Public Health, Zhejiang University,<br />

Hangzhou, China; 6 Center for Clinical Informatics, Stanford University<br />

School of Medicine, Palo Alto, CA<br />

Background & Aim: Utilization of PEG IFN+RBV (P/R) is generally<br />

low for CHC due to adverse events (AE) and low SVR.<br />

More efficacious triple P/R+PI therapy was approved in 2011<br />

but treatment rate may also remain low due to AE. Our goal<br />

was to measure HCV treatment rates in a large ethnically<br />

diverse U.S. single-center cohort and changes from 2005-<br />

2014. Methods: ICD-9 query identified 9207 patients and<br />

individual charts were reviewed and confirmed HCV diagnosis<br />

in 7105 patients. Of these, 929 received anti-HCV treatment<br />

(P or P/R-based DAA-based). Results: Overall, treatment rate<br />

was low (Figure). From 2005-2013, the proportion of treated<br />

patients per all CHC patients with at least one clinical encounter<br />

ranged 1.0-3.7% and 3.7-14.3% per new CHC patients.<br />

Treatment rates per new HCV patients from 2005-07, 2008-<br />

10, 2011-13, and 2014 were significantly different: 4.4%,<br />

11.5%, 9.3%, and 46.5% (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1095A<br />

E (clinical illness, with anti-HEV IgM +ve) a year ago, when<br />

their unit was affected by an outbreak (n=67), (B) other soldiers<br />

in the unit who did not have symptomatic hepatitis (n=133),<br />

and (C) soldiers in another unit located elsewhere in the city,<br />

which was not affected by the outbreak (n=196). All sera were<br />

tested for anti-HEV IgG using 3 commercial assays (Wantai<br />

[W], China; DSI [D], Italy; and MP Biomedical [M], Singapore);<br />

the manufacturers’ protocols were followed. Results: In<br />

each group, assay W had the highest sensitivity (Table). After<br />

excluding indeterminate results, which were most common with<br />

assay M, the 3 assays showed positive concordant results in<br />

55, 16 and 15 sera, and negative concordant results in 2, 75<br />

and 133 sera, in groups A, B and C, respectively. Discordant<br />

results were found in 10, 30 and 36 sera. Using positive results<br />

in ≥2 assays as the criterion-standard, the sensitivity rates of<br />

assays W, D and M were 100%, 100%, and 87%, respectively<br />

in group A; 100%, 95% and 47% in group B; and 97%,<br />

94% and 58% in group C. Similarly, using negative results<br />

in ≥2 assays as the criterion-standard, the specificity rates of<br />

assays W, D and M were 75%, 75%, and 100%, respectively<br />

in group A; 96%, 99% and 95% in group B; and 92%, 100%<br />

and 95% in group C. Overall assays W and D had good concordance<br />

(97%, 95% and 91% in A, B and C). Conclusion: Of<br />

the three assays, W and D had a higher sensitivity and high<br />

degree of concordance with each other. In comparison, assay<br />

M was less sensitive and had less concordance with the other<br />

two assays. Also, assay M had more indeterminate test results.<br />

These data indicate that assays W and D may be preferred<br />

over assay M for anti-HEV seroprevalence <strong>studies</strong>.<br />

Results with the three serological assays<br />

Indet. = Indeterminate<br />

Disclosures:<br />

The following authors have nothing to disclose: Rakesh Aggarwal, Amit Goel,<br />

Gurdeep Singh<br />

1818<br />

Hepatitis C Virus (HCV) Prevalence among People who<br />

Inject Drugs (PWIDs) in Switzerland<br />

Francesco Negro 1 , Philip Bruggmann 2 , Sarah Blach 3 , Pierre<br />

Deltenre 4 , Jan Fehr 5 , Roger Kouyos 5 , Daniel Lavanchy 6 , Beat Mullhaupt<br />

7 , Homie Razavi 3 , Patrick Schmid 8 , David Semela 9 , Martin<br />

Stoeckle 10 , Andri Rauch 11 ; 1 Gastroenterology, Hepatology and<br />

Clinical pathology, University Hospitals, Geneva-14, Switzerland;<br />

2 ARUD Center for addiction medicine, Zurich, Switzerland; 3 Center<br />

for Disease Analysis, Lousville, CO; 4 Gastroenterology and<br />

hepatology, University Hospital, Lausanne, Switzerland; 5 Infectious<br />

Diseases & Hospital Epidemiology, University Hospital, Zurich,<br />

Switzerland; 6 Consultant, Denges, Switzerland; 7 Gastroenterology<br />

and hepatology, University Hospital, Zurich, Switzerland;<br />

8 Infectious diseases, Cantonal Hospital, St Gallen, Switzerland;<br />

9 Gastroenterology, Cantonal Hospital, St. Gallen, Switzerland;<br />

10 Infectious Diseases & Hospital Epidemiology, University Hospital,<br />

Basel, Switzerland; 11 Infectious diseases, University Hospital,<br />

Berne, Switzerland<br />

Background - In Switzerland, HCV among PWIDs has been<br />

decreasing due to active harm reduction efforts and an aging<br />

population (average age – 44 yrs). Expert consensus and estimates<br />

from the Swiss Federal Office of Public Health suggest<br />

that there are between 8,000 and 12,000 active PWIDs in<br />

Switzerland, and that 42% (27%-58%) of PWIDs are HCV<br />

infected. In addition, an estimated 17,000 – 25,700 individuals<br />

were enrolled in opioid substitution therapy (OST) and<br />

1,598 in heroin substitution therapy (HeGeBe) with 300,000<br />

syringes distributed monthly in 2012. Among individuals on<br />

OST and HeGeBe, an estimated 27.4% and 54%, respectively,<br />

continued to inject while on treatment. Understanding HCV<br />

transmission dynamics among high-risk populations requires<br />

robust epidemiological data and country-specific mathematical<br />

modeling to assess the potential impact of new HCV treatment<br />

strategies. Recent therapeutic advances promise greater convenience<br />

(oral therapies) with higher efficacy (>90% sustained<br />

viral response) and shorter duration of treatment. Methods -<br />

HCV transmission was modeled using cohorts to track HCV<br />

incidence and prevalence among PWIDs in the general population,<br />

as well as active PWIDs enrolled in OST and/or needle<br />

exchange programs (NEP). Model assumptions were derived<br />

from published literature and expert consensus. The relative<br />

impact of increasing treatment among PWIDs was considered,<br />

including the annual number needed to treat in order to reduce<br />

the HCV-infected PWID population by 2030. Results - If the<br />

current transmission paradigm continues, there are projected<br />

to be 3,150 HCV infected PWIDs in 2030. Annually treating<br />

45 HCV-infected PWIDs (1% of HCV-infected PWID population<br />

in 2014) resulted in an 11% reduction in HCV-infected<br />

PWIDs by 2030, while annual treatment of 200 PWIDs (4.5%<br />

of 2014 population) resulted in a reduction of over 50% by<br />

2030. Treating 322 PWIDs annually (7.5% of 2014 population)<br />

resulted in a >90% reduction in HCV-infected PWIDs by<br />

2030. Targeting treatment to PWIDs engaged in OST and NEP<br />

provided the greatest reduction in prevalence for the number of<br />

individuals treated. To achieve a 1-person reduction in overall<br />

prevalence by 2030, it was necessary to treat 1.6 PWIDs in<br />

OST/NEP as compared with 3.4 PWIDs in the general population.<br />

Conclusions - Treating a relatively small number of PWIDs<br />

results in substantial decreases in the HCV infected PWID population<br />

by 2030. Additionally, the impact of treatment is higher<br />

among PWIDs engaged in harm reduction programs. These<br />

data support implementation of a screening and treatment strategy<br />

among PWIDs, and particularly among PWIDs engaged<br />

in OST and NEP.<br />

Disclosures:<br />

Francesco Negro - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

MSD, Gilead, AbbVie; Grant/Research Support: Gilead<br />

Philip Bruggmann - Advisory Committees or Review Panels: Merck, Gilead, BMS,<br />

Abbvie, BMS; Grant/Research Support: BMS, Merck, Janssen, Gilead, Abbvie;<br />

Speaking and Teaching: BMS<br />

Sarah Blach - Employment: Center for Disease Analysis<br />

Pierre Deltenre - Consulting: Abbvie, Gilead, BMS; Grant/Research Support:<br />

Abbvie, Gilead, BMS, Janssen, MSD; Speaking and Teaching: Abbvie, Gilead,<br />

BMS<br />

Beat Mullhaupt - Consulting: MSD, Novartis, MSD, Janssen; Grant/Research<br />

Support: Bayer, Gillead<br />

Homie Razavi - Management Position: Center for Disease Analysis<br />

Andri Rauch - Advisory Committees or Review Panels: Gilead Sciences, Abbvie,<br />

MSD, Janssen Cilag<br />

The following authors have nothing to disclose: Jan Fehr, Roger Kouyos, Daniel<br />

Lavanchy, Patrick Schmid, David Semela, Martin Stoeckle


1096A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1819<br />

The clinical profiles of Japanese patients with L31M and<br />

Y93H variants in NS5A region of hepatitis C virus<br />

Kei Ogata 1 , Tatsuya Ide 1 , Teruko Arinaga-Hino 1 , Ichiro Miyajima<br />

1 , Reiichiro Kuwahara 1 , Keisuke Amano 1 , Takumi Kawaguchi<br />

1 , Toru Nakamura 1 , Masahito Nakano 1 , Hironori Koga 1 , Ryoko<br />

Kuromatsu 1 , Yuichiro Eguchi 2 , Masaru Harada 3 , Takuji Torimura 1 ;<br />

1 Division of Gastroenterology, Department of Medicine, Kurume<br />

Univ.School of Medicine, Kurume, Japan; 2 Department of Internal<br />

Medicine, Faculty of Medicine, Saga University, Division of Hepatology,<br />

Saga, Japan; 3 The Third Department of Internal Medicine,<br />

School of Medicine, University of Occupational and Environmental<br />

Health, Kitakyusyu, Japan<br />

Background and aims: Therapy for hepatitis C patients has<br />

been developed, and several direct antiviral agents (DAAs)<br />

have been approved clinically. The NS5A replication complex<br />

inhibitor is one of the important DAAs. It is known that<br />

resistance-associated variants (RAVs) naturally occur at NS5A<br />

positions L31 and Y93. Since the clinical background and<br />

profiles are still unclear, we investigated RAVs by direct<br />

sequencing and cycleave methods. Patients and method:<br />

Patients with genotype 1 and who were treatment naive for<br />

NS5A inhibitors were selected. We measured the Y93 positions<br />

in 1,204 patients and L31 positions in 1,039 patients<br />

by direct sequencing. We also measured the Y93 positions in<br />

1,039 patients by the cycleave method (PLoS One. 2014, Nov<br />

14:e112647). Among them, 520 patients of age, sex, with or<br />

without cirrhosis, HCV RNA level, ALT, platelet count, albumin,<br />

gamma-GTP, type IV collagen, prothrombin time and αfetoprotein,<br />

Il28B were also analyzed as clinical features. Results:<br />

In the 1,204 patients, the Y93H variant was detected in 165<br />

patients (13.7%). Among these 165 patients, 150 patients<br />

were measured by the cycleave method : 25.3% (38/150) of<br />

the patients had 100% of the Y93H variant, and the remaining<br />

patients had a mixed Y93H and Y93 wild type (Y93H variant,<br />

0-20% ; 52 patients (34.6%), 21-40% ; 18 patients (12.0%),<br />

41-60% ; 15 patients (10.0%), and 61-99% ; 27 patients<br />

(18.0%). Patients with the Y93H variant had a significantly<br />

higher HCV RNA level (6.4 ± 0.3 vs. 6.0 ± 0.7 log IU / ml,<br />

respectively, p = 0.0016), lower platelet count (106 ± 45 vs.<br />

134 ± 66 x10 3 /ml, respectively, p = 0.0048), and an older<br />

age (72.1 ± 7.9 vs. 68.6 ± 8.9, yrs, respectively, p = 0.011)<br />

than patients with the Y93 wild type. On the other hand, the<br />

L31M variant was detected in 4.5% of patients (47/1,039).<br />

Patients with the L31M variant had a significantly lower HCV<br />

RNA level (5.5 ± 0.8 vs. 6.0 ± 0.7, log IU/ml, respectively,<br />

p = 0.0021), higher gamma-GTP (75.3 ± 79 vs. 50.2 ± 57,<br />

IU/L, respectively, p = 0.0074), and a higher proportion of<br />

males (10.1vs. 2.7%, respectively) than patients with the L31<br />

wild - type. Conclusion: We demonstrated the clinical features<br />

of naturally occurring L31 and Y93 variants. HCV RNA levels<br />

are closely related to RAVs. Although careful attention should<br />

be paid on treatment with NS5A inhibitors, the interaction<br />

among RAVs, these clinical features, and treatment efficacy<br />

should be investigated.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kei Ogata, Tatsuya Ide, Teruko<br />

Arinaga-Hino, Ichiro Miyajima, Reiichiro Kuwahara, Keisuke Amano, Takumi<br />

Kawaguchi, Toru Nakamura, Masahito Nakano, Hironori Koga, Ryoko Kuromatsu,<br />

Yuichiro Eguchi, Masaru Harada, Takuji Torimura<br />

1820<br />

Hepatitis C Virus Genotype Analysis in Chronic Hepatitis<br />

Patients and Individuals with Spontaneous Virus Clearance<br />

Using a Newly Developed Serotyping Assay<br />

Ruifeng Yang, Huiying Rao, Lai Wei; Peking University People’s<br />

Hospital, Peking University Hepatology Institute, Beijing, China<br />

Objective We developed and evaluated the performance of<br />

an HCV serotyping assay in chronic hepatitis C(CHC) patients,<br />

and detected the genotypes in people with spontaneous viral<br />

clearance(SVC). Methods The novel serotyping assay was<br />

developed based on competitive ELISA principle. 997 CHC<br />

patients were enrolled in a nationwide cross-sectional study.<br />

Their samples were genotyped originally using the LiPA 2.0<br />

assay. For samples with inconsistent results yielded by the serotyping<br />

assay, sequences of the NS5B or 5’-UTR were analyzed<br />

to confirm the genotype results. In these patients, 190 patients<br />

achieved sustained virological response(SVR). The paired serotyping<br />

results were compared before and after treatment. Moreover,<br />

326 serum samples from a long-term follow-up cohort<br />

were also serotyped, among whom 66 were from SVC individuals,<br />

and the remaining 260 had persistent HCV RNA. All<br />

patients provided informed consent. Results There were 958<br />

out of 997 samples from CHC patients which yielded both<br />

genotyping and serotyping results. The overall consistency<br />

between the genotyping and serotyping assays was 90.19%<br />

(864/958). The specificity and sensitivity were 99.74% and<br />

88.77% for genotype 1, 96.97% and 93.97% for genotype<br />

2, 100% and 100% for mixed genotype 1/2, and 94.15%<br />

and 89.61% for non-genotype 1 or 2. In 190 SVR patients, no<br />

statistical difference existed in the serotype results after therapy(P=0.64).<br />

In the 326 follow-up patients, there was no difference<br />

in the proportion of genotypes between the groups with or<br />

without SVC(P=0.23, Tab 1). Conclusions The novel serotyping<br />

assay is sensitive and specific compared with the molecular<br />

genotyping assays. Using this assay, genotype distribution in<br />

SVC patients can also be revealed, which is similar with that<br />

in CHC patients. Genotype might not play an important role<br />

in SVC.<br />

Characteristics of the individuals with and without SVC<br />

Disclosures:<br />

Lai Wei - Advisory Committees or Review Panels: Gilead, AbbVie; Grant/<br />

Research Support: BMS<br />

The following authors have nothing to disclose: Ruifeng Yang, Huiying Rao


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1097A<br />

1821<br />

Mathematical model for early detection of hepatocellular<br />

carcinoma on top of HCV infection<br />

Abdelrahman Zekri, Amira S. Youssef, Yasser M. Bakr, Reham<br />

Mohamed Gabr, Ola Ahmed, Mohamed Abouelhoda; Virology<br />

and immunology Unit, Cancer Biology, National Cancer Institute,<br />

Cairo, Egypt<br />

Background: We investigated whether combinations of serum<br />

Intercellular Adhesion Molecules (sICAM-1), Beta catenin<br />

(β-catenin), Interleukin 8 (IL-8), Proteasome and soluble Tumor<br />

Necrosis Factor Receptor II (sTNF-RII) or subsets of them with<br />

AFP, used with logistic disease predictor models, could facilitate<br />

the early detection of hepatocellular carcinoma (HCC).<br />

Methods: Serum levels of IL-8, sICAM-1, sTNF-RII, Proteasome<br />

and β-catenin were measured in 479 subjects (192<br />

HCC associated with HCV infection; 96 HCV related liver<br />

cirrhosis (LC); 96 chronic hepatitis C (CHC) and 95 healthy<br />

controls). The R package and different modules for binary<br />

and multi-class classifiers based on generalized linear models<br />

were used to model the data. ROC curve analysis was<br />

used to find the best cutoffs differentiating among different<br />

groups. Results: We revealed mathematical models, based on<br />

binary classifier, made up of unique panel of biomarkers that<br />

improved the individual performance of AFP for the discrimination<br />

of HCC from benign liver disease. For the discrimination<br />

of HCC group from LC group, using a mathematical model<br />

[-1.133(E+01) +7.380* Proteasome + (1.081E-03)*sICAM-1+<br />

(2.574E-01)*β-catenin + (1.597E-02)*AFP] with cutoff 0.6552<br />

has achieved 98.8% specificity and 89.1% sensitivity. For the<br />

discrimination of HCC group from CHC group, a mathematical<br />

model [-1.040*(e+01) + (1.416e+00)*Proteasome +<br />

(2.024e-03)*IL-8+ (4.096e-03)*sICAM-1+ (4.251e-04)*sT-<br />

NF-RII+ (2.567e-01)*β-catenin + (2.442e-02)*AFP] with cutoff<br />

0.744 has achieved 96.8% specificity and 89.7% sensitivity.<br />

Also, we have derived an algorithm for resolving the multi-class<br />

classification problem by using three successive mathematical<br />

models. First, use the mathematical model [-1.10E+01+6.83*<br />

Proteasome + (1.29E-03)*sTNF—RII + (2.83E-01)*AFP]<br />

with cutoff 0.764 to discriminate healthy controls from disease<br />

cases. For disease cases, discriminate HCC cases from<br />

HCV cases (cases with LC and CHC) using the mathematical<br />

model [-1.191(E+01) + (3.222E+00)* Proteasome +<br />

(3.813E-04)*IL-8 + (1.518E-03)*sICAM-1 + (6.481E-04)*sT-<br />

NF-RII + (2.906E-01)* β-Catenin + (1.931e-02)*AFP] with<br />

cutoff 0.712. If the cases were HCV, proceed to discriminate<br />

LC cases from CHC cases using the mathematical model<br />

[-1.016e+01 + (-4.487e+00)*Proteasome + (2.086e-03)*sI-<br />

CAM-1 + (1.858e-03)*sTNF-RII+ (-2.984e-01)* β-Catenin +<br />

(2.169e-02)*AFP] with cutoff 0.582. Conclusion: Our proposed<br />

mathematical models can differentiate not only between<br />

HCC and other studied groups but also between all studied<br />

groups in non-invasive, inexpensive and rapid manner. Key<br />

words: Hepatocellular carcinoma, AFP, sICAM-1, β-catenin,<br />

IL-8, sTNF-RII, Proteasome.<br />

Disclosures:<br />

The following authors have nothing to disclose: Abdelrahman Zekri, Amira<br />

S. Youssef, Yasser M. Bakr, Reham Mohamed Gabr, Ola Ahmed, Mohamed<br />

Abouelhoda<br />

1822<br />

Hepatitis C Seropositivity is not as high as Western<br />

Countries in the Population Born Between 1945-1965<br />

According to Turkey Nationwide Hepatitis Prevalence<br />

Study (TURKHEP 2010)<br />

Osman C. Ozdogan 1 , Nurdan Tozun 2 , Sabahattin Kaymakoglu 3 ,<br />

Ramazan Idilman 4 , Zeki Karasu 5 , Ulus S. Akarca 5 ; 1 Gastroenterology,<br />

Marmara University Medical School, Istanbul, Turkey;<br />

2 Gastroenterology, Acibadem University Medical School, Istanbul,<br />

Turkey; 3 Gastroenterology, Istanbul University Medical Faculty,<br />

Istanbul, Turkey; 4 Gastroenterology, Ankara Universit Faculty of<br />

Medicine, Istanbul, Turkey; 5 Gastroenterology, Ege University<br />

Medical Faculty, Istanbul, Turkey<br />

Aim&Methods: Hepatitis C infection is more prevalent in the<br />

“Baby Boomer Population” those born between 1945 and<br />

1965 in USA and Western Countries. Although Turkey was<br />

not involved in World War II, birth rates were increased after<br />

1945 in Turkey. However, this population can not be defined<br />

as “Baby Boomers” since they have different properties. We<br />

aimed to determine the anti-HCV seropositivity in the population<br />

born between the years of 1945-1965 from the data<br />

of Turkey Nationwide Hepatitis Prevalence Study (TURKHEP)<br />

which was presented in AASLD in 2010 by Turkish Association<br />

for the Study of Liver Disease (TASL). A total of 5533 volunteers<br />

were screened in TURHEP Study, performed via home<br />

visits in 2009-2010 with volunteers living in urban and rural<br />

areas of 23 cities located in EUROSTAT NUTS 2 regions of<br />

Turkey which were determined based on two stage stratified<br />

sampling method. The participants were selected according<br />

to randomized sampling method of Statistics Institute of Turkey<br />

(TUIK). Participants’ socio-demographic features, anthropometrics,<br />

medical history, concomitant medications and presence<br />

of risk factors were recorded and blood samples were<br />

collected for the analysis of serum HBV, HDV, HCV markers,<br />

ALT and AST levels. Anti-HCV seropositivity and HCV RNA<br />

status was recruited from the TURKHEP data and compared<br />

with the population born after 1965. Results: TURKHEP data<br />

analysis included 5460 participants with data available on<br />

seroprevalence and risk factors for hepatitis C infection inwhich<br />

1980 participants (1037 female/943 male) borned between<br />

the years of 1945-1965 were recruited. Overall, 0.95% of the<br />

subjects were reactive for anti-HCV antibodies in whole population<br />

whereas Anti-HCV seropositive subjects were 28 out of<br />

1980 subjects in the population born between 1945-1965 at<br />

a rate of 1.4 % which is 1.5 times higher than the whole population.<br />

Half of these subjects were HCV RNA positive by the<br />

Real time PCR instrument, Rotor-Gene Q (QIAGEN, Germany).<br />

However among the subjects borned after 1965, Anti-HCV<br />

seropositive subjects were 23 out of 3460 subjects at a rate of<br />

0.7%. Conclusion: Eventhough the HCV seropositivity rate was<br />

higher than the whole Turkish population in Turkish population<br />

born between the years of 1945-1965, this rate of increase did<br />

not reach the higher rates of Western Baby Boomer Population<br />

showing the different transmission routes of HCV infection in<br />

Turkey.<br />

Disclosures:<br />

Ulus S. Akarca - Advisory Committees or Review Panels: GILEAD, BMS, MSD,<br />

AbbVie<br />

The following authors have nothing to disclose: Osman C. Ozdogan, Nurdan<br />

Tozun, Sabahattin Kaymakoglu, Ramazan Idilman, Zeki Karasu


1098A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1823<br />

Age and Risk Factor-based Serologic Screening for Hepatitis<br />

C Virus among an Urban High-risk Population<br />

Rositsa B. Dimova 2,3 , Anthony D. Martinez 2,1 , Ethan Weinberg 4 ,<br />

Ann Drobnik 5 , Andrew H. Talal 2,1 ; 1 Division of Gastroenterology,<br />

Dept of Medicine, SUNY-Buffalo, Buffalo, NY; 2 Center for Clinical<br />

Care and Research in Liver Disease, University at Buffalo, Buffalo,<br />

NY; 3 Department of Biostatistics, University at Buffalo, Buffalo,<br />

NY; 4 Department of Medicine, Weill Cornell Medical College,<br />

New York, NY; 5 New York City Department of Health and Mental<br />

Hygeine, New York, NY<br />

Background and aims: Hepatitis C virus (HCV) screening<br />

among individuals born between 1945 and 1965 (i.e. birth<br />

cohort) has been proposed to augment risk factor-based screening.<br />

Among an urban, high-risk population, we compared the<br />

frequency of HCV seropositivity among injection drug users<br />

(IDUs) and birth cohort members. Methods: We assessed HCV<br />

seropositivity in 7722 subjects from New York City drawn from<br />

HIV prevention programs, syringe exchange programs, drug<br />

treatment programs, and community health centers. HCV serology<br />

was assessed by immunoblot and data analysis employed<br />

logistic regression. Results: Mean age was 40.7 years, standard<br />

deviation (SD) 10.9 years, 47.4% were born between<br />

1945-1965, 81.2% were born in the US, 70.3% were male,<br />

53% were African American, 21.5% were Caucasian, 36.8%<br />

reported history of injection drug use, and 9.8% were HIV<br />

infected. Among men, 12.4% were men who had sex with<br />

men (MSM). Most participants were from HIV (72.7%) and substance<br />

use (56.4%) treatment sites. Overall, 26.6% were HCV<br />

seropositive, 55.8% of whom were born between 1945 and<br />

1965, and 82.2% had ever injected drugs. Among all participants,<br />

HCV seropositivity was significantly higher among IDUs<br />

than non-IDUs (60.5% versus 7.7%, odds ratio (OR) = 18.5,<br />

95% confidence interval (CI) [16.2, 21.1], p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1099A<br />

1825<br />

Hepatitis C Virus (HCV) Eradication in Patients with<br />

Varying Severity of Cirrhosis: Impact on Portal Hypertensive<br />

Complications, Liver Transplant (LT) and Death<br />

Varun Saxena 1 , Lisa M. Nyberg 2 , Aditi Dasgupta 1 , Stephanie Straley<br />

1 , Lisa Catalli 1 , Anders H. Nyberg 2 , Norah Terrault 1 ; 1 Gastroenterology,<br />

University of California San Francisco, San Francisco,<br />

CA; 2 Kaiser Permanente Southern California, San Diego, CA<br />

Background: Recent all-oral antiviral therapy for HCV results in<br />

high rates of sustained virologic response (SVR) among patients<br />

with varying severity of cirrhosis. The impact of SVR on longer-term<br />

hepatic outcomes (beyond SVR12) in patients with<br />

cirrhosis, especially those who are Child-Pugh (CP) B/C, is<br />

currently unknown. Aims: We aimed to compare the liver-related<br />

events in a cohort of HCV-infected patients with varying<br />

severity of cirrhosis who achieved SVR with all oral antiviral<br />

therapy vs. those not achieving SVR / untreated matched controls.<br />

Methods: Adults with HCV genotype 1 cirrhosis treated<br />

with simeprevir (SMV) + sofosbuvir (SOF) ± ribavirin (RBV)<br />

were followed for a median of 0.80 yrs (IQR: 0.65-0.93) after<br />

treatment discontinuation. Randomly selected, untreated controls<br />

(1-3/pt) matched on site, age ± 5yrs, CP (A vs. B/C) and<br />

model for end-stage liver disease (MELD) score ± 2 were followed<br />

for a similar duration. Safety outcomes included change<br />

in MELD and CP scores, number of emergency room (ER) visits<br />

and hospitalizations, presence of portal hypertensive complications<br />

(new/worsening ascites, varices, encephalopathy and/or<br />

spontaneous bacterial peritonitis), need for LT and death. Survival<br />

from treatment discontinuation was compared in treated<br />

patients and matched controls over equivalent timeframe.<br />

Results: Of 371 patients, 93 were treated with SMV+SOF±RBV<br />

and 278 were matched, untreated controls. Cohort median<br />

age was 62yrs (IQR: 58-68), 32% female, 25% Hispanic, 5%<br />

Black, 24% diabetes, median (range) platelet count 133K/<br />

mm 3 (20-341), albumin 3.5g/dL (2.3-4.7), MELD 8 (6-18),<br />

35% CP-B/C, 22% ascites, 22% encephalopathy, 8% varices.<br />

Of the SMV+SOF±RBV treated patients, 78 (84%) achieved<br />

SVR12. Liver-related outcomes between groups are shown in<br />

the Table. In bivariate Cox survival analysis, achievement of<br />

SVR vs. non-SVR/untreated (HR 0.15, 95% CI 0.02-0.97) and<br />

CP-A vs. CP-B/C (HR: 0.57, 95% CI 0.28-0.96) protected from<br />

LT/death. Conclusions: Achieving HCV cure among patients<br />

with compensated and decompensated cirrhosis significantly<br />

reduced the frequency of liver-related outcomes including portal<br />

hypertensive complications and LT/death.<br />

Liver-Related Outcomes in Patients Achieving SVR vs. Non-SVR/<br />

Untreated Matched Controls<br />

*unadjusted log-rank<br />

Disclosures:<br />

Lisa M. Nyberg - Grant/Research Support: Merck, Gilead, Abbvie<br />

Lisa Catalli - Advisory Committees or Review Panels: Gilead, Kadmon<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

The following authors have nothing to disclose: Varun Saxena, Aditi Dasgupta,<br />

Stephanie Straley, Anders H. Nyberg<br />

1826<br />

Can HCV Antigen Testing Replace HCV RNA PCR Analysis,<br />

for Diagnosis and Monitoring of Treatment for<br />

Hepatitis C?<br />

Andrew M. Hill 1 , Teri Roberts 2 , Valerianna Amorosa 3 , Kamron<br />

Pourmand 3 , Gail Matthews 4 , Graham Cooke 5 , Harun Khan 5 ,<br />

Janice Main 5 , Ashley Brown 5 , Joy A. Peter 9 , David R. Nelson 6 ,<br />

Michael W. Fried 7 , Jennifer Cohn 8 , Camilla S. Graham 10 ; 1 Pharmacology<br />

and Therapeutics, Liverpool University, Liverpool, United<br />

Kingdom; 2 Treatment Access Campaign, Medecins Sans Frontieres,<br />

Geneva, Switzerland; 3 Philadelphia VAMC, University of<br />

Pennsylvania, PA; 4 University of New South Wales, Kirby Institute,<br />

Sydney, NSW, Australia; 5 Faculty of Medicine, Imperial College,<br />

London, United Kingdom; 6 Clinical and Translational Science Institute,<br />

University of Florida, Gainesville, FL; 7 Liver Center, University<br />

of North Carolina, Chapel Hill, NC; 8 Division of Infectious Diseases,<br />

University of Pennsylvania, Philadelphia, PA; 9 Hepatology<br />

Research, University of Florida, Gainsville, FL; 10 Beth Israel Deaconess<br />

Medical Centre, Boston, MA<br />

Background: Simplified systems of diagnostic testing for hepatitis<br />

C virus (HCV) could make widespread implementation of<br />

Direct Acting Antiviral (DAA) based treatment more feasible in<br />

low and middle income countries. HCV core antigen tests (e.g.<br />

Abbott Architect assay) are cheaper and simpler to perform,<br />

but have lower limits of detection in the range of 1000 IU/mL.<br />

More expensive, complex HCV RNA nucleic acid assays (e.g.<br />

Roche TAQMAN) can detect HCV virus at levels of 15-25 IU/<br />

mL. Methods: A systematic review identified <strong>studies</strong> testing samples<br />

in duplicate for HCV antigen and HCV RNA. Clinical data<br />

from 4 cohort <strong>studies</strong> evaluating patients relapsing on PEG-<br />

IFN or DAA based treatment for acute and chronic HCV were<br />

reviewed to determine the percentage of relapsing patients<br />

with HCV RNA levels above 1000 IU/mL before treatment and<br />

at End of Treatment (EOT)+12-48 week time points. Results: In<br />

24 <strong>studies</strong> evaluating 8142 samples tested in parallel by the<br />

Abbott Architect antigen and HCV RNA assays, 46 samples<br />

(0.6%) were HCV antigen positive but HCV RNA undetectable.<br />

There were 280 samples (3.5%) HCV antigen negative but<br />

HCV RNA detectable: these samples generally showed HCV<br />

RNA results 1000 IU/mL in 31 (100%) before treatment, 28 (87%) at<br />

EOT+12, 29 (94%) at EOT+24, and 31 (100%) at EOT+48.<br />

In a Philadelphia chronic HCV treatment cohort, among 90<br />

relapsing patients HCV RNA levels were >1000 IU/mL in<br />

90 (100%) before treatment, and 89 (99%) at EOT+. In an<br />

Australian acute/early chronic HCV treatment cohort, among<br />

13 relapsing patients, HCV RNA levels were >1000 IU/ml in<br />

12 (92%) before treatment, and 9 (69%) at EOT+12 (n=11)<br />

or EOT+24 (n=2). In the HIV-TARGET observational cphprt<br />

study, all 232 relapsing patients had HCV RNA levels >1000<br />

IU/mL before treatment. There were 92 relapser patients with<br />

HCV RNA data available at least 10 weeks after EOT: 89/92<br />

(97%) had HCV RNA >1000 IU/mL at first relapse. Summary:<br />

In the 4 cohort <strong>studies</strong>, 365/366 patients (99.7%) had HCV<br />

RNA levels >1000 IU/mL before treatment; 214/224 relapsing<br />

patients (96%) had HCV RNA levels above 1000 IU/mL<br />

12 weeks after EOT, and so could have been detected by the<br />

Abbott Architect assay. A minority of patients relapsed with<br />

HCV RNA>1000 IU/mL 24-48 weeks after EOT. Antigen testing<br />

before and after treatment could potentially replace HCV<br />

RNA PCR analysis, but further validation <strong>studies</strong> are required in<br />

acute and chronic infection. New, low-cost Point-of-Care HCV<br />

antigen assays are needed for use in low income countries,<br />

with lower detection limits in the range of 1000 IU/mL.


1100A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Andrew M. Hill - Consulting: Janssen<br />

David R. Nelson - Advisory Committees or Review Panels: Merck; Grant/Research<br />

Support: Abbot, BMS, Beohringer Ingelheim, Gilead, Genentech, Merck, Bayer,<br />

Idenix, Vertex, Jansen<br />

Michael W. Fried - Consulting: Merck, Abbvie, Janssen, Bristol Myers Squibb,<br />

Gilead; Grant/Research Support: Merck, AbbVie, Janssen, Bristol Myers Squibb,<br />

Gilead; Patent Held/Filed: HCCPlex<br />

The following authors have nothing to disclose: Teri Roberts, Valerianna Amorosa,<br />

Kamron Pourmand, Gail Matthews, Graham Cooke, Harun Khan, Janice<br />

Main, Ashley Brown, Joy A. Peter, Jennifer Cohn, Camilla S. Graham<br />

FRY graph<br />

1827<br />

Comparative analysis of online patient education<br />

resources relating to Hepatitis C<br />

Rishabh Gulati, Mohammad Nawaz, Sowjanya Kanna, Nikolaos<br />

Pyrsopoulos; Medicine, Rutgers New Jersey Medical School, Newark,<br />

NJ<br />

Introduction: Role of the Internet is ever-increasing in the present<br />

era as the first source of medical information. Imprecise,<br />

partial comprehension of textual information limits its efficacy<br />

in communicating the disease process to the patient. Here, we<br />

report a comparative analysis of readability of patient-centered<br />

text pertaining to Hepatitis C available online. Methods: In<br />

March 2015, patient-centered information from websites of<br />

American College of Gastroenterology (ACG), Centers for Disease<br />

Control & Prevention (CDC), American Liver Foundation<br />

(ALF), Mayo Clinic, National Institutes of Health (NIH), Uptodate,<br />

HCVAdvocate & WebMD was downloaded & processed<br />

in Microsoft Word. All data were formatted & categorized<br />

into subsections. Proper nouns, copyright information & certain<br />

medical terms were omitted to limit bias. Text was then analyzed<br />

for their specific level of readability using 7 quantitative<br />

scales: Flesch Reading Ease, Flesch–Kincaid level, Gunning fog<br />

index, SMOG, Coleman-Liau, FRY & New Dale–Chall using<br />

Readability Studio software. Results: Modified documents had<br />

a mean grade level that was 1 less than their original counterparts.<br />

ACG had the highest mean grade level of readability<br />

of it’s content (13.4±0.61), with the lowest being for WebMD<br />

(8.4±0.40). When compared with all subsets, the treatment<br />

subsection had the highest mean grade level (12.2±0.85).<br />

ANOVA analysis showed that there were significant differences<br />

in the grade level depending on the source website (p <<br />

0.05), however there was no significant impact by subsection<br />

when compared with all readability tests. Post hoc Turkey HSD<br />

Analysis showed ACG was written at a significantly higher<br />

grade level than other websites. The treatment section was usually<br />

the most difficult section written when compared with other<br />

subsections (p < 0.05). Conclusion: Patient material is above<br />

the recommended 6 th grade level across all websites. Treatment<br />

section is often the most difficult section to comprehend.<br />

Greater emphasis on clear & simple language is warranted to<br />

increase quality & comprehension of online patient education<br />

resources.<br />

Disclosures:<br />

Nikolaos Pyrsopoulos - Advisory Committees or Review Panels: GILEAD, BMS,<br />

ABBVIE, VITAL THERAPIES, Quest; Grant/Research Support: ABBVIE<br />

The following authors have nothing to disclose: Rishabh Gulati, Mohammad<br />

Nawaz, Sowjanya Kanna<br />

1828<br />

Benefits of Antiviral Treatment and Predictors of Subsequent<br />

Hepatocellular Carcinoma Risk in Chronic Hepatitis<br />

C Patients with or without Sustained Virological<br />

Response to Peg-interferon plus Ribavirin Therapy<br />

Mei-Hsuan Lee 1 , Sheng-Nan Lu 2 , Ming-Lung Yu 3 , Cheng-Yuan<br />

Peng 4 , I-Shyan Sheen 5 , Chen-Hua Liu 6 , Jia-Horng Kao 6 , Wan-Long<br />

Chuang 3 , Hwai-I Yang 7 , Gilbert L’Italien 8 , Yong Yuan 9 , Chien-Jen<br />

Chen 7 ; 1 National Yang-Ming University, Institute of Clinical Medicine,<br />

Taipei, Taiwan; 2 Division of Hepatogastroenterology, Department<br />

of Internal Medicine, Kaohsiung Chang-Gung Memorial<br />

Hospital and Chang Gung University College of Medicine, Kaohsiung,<br />

Taiwan; 3 Hepatobiliary Division, Department of Internal Medicine<br />

and Hepatitis Center, Kaohsiung Medical University Hospital<br />

and Kaohsiung Medical University college of Medicine, Kaohsiung,<br />

Taiwan; 4 Division of Hepatogastroenterology, Department of<br />

Internal Medicine, China Medical University Hospital, Taichung,<br />

Taiwan; 5 Division of Hepatology, Department of Gastroenterology<br />

and Hepatology, Linkou Medical Center, Chang Gung Memorial<br />

Hospital, Tao-Yuan, Taiwan; 6 Graduate Institute of Clinical Medicine,<br />

Department of Internal Medicine and Hepatitis Research Center,<br />

National Taiwan University College of Medicine and Hospital,<br />

Taipei, Taiwan; 7 Genomics Research Center, Academia Sinica,<br />

Taipei, Taiwan; 8 Yale University School of Medicine, New Haven,<br />

CT; 9 Global Health Economics and Outcome Research, Bristol-Myers<br />

Squibb, Princeton, NJ<br />

Background & Aims: The aims of this large follow-up study<br />

were to evaluate the antiviral treatment efficacy in chronic hepatitis<br />

C patients by two distinctive cohorts, and to investigate<br />

the predictability of post-treatment seromarkers for HCC risk<br />

stratified by patients with sustained virological response (SVR)<br />

and non-SVR to antiviral treatment. Methods: The treatment<br />

cohort included 3,133 chronic hepatitis C patients treated with<br />

peg-interferon plus ribavirin (PR) for 6-12 months. In addition,<br />

the untreated cohort consisted of 1,366 anti-HCV seropositives<br />

who participated in a community-based liver cancer screening.<br />

All of them were seronegative for HBsAg and free of HCC<br />

at study entry. The treated patients were followed from the<br />

date starting PR therapy whereas the untreated patients were<br />

followed from the date of enrollment. Both cohorts were followed<br />

to the date of HCC diagnosis, death, or the end of<br />

2012, whichever came first. The Cox’s proportional hazards<br />

model was utilized to estimate the adjusted hazard ratio (HR)<br />

and 95% confidence interval (CI) associated with HCC by<br />

comparing patients with or without antiviral treatment. Among


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1101A<br />

the treated patients, the post-treatment seromarkers were evaluated<br />

for HCC risk by patients with or without SVR. Results:<br />

The patients in the treatment cohort had increased likelihood<br />

of advanced age, male gender, HCV genotype 1 infection,<br />

elevated serum levels of HCV RNA, AST, ALT, platelet count,<br />

and alfa-fetoprotein (AFP) at baseline (p


1102A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ratio for models including behavioural risk heterogeneity. Conclusions:<br />

Preferential mixing by HIV status and behavioural risk<br />

heterogeneity could be the main factors that have resulted in<br />

much higher HCV prevalence amongst HIV-positive MSM than<br />

in HIV-negative MSM. Greater HCV transmission could occur<br />

amongst HIV-negative MSM if HIV-negative MSM start using<br />

condoms less or mix more with HIV-positive MSM, something<br />

that may occur with widespread use of HIV pre-exposure prophylaxis.<br />

The effects of new HIV prevention interventions on<br />

HCV transmission should be considered.<br />

Disclosures:<br />

The following authors have nothing to disclose: Louis W. MacGregor, Peter Vickerman,<br />

Natasha K. Martin, Ford Hickson, Peter Weatherburn<br />

1831<br />

Positive Impact of Point of Care Testing and an Informational<br />

Brochure on Screening of the HCV Birth Cohort in<br />

a Rural New England Tertiary Care Center<br />

Arvind Suguness 1 , Casey M. Kolb Nava 1 , Kristen Ray 2 , Rolland C.<br />

Dickson 2 ; 1 Internal Medicine, Dartmouth Hitchcock Medical Center,<br />

White River Junction, VT; 2 Gastroenterology and Hepatology,<br />

Dartmouth Hitchcock Medical Center, Lebanon, NH<br />

Background. Hepatitis C Virus (HCV) screening of the Birth<br />

Cohort (BC) has been recommended by the CDC and USPSTF<br />

due to the high prevalence of HCV and advanced liver disease<br />

in this population. The use of risk-based screening was<br />

previously shown to be ineffective at identifying patients with<br />

HCV infection. To address these guidelines, we implemented<br />

a point-of-care testing (POCT) strategy for HCV, and surveyed<br />

patients within the cohort. The aim of this study was to assess<br />

factors in a HCV testing algorithm that influenced screening.<br />

Methods. A HCV BC screening initiative was fully implemented<br />

in an Internal Medicine Residents Clinic in a rural tertiary care<br />

center in the NE United States on 4-1-2014. BC patients (1945-<br />

1965) were identified by a prompt in the electronic medical<br />

record prior to arrival to clinic. They were given a brochure on<br />

HCV natural history, POCT, and HCV treatment at the time of<br />

check in. When patients were roomed they were offered testing<br />

with the Orasure finger prick test. The patients were given a<br />

questionnaire while they waited for the provider. HCV screening<br />

(+ or -) results were available within 20 minutes during the<br />

provider visit. Patients were informed of the results and those<br />

with positive results were offered same day PCR testing, with<br />

genotype performed if viral load was present. Results. There<br />

were 325 total respondents to the survey. Sixty three (19.4%)<br />

reported being asked by a health care provider to be tested<br />

for HCV in the past. Twenty one reported prior HCV testing<br />

(6.5%) and 8 reported a prior positive test (2.5%). Of the total<br />

respondents, 86 (26%) opted out of screening entirely, while<br />

239 patients (73.5%) chose to be screened at that visit. Of<br />

those screened, 117 (49.0%) said the information in the brochure<br />

or having test results immediately available influenced<br />

their decision to be tested. At least one reported risk factor<br />

was present in 142 patients (43.69%). Patients with at least<br />

one risk factor were no more likely to desire screening than<br />

those without risk factors (73.2% vs 73.8%, p=0.914). Patients<br />

with risk factors were no more likely than those without risk<br />

factors to report that the brochure or the immediately available<br />

results influenced their decision to get tested (43.2% vs 53.3%,<br />

p=0.123). Summary and Conclusion. Prior to initiation of a<br />

POCT strategy, relatively few patients in the birth cohort at<br />

our institution had been approached by health care providers<br />

about HCV screening and fewer still had been tested. The presence<br />

of an informational brochure, combined with point-of-care<br />

testing, significantly increased the number of patients with and<br />

without risk factors to undergo testing.<br />

Disclosures:<br />

Casey M. Kolb Nava - Consulting: AbbVie<br />

Kristen Ray - Advisory Committees or Review Panels: Abbvie<br />

Rolland C. Dickson - Advisory Committees or Review Panels: Biotest; Speaking<br />

and Teaching: gilead<br />

The following authors have nothing to disclose: Arvind Suguness<br />

1832<br />

Stability and prevalence of the NS3 Q80K polymorphism<br />

over time within HCV genotype 1a infected<br />

patients in Canada<br />

Jeffrey B. Joy 1 , Weiyan B. Dong 1 , Celia B. Chui 1 , Chanson J.<br />

Brumme 1 , Art Poon 1,3 , Huong Hew 2 , Richard Harrigan 1,3 ; 1 BC<br />

Centre for Excellence in HIV/AIDS, Vancouver, BC, Canada;<br />

2 Clinical Affairs, Janssen Incorporated, North York, ON, Canada;<br />

3 Medicine, University of British Columbia, Vancouver, BC, Canada<br />

Background: HCV genotype 1a (GT1a) infections harbouring a<br />

baseline Q80K polymorphism in the NS3 gene display reduced<br />

virologic response to IFN-based HCV treatments containing<br />

simeprevir. The prevalence of this and other NS3 resistance<br />

mutations in Canadian populations is not well described. Further,<br />

in the context of individual infections, the stability of this<br />

polymorphism among untreated patients over time is unknown.<br />

Methods: Using plasma samples serially collected over a<br />

10-year period from 121 HCV treatment naïve GT1a infected<br />

seroconverting injection drug users, we sought to investigate<br />

gain or loss of the Q80K polymorphism over time. A 1200-<br />

or 564-bp HCV NS3 fragment was sequenced by Sanger<br />

methods. Each sample was HLA typed to confirm database<br />

annotations on source individuals. HCV sequences were multiply<br />

aligned using MAFFT v7.154b. Phylogenetic trees were<br />

inferred using an approximate maximum likelihood method<br />

(FastTree2) and rooted under a molecular clock model using<br />

Path-O-Gen. To establish the prevalence of Q80K in Canadian<br />

Provinces, samples from GT1 LiPA HCV+ individuals were<br />

examined. Sequences were also screened for mutations associated<br />

with boceprevir/telaprevir resistance. Results: No patients<br />

whose first and last samples formed a monophyletic group<br />

altered their Q80K status. Nine patients changed genotypes<br />

(6 GT3a to GT1a, 2 GT1a to GT3a, and 1 GT1b to GT1a).<br />

Furthermore, in 10 patients, GT1a infections did not form a<br />

monophyletic group. Both between genotype and within genotype<br />

changes in viral lineage between collection dates suggest<br />

either (1) clearance followed by reinfection with a new variant<br />

or (2) a mixed infection. In sum, we observed 9 changes in<br />

Q80K in 121 patients, but in every case this resulted from<br />

patients switching HCV lineages rather than a mutation in<br />

their original HCV lineage. Prevalence of Q80K in Canada<br />

was 956/1835 (52%) in GT1a and provinces displayed no<br />

significant difference in prevalence of Q80K. Mutations associated<br />

with boceprevir or telaprevir resistance in this newly<br />

diagnosed population were also observed. Conclusions: These<br />

results suggest that, in the absence of therapy, Q80K is highly<br />

stable within HCV lineages and does not evolve in response<br />

to immune or other host specific effects. Observed changes<br />

in infection status amongst these patients supports genotypic<br />

and resistance testing of patients prior to starting IFN-based<br />

therapy, particularly amongst those at high risk of exposure to<br />

new variants of HCV such as intravenous drug users.<br />

Disclosures:<br />

Huong Hew - Employment: Janssen<br />

Richard Harrigan - Consulting: ViiV Health Care, Tobira Therapeutics, Selah<br />

Genomics Inc, Quest Diagnostics; Stock Shareholder: Merck, Gilead


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1103A<br />

The following authors have nothing to disclose: Jeffrey B. Joy, Weiyan B. Dong,<br />

Celia B. Chui, Chanson J. Brumme, Art Poon<br />

1833<br />

Performance of HCV Ag quantification as a screening<br />

tool in HCV mono-infected, HBV-HCV and HIV-HCV<br />

co-infected patients from Cameroon: the ANRS 12336<br />

study<br />

Léa Duchesne 1 , Richard Njouom 2 , Frédéric S. Lissock 2 , Ghislaine<br />

Flore Tamko-Mella 2 , Sandrine Rallier 3 , Lila Poiteau 3 , Alexandre<br />

Soulier 3 , Stéphane Chevaliez 3 , Guy Vernet 2 , Nicolas Rouveau 4 ,<br />

Pierre-Marie Girard 1 , Karine Lacombe 1 ; 1 Inserm UMR-S1136,<br />

Paris, France; 2 Pasteur Center of Cameroon, Yaounde, Cameroon;<br />

3 National Reference Center for Viral Hepatitis B, C and Delta,<br />

Department of Virology, Hôpital Henri Mondor, Créteil, France;<br />

4 National Agency of Research on AIDS and viral hepatitis, Paris,<br />

France<br />

Background: HCV chronic infection diagnosis currently relies<br />

on anti-HCV antibody (HCV-Ab) detection. As it cannot differentiate<br />

an active infection from a resolved one, its diagnosis<br />

must be confirmed by HCV-RNA measurement which<br />

is scarcely available in resource-limited settings. Quantifying<br />

HCV core antigen (HCV-cAg), a marker of viral replication,<br />

could shorten this algorithm if used as a one-step tool. Aim: To<br />

assess the performance of the HCV-cAg quantification for the<br />

diagnostic of chronic HCV in a serum bank of HCV mono- and<br />

HCV-HBV or HCV-HIV co-infected patients from Cameroon and<br />

the influence of co-infections on the test’s results. Methods: The<br />

quantification of the HCV-cAg was performed by an automated<br />

assay (Abbott Diagnostics) in 476 HCV-Ab negative samples<br />

and 548 HCV-Ab and HCV-ARN positive samples (n=1024)<br />

collected in patients from the Pasteur Center of Cameroon. Its<br />

performance was assessed by calculating its sensitivity and<br />

specificity, and building ROC curves in order to compare its<br />

results to the gold standard (ELISA and/or PCR) with the Area<br />

Under the Curve (AUC). Results: Among these 1024 sera, 493<br />

were HCV mono-infected, 27 of 40 HIV-infected were co-infected<br />

with HCV and 28 of 47 HBV –infected were HCV-HBV.<br />

No statistical association was found between the HCV-cAg<br />

level and our covariates (age, gender, HBV or HIV co-infection).<br />

The correlation between HCV-cAg and HCV ARN was<br />

good in the mono-infected group (r=0.75, p


1104A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

care clinic to screen persons born between 1945-1965 for<br />

HCV antibodies, and linking of viremic individuals to care.<br />

Emory University and Morehouse School of Medicine residents<br />

practicing in the Grady Memorial Hospital Primary Care Center<br />

(Atlanta, GA, USA) received one-on-one training regarding<br />

birth cohort screening, as well as an electronic medical record<br />

reminder prompt. Residents screened primary care patients,<br />

and project staff followed up positive HCV Ab tests and performed<br />

outreach and linkage to care. HCV Ab positive persons<br />

who attended an appointment at the Grady Primary Care<br />

Center, Grady Liver Clinic, or Infectious Disease Clinic during<br />

which the positive HCV test was addressed and further work-up<br />

ordered were counted as linked to care. Results: In a birth<br />

cohort population that was 92.5% Black/African American,<br />

412 (7.9%) of 5,239 patients screened for hepatitis C had<br />

HCV antibodies. HCV RNA testing was completed for 90%<br />

of the seropositive patients, and 70% were viremic. Of 258<br />

patients with a positive RNA test, 229 (89%) were referred to<br />

care, and 212 (82%) referred patients attended a linkage visit.<br />

Conclusions: The TILT-C screening program was found to be<br />

feasible and effective. The program was successful in detecting<br />

previously undiagnosed chronic hepatitis C infections and linking<br />

persons to care in an urban primary care center.<br />

TILTC: HCV Testing and Linkage Cascade<br />

Disclosures:<br />

Lesley Miller - Advisory Committees or Review Panels: Bristol Myers Squibb<br />

Anne C. Spaulding - Grant/Research Support: Gilead Sciences, BMS<br />

The following authors have nothing to disclose: Brandi Park, Nyiramugisha Niyibizi,<br />

April D. Elam, Francois Rollin, Shelly-Ann Fluker<br />

1836<br />

WITHDRAWN<br />

1837<br />

Cannabinoid receptor 2 (CB2) 63 RR variant is associated<br />

with immune-mediated disorders in patients with<br />

chronic HCV infection<br />

Nicola Coppola 1 , Rosa Zampino 2 , Giulia Bellini 3 , Maria Stanzione<br />

4 , Nicolina Capoluongo 1 , Aldo Marrone 2 , Margherita<br />

Macera 1 , Adriana Boemio 2 , Sabatino Maione 3 , Luigi Adinolfi 2 ,<br />

Emanuele Miraglia del Giudice 5 , Evangelista Sagnelli 1 , Francesca<br />

Rossi 5 ; 1 Mental Health and Public Medicine, Second University of<br />

Naples, Naples, Italy; 2 Internal Medicine and Hepatology, Second<br />

University of Naples, Naples, Italy; 3 Department of Experimental<br />

Medicine, Second University of Naples, Naples, Italy; 4 Department<br />

of Clinical and Experimental Medicine and Surgery, Second<br />

University of Naples, Naples, Italy; 5 Department of Woman, Child<br />

and of General and Specialized Surgery, Second University of<br />

Naples, Naples, Italy<br />

Background: Patients with HCV chronic infection frequently<br />

show immune-mediated disorders (IMDs). The Cannabinoid<br />

(CB) receptor 2, predominantly expressed in the immune cells,<br />

plays an important role on the function of the immune system.<br />

In particular, the CB2-63 variants (rs35761398) affects the<br />

ability of the CB2 receptor to exert its inhibitory function on T<br />

lymphocyte. Aims: to evaluate whether CB2 variants are associated<br />

with the presence of IMDs in patients with chronic HCV<br />

infection. Methods: Considering that nearly 30% of anti-HCV<br />

positive patients are affected by IMDs, we planned a 12-month<br />

recruitment period for treatment-naïve anti-HCV positive patients<br />

with signs of IMD and a 4-month period for treatment-naïve<br />

anti-HCV patients lacking these signs. The enrollment stared<br />

in September 2013 and at the end of the recruitment periods<br />

168 patients have been selected, 81 anti-HCV/HCV-RNA positive<br />

with signs of IMDs and 87 anti-HCV/HCV-RNA positive<br />

with no sign of IMDs. The presence of IMDs was defined by<br />

at least one of the following conditions: ANA positivity (titers<br />

≥1:160) observed in 22 (27.2%) cases, SMA positivity (titers<br />

≥1:160) in 3 (3.7%), a cryocrite >2% in 24 (29.6%), history<br />

or active autoimmune thyroiditis in 25 (30.9%), psoriasis in 4<br />

(4.9%), B-cells non-Hodgkin lymphoma in 2 (2.5%) and autoimmune<br />

hemolytic anemia in 1 (1.2%) case; no patient showed<br />

signs of lichen planus nor Syogren syndrome. All patients were<br />

screened for the CNR2 rs35761398 SNP by a TaqMan Assay<br />

Results: Compared with the 87 patients lacking IMDs, the 81<br />

in the IMDs group more frequently were females (65% vs 45%,<br />

p=0.01), but not other significant difference was found in initial<br />

demographic, epidemiologic, serological, biochemical and<br />

virological data. In particular, the age (mean+SD: 53±14.1 vs.<br />

52.9±13.4 years), ALT serum levels, HCV viral load and in distribution<br />

of HCV genotypes were similar in these two groups.<br />

Instead, the prevalence of the patients with the CB2-63 RR variant<br />

was significantly higher in patients in the IMD group than<br />

in those in the non-IMD group (49.4% vs 24.1%, p=0.001). A<br />

logistic regression analysis including the CB2-63 receptor (RR<br />

vs QR or QQ), age and sex, identified the CB2-63 RR as the<br />

only independent predictor of IMDs (p =0.005). Conclusions:<br />

the data suggest a significant previously unknown association<br />

between CB2-63 RR variant and IMDs in anti-HCV patients, an<br />

observation deserving further investigation on a larger series of<br />

patients to define its clinical value<br />

Disclosures:<br />

The following authors have nothing to disclose: Nicola Coppola, Rosa Zampino,<br />

Giulia Bellini, Maria Stanzione, Nicolina Capoluongo, Aldo Marrone, Margherita<br />

Macera, Adriana Boemio, Sabatino Maione, Luigi Adinolfi, Emanuele<br />

Miraglia del Giudice, Evangelista Sagnelli, Francesca Rossi<br />

1838<br />

Serum levels of Wisteria floribunda agglutinin–positive<br />

human Mac-2 binding protein are useful for evaluating<br />

early liver fibrosis in hepatitis C patients<br />

Kohei Oda 1 , Hirofumi Uto 1,2 , Seiichi Mawatari 1 , Rie Ibusuki 1 , Sho<br />

Ijuin 1 , Hiroka Onishi 1 , Haruka Sakae 1 , Kaori Muromachi 1 , Akihiko<br />

Oshige 1 , Kotaro Kumagai 1 , Tsutomu Tamai 1 , Akihiro Moriuchi<br />

3 , Akio Ido 1 ; 1 Digestive and Lifestyle Diseases, Department of<br />

Human and Environmental Sciences, Kagoshima University Graduate<br />

School of Medical and Dental Sciences, Kagoshima, Japan;<br />

2 Center for Digestive and Liver Disease, Miyazaki Medical Center<br />

Hospital, Miyazaki, Japan; 3 Department of HGF Tissue Repair and<br />

Regenerative Medicine, Kagoshima University Graduate School of<br />

Medical and Dental Sciences, Kagoshima, Japan<br />

Objective: Liver fibrosis is the most important risk factor for<br />

liver cancer, and the rate of liver carcinogenesis rises significantly<br />

in cases of mild to advanced liver fibrosis (stage F2<br />

or greater) in patients with hepatitis C. Therefore, it is very<br />

important to evaluate liver fibrosis precisely, but appropriate<br />

biomarkers for evaluating less-advanced liver fibrosis have


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1105A<br />

not been identified. Wisteria floribunda agglutinin–positive<br />

human Mac-2 binding protein (WFA(+)-M2BP) was reported<br />

to be a novel serum marker of liver fibrosis identified in glycoproteomic<br />

biomarker screening <strong>studies</strong>, and was effective for<br />

evaluating advanced liver fibrosis. However, its usefulness in<br />

less-advanced liver fibrosis has not been fully elucidated. In this<br />

study, we examined the utility of WFA(+)-M2BP as liver fibrosis<br />

marker in patients with less-advanced liver fibrosis. Methods:<br />

One hundred forty-three patients with biopsy-proven chronic<br />

hepatitis C were enrolled in this study. We examined the association<br />

between WFA(+)-M2BP and the results of blood biochemistry<br />

and liver histology examinations. Results: Fifty-nine<br />

patients were male, and the mean age of the entire group<br />

was 60 years. The mean platelet count was 16.6×10 4 /mL,<br />

serum ALT was 48 IU/L, AFP was 4.9 ng/dL, and liver fibrosis<br />

scores of F0–1/F2/F3/F4, respectively, were 80/42/14/7.<br />

WFA(+)-M2BP rose with liver fibrosis progression, was significantly<br />

correlated with liver fibrosis markers and scores, and<br />

was significantly associated with histopathology findings of<br />

liver fibrosis. For cases of less-advanced liver fibrosis (stage<br />

F2 or less), WFA(+)-M2BP was significantly elevated by a<br />

liver fibrosis extension case. Receiver operating characteristic<br />

(ROC) analysis revealed that the WFA(+)-M2BP cut-off value of<br />

1.94 [C.O.I.] discriminated between stages F2 and F0–1 (area<br />

under the ROC curve, 0.768) better than other serum markers<br />

such as platelet count (0.613), hyaluronic acid (0.661), type<br />

IV collagen (0.627), and procollagen III peptide (0.610), and<br />

better than liver fibrosis scores such as the FIB4 index (0.645)<br />

and APRI (0.660). In the multivariate analysis, high WFA(+)-<br />

M2BP was an independent factor associated with mild liver<br />

fibrosis (F2) (OR, 6.37; 95% CI, 1.85–22.00; P=0.003). Furthermore,<br />

in the 15 cases we could evaluate, WFA(+)-M2BP<br />

improved significantly following interferon treatment (from<br />

3.21 to 1.94;P=0.015). Conclusions: In chronic hepatitis C,<br />

WFA(+)-M2BP should prove useful in evaluating relatively early<br />

cases of liver fibrosis. In addition, WFA(+)-M2BP may serve as<br />

an indicator of improvement in liver fibrosis at an early stage.<br />

Disclosures:<br />

The following authors have nothing to disclose: Kohei Oda, Hirofumi Uto, Seiichi<br />

Mawatari, Rie Ibusuki, Sho Ijuin, Hiroka Onishi, Haruka Sakae, Kaori Muromachi,<br />

Akihiko Oshige, Kotaro Kumagai, Tsutomu Tamai, Akihiro Moriuchi, Akio<br />

Ido<br />

1839<br />

Factors Associated with Hepatitis C Virus Infection<br />

among Persons Ages 18 to 29 who Inject Drugs in Suburban<br />

and Rural Wisconsin<br />

David B. Rein 1 , Danielle Liffmann 1 , Jim Brunner 2 , Emily Wievel 2 ,<br />

Gregory Scott 3 , Daniel Raymond 4 , Lisa Danelski 2 , Scott Stokes 2 ,<br />

Joanne E. Brady 1 , Jon E. Zibbell 5 ; 1 NORC at the University of<br />

Chicago, Atlanta, GA; 2 AIDS Resource Center of Wisconsin,<br />

Green Bay, WI; 3 Sociology, DePaul University, Chicago, IL; 4 Harm<br />

Reduction Coalition, New York, NY; 5 Division of Viral Hepatitis,<br />

CDC, Atlanta, GA<br />

Background: Due to national increases in heroin and injection<br />

drug use, along with the recent HIV outbreak among HCV-infected<br />

persons who inject drugs (PWID) in Indiana, information<br />

on behavioral risk factors for HCV infection among young<br />

PWID in non-urban settings is needed. Methods: Between September<br />

2014 and May 2015, we collected survey data and<br />

hepatitis C antibody testing information from young adults ages<br />

18 to 29 who had injected drugs illicitly at least 1 time within<br />

the last year. Using a respondent driven snowball sampling<br />

design, the study collected HCV antibody status using rapid test<br />

kits donated by OraSure® as well as self-reported information<br />

on demographics, injecting behaviors, social network size,<br />

personal characteristics, and drug using histories. This initial<br />

analysis estimated descriptive statistics and stepwise multivariate<br />

logistic regressions (using an alpha of 0.05 for selection)<br />

of the risk of HCV infection. For selection, we included demographics,<br />

number of lifetime injections, history of ever sharing<br />

needles, filters, mix water, or cottons, seeing other people’s<br />

blood while injecting, reported size of injecting network, a<br />

history of ever injecting heroin, prescription opioids, cocaine,<br />

and methamphetamines, and scores on abbreviated scales of<br />

self-efficacy, extroversion, addiction, sensation seeking, and<br />

socio-normative aspirations. Results: Through May, 2015,<br />

we enrolled 265 persons: 58.4% male, 81.9% white, 14.7%<br />

American Indian and 3.4% of other races, with a mean age<br />

of 23.4. Of those, 34.0% were HCV+, 56.2% reported more<br />

than 100 lifetime injections, and 66.4%, 55.9%, 59.3%, and<br />

46.4% reporting ever sharing needles, filters, mix water, and<br />

cookers respectively. In multivariate analyses, injecting more<br />

than 100 times (OR 4.9; 95% CI 2.7-9.2) and having ever<br />

injected cocaine (OR 2.7; 95% CI 1.4-4.9) were significantly<br />

related to HCV+. Among those with less than 100 lifetime injections,<br />

sharing mix water (OR 4.6; 95% CI 1.3-16.1), injecting<br />

heroin (OR 9.5; 95% CI 1.0-89.3), and injecting cocaine<br />

(OR 4.3; 95% CI 1.2-15.3) were significantly associated with<br />

HCV+. Among persons who reported more than 1,000 lifetime<br />

injections, sharing filters (OR 3.1; 95% CI 1.3-7.5) and<br />

injecting cocaine (OR 3.7, 95% CI 1.3-10.0) were significantly<br />

associated with HCV+. Conclusions: We identified an HCV+<br />

prevalence 34.0% among rural/suburban PWID ages 18 to<br />

29. Injecting cocaine was significantly associated with HCV<br />

antibody positivity in the full population. Sharing injection<br />

equipment other than syringes was significantly associated with<br />

higher HCV+ risk when the sample was stratified by number of<br />

lifetime injections.<br />

Disclosures:<br />

David B. Rein - Grant/Research Support: Gilead Sciences, Inc.<br />

Daniel Raymond - Grant/Research Support: Gilead, Janssen<br />

The following authors have nothing to disclose: Danielle Liffmann, Jim Brunner,<br />

Emily Wievel, Gregory Scott, Lisa Danelski, Scott Stokes, Joanne E. Brady, Jon<br />

E. Zibbell<br />

1840<br />

Hyperbilirubinemia in HIV-HCV Co-infected Patients on<br />

cART - Drug Effect or Liver Disease Severity?<br />

Mathew Kaspar, Richard K. Sterling; Virginia Commonwealth University,<br />

Richmond, VA<br />

Background: Hyperbilirubinia (HB) is common and a known<br />

side effect of many medications used in the management of<br />

HIV, particularly the protease inhibitors (PI) Atazanavir and<br />

Indinavir. While typically benign, the sudden development of<br />

HB in HIV-HCV presents a unique challenge to determine if it<br />

is a benign drug effect, related to liver disease progression, or<br />

combination of the two. To date, there have been no <strong>studies</strong><br />

investigating the impact of underlying liver histology to distinguish<br />

drug effect from progression of liver disease in the<br />

HIV-HCV population. Objective: To determine if HB in the HIV-<br />

HCV pts being treated with PI is primarily due to drug effect,<br />

progression of liver disease, or both. Methods: We performed<br />

a retrospective analysis of 344 HIV-HCV pts undergoing liver<br />

biopsy. Pts with HBsAg +, hepatic decompensation, or HCC<br />

were excluded. Demographic, clinical, laboratory, and biopsy<br />

data were collected. The primary endpoint was HB (defined as<br />

serum bilirubin greater than 1.2mg/dL, the ULN). Univariate<br />

and multivariate analyses was used to determine factors associated<br />

with HB. Advanced fibrosis (AF) was defined by F3 or<br />

F4 histology (Knodell). Results: The mean age was 46 years,


1106A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

75% were male, 84% were black, 91% were GT 1, 19% had<br />

>5% steatosis, and 31% had AF. 155 pts (46%) were being<br />

treated with a PI: Atazanavir (n=40), Kaletra (n=40), Indinavir<br />

(n=15), Nelfinavir (n=34), Darunavir (n=6), and other (n=20).<br />

The prevalence of HB (range 1.3 to 9.4 mg/dL) was 15.4%.<br />

Those with HB had higher AST (132 vs 75; p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1107A<br />

of increasing the number of PWIDs treated with new oral DAAs<br />

was considered, including the annual number needed to treat<br />

in order to reduce the HCV-infected PWID population by 2030.<br />

Results: If the current transmission paradigm continues, there<br />

are projected to be 3,620 HCV infected PWIDs in 2030.<br />

Annually treating 40 HCV-infected PWIDs with new oral DAAs<br />

(1% of HCV-infected PWID population in 2014) resulted in a<br />

5% reduction in HCV-infected PWIDs by 2030, while annual<br />

treatment of 200 PWIDs (5% of 2014 population) resulted in a<br />

reduction of over 25% by 2030. Treating 387 PWIDs annually<br />

(17% of 2014 population) resulted in a >90% reduction in<br />

HCV-infected PWIDs by 2030. Targeting treatment to PWIDs<br />

engaged in OST and NEP would provide the greatest reduction<br />

in prevalence for the number of individuals treated (2.2 treated<br />

in OST/NEP to reduce prevalence by 1, as compared with 6.8<br />

treated in the general population). Conclusions: The results<br />

show that treating yet a small amount of PWIDs resulted in<br />

substantial decreases in the HCV infected PWID population by<br />

2030. Furthermore, the relative impact of treatment was greatest<br />

when focused on the population engaged in OST and NEP.<br />

Treatment is expected to increase the rate of reinfection; however,<br />

reinfection will decline as HCV prevalence decreases.<br />

This analysis supports the implementation of a screening and<br />

treatment strategy among PWIDs when combined with an<br />

expansion of harm reduction programs.<br />

Disclosures:<br />

Stefan Bourgeois - Advisory Committees or Review Panels: AbbVIe, Gilead; Consulting:<br />

Roche, MSD; Speaking and Teaching: Janssen, BMS<br />

Sarah Blach - Employment: Center for Disease Analysis<br />

Homie Razavi - Management Position: Center for Disease Analysis<br />

Geert Robaeys - Advisory Committees or Review Panels: Gilead; Speaking and<br />

Teaching: Merck, Janssens<br />

The following authors have nothing to disclose: Catharina Mathei, Christian<br />

Brixko, Jean-Pierre Mulkay, Thomas Sersté<br />

1843<br />

Non-Invasive Fibrosis Scores: Do They Predict Antiviral<br />

Treatment Response, Decompensation, Hepatocellular<br />

Carcinoma and Significant Liver Related Adverse Events<br />

in Chronic Hepatitis C?<br />

Ragesh B. Thandassery 1 , Madiha E. Soofi 1 , Anil John 1 , Samir S.<br />

Nairat 1 , Abdulrahman A. Alfadda 2 , Saad R. Al Kaabi 1 ; 1 Hamad<br />

Medical Corporation, Doha, Qatar; 2 King Faisal Specialist Hospital<br />

and Research Center, Riyadh, Saudi Arabia<br />

Background The role of non-invasive liver fibrosis scores (NIF)<br />

in predicting antiviral treatment (AVT) response and post treatment<br />

significant liver related events (SLRE) in chronic hepatitis<br />

C (CHC) is less studied. Aim To compare 12 simple NIF,<br />

derived from routine blood investigations, for predicting<br />

response to AVT and SLRE. Methods 1605 patients underwent<br />

liver biopsy (LB, Scheuer classification) and received AVT<br />

(pegylated interferon and ribavirin). 12 NIF [AST-platelet count<br />

ratio index (APRI), Fibrosis-4 (FIB-4) score, Lok score, GUCI<br />

score, Fibroalpha score, modified APRI, King score, AST-ALT<br />

ratio (AAR), Fibrosis Index (FI), Fibro score, Fibrosis cirrhosis<br />

index (FCI) and Globulin platelet index (GPI)] were calculated<br />

prior to AVT. AUROCs were calculated for each of these NIF<br />

for predicting non-response to AVT and development of SLRE<br />

(defined as development of any event requiring intervention;<br />

decompensation and hepatocellular carcinoma, HCC) on<br />

follow-up Results Mean age 41.9years, predominantly genotype<br />

4(65%).1089(67.8%) were responders, 482(30%) non<br />

responders and 34(2.1%) relapsers. After median follow-up of<br />

6580.5 patient-years; 52(3.2%) had decompensation (bleed-<br />

9, ascites-39, jaundice-22, hepatic encephalopathy-7, spontaneous<br />

bacterial peritonitis-10, hepatorenal syndrome-4),<br />

11(0.7%) had HCC and 60(3.7%) had SLRE. The predictive<br />

accuracy of NIF and LB for non-response to AVT was low.<br />

FibroQ, King score and FIB-4 showed high accuracy for predicting<br />

adverse events. For predicting decompensation, HCC<br />

and SLRE, FibroQ (0.88), King score (0.90) and FibroQ (0.87)<br />

had highest AUROC respectively. Conclusions Predictive accuracy<br />

of NIF for non-response to treatment was low. Some NIF<br />

have high accuracy for predicting development of decompensation,<br />

HCC and SLRE on follow-up. Application of these simple<br />

scores can improve assessment of liver prognosis in patients<br />

treated for CHC.<br />

Non-invasive scores in predicting post treatment events<br />

* AUROC(95% Confidence interval)<br />

Disclosures:<br />

The following authors have nothing to disclose: Ragesh B. Thandassery, Madiha<br />

E. Soofi, Anil John, Samir S. Nairat, Abdulrahman A. Alfadda, Saad R. Al Kaabi


1108A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1844<br />

Male gender, genotype 3, previous alcohol use,<br />

increased BMI, and diabetes are factors independently<br />

correlated to advanced HCV chronic liver disease in<br />

Italy: data from PITER (Piattaforma Italiana per lo studio<br />

della Terapia delle Epatiti viRali) cohort study<br />

Loreta A. Kondili 1 , Stefano Rosato 2 , Liliana E. Weimer 1 , Maria<br />

Giovanna Quaranta 1 , Loredana Falzano 1 , Alessandra Mallano<br />

1 , Maria Elena Tosti 2 , Maurizio Massella 1 , Maurizia R. Brunetto<br />

31 , Anna Linda Zignego 4 , Mario Rizzetto 29 , Alfredo Di Leo 24 ,<br />

Giovanni Raimondo 3 , Carlo Ferrari 25 , Antonio Craxi 30 , Gloria<br />

Taliani 5 , Pierluigi Blanc 6 , Antonio Gasbarrini 7 , Luchino Chessa 8 ,<br />

Elke M. Erne 9 , Giovanna Fattovich 10 , Pietro Andreone 11 , Maria<br />

Vinci 12 , Francesco P. Russo 14 , Erica Villa 13 , Giovanni B. Gaeta 15 ,<br />

Teresa A. Santantonio 19 , Guglielmo Borgia 20 , Gabriella Verucchi<br />

21 , Carmine Coppola 22 , Marcello Persico 17 , Liliana Chemello 28 ,<br />

Alfredo Alberti 14 , Vincenzo De Maria 26 , Massimo Puoti 27 , Raffaele<br />

Bruno 18 , Paolo Caraceni 11 , Massimo Andreoni 16 , Marco Marzioni<br />

23 , Stefano Vella 1 ; 1 Therapeutic Research and Medicines Evaluation,<br />

Istituto Superiore di Sanità, Rome, Italy; 2 Istituto Superiore<br />

di Sanità, Rome, Italy; 3 University of Messina, Messina, Italy; 4 University<br />

of Florence, Florence, Italy; 5 Sapienza University of Rome,<br />

Rome, Italy; 6 Santa Maria Annunziata Hospital, Florence, Italy;<br />

7 Policlinico Universitario Agostino Gemelli, Rome, Italy; 8 Azienda<br />

Ospedaliero-Univesitaria di Cagliari, Cagliari, Italy; 9 Azienda<br />

Ospedaliera di Padova, Padova, Italy; 10 Azienda Ospedaliero<br />

Universitaria Integrata di Verona, Verona, Italy; 11 Azienda<br />

Ospedaliero Universitaria di Bologna, Bologna, Italy; 12 Niguarda<br />

Hospital, Milano, Italy; 13 University of Modena, Modena, Italy;<br />

14 Azienda Ospedaliero Universitaria di Padova, Padova, Italy;<br />

15 Second University of Napoli, Napoli, Italy; 16 University Tor Vergata,<br />

Rome, Italy; 17 University of Salerno, Salerno, Italy; 18 University<br />

of Pavia, Pavia, Italy; 19 University of Foggia, Foggia, Italy;<br />

20 Federico II University of Naples, Naples, Italy; 21 Sant’Orsola<br />

Malpighi Hospital, Bologna, Italy; 22 Gragnano Hospital, Napoli,<br />

Italy; 23 Università Politecnica delle Marche, Ancona, Italy; 24 University<br />

of Bari, Bari, Italy; 25 Azienda Ospedaliero-Universitaria di<br />

Parma, Parma, Italy; 26 Azienda Ospedaliero-Universitaria Mater<br />

Domini, Catanzaro, Italy; 27 Azienda Ospedaliera Niguarda-Cà<br />

Granda, Milano, Italy; 28 University of Padua, Padova, Italy; 29 University<br />

of Torino, Torino, Italy; 30 University of Palermo, Palermo,<br />

Italy; 31 Azienda Ospedalero-Universitaria Pisana, Pisa, Italy<br />

Aim: Italy has the highest prevalence rates of HCV infection<br />

in Europe. In order to evaluate factors associated with severe<br />

liver disease in Italian patients in care, data derived from the<br />

PITER cohort study were analysed. Methods: The relationship<br />

between severe fibrosis stage/cirrhosis and sociodemographic<br />

characteristics, HCV RNA genotype, alcohol, body mass index<br />

(BMI), ALT, AST, GGT, platelets, diabetes, cardiovascular,<br />

neurological/psychiatric, autoimmune/reumatological and<br />

neoplastic diseases were evaluated by univariate and logistic<br />

regression statistical models. The regression model’s goodness<br />

of fit (calibration and sensitivity) was also estimated. Results: To<br />

date the cohort consists of 6831 consecutive HCV chronic liver<br />

disease patients (mean age 59±19 years; 3797 males) on clinical<br />

care. HCV genotype 1b (58%) and genotype 2 (15%) are<br />

significantly prevalent in older ages (older than 59 years) compared<br />

to genotypes 3 (10%) and 4 (7%), which are prevalent in<br />

younger ages. F4/cirrhosis stage was present in 2579 (38%)<br />

patients. It increased by age as expected, although 32% of<br />

patients younger than 60 years (3371 patients) had F4/cirrhosis.<br />

Of the enrolled patients, 48% were treatment-experienced.<br />

The independent role of each factor related with F4/cirrhosis,<br />

as defined by the logistic regression analysis, is as follows:<br />

male vs female (3797/3034) OR:1.42 (95% CI:1.25-01.66);<br />

BMI≥25 vs BMI


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1109A<br />

1845<br />

Effectiveness and cost-effectiveness of improvements in<br />

harm reduction interventions, a better cascade of care,<br />

and Treat as Prevention of chronic hepatitis C in people<br />

who inject drugs (PWID) in France (ANRS 95146)<br />

Anthony Cousien 1 , Viet Chi Tran 2 , Sylvie Deuffic-Burban 1,3 , Marie<br />

Jauffret-Roustide 4,5 , Guillaume Mabileau 1 , Jean-Stéphane Dhersin<br />

6 , Yazdan Yazdanpanah 1,7 ; 1 IAME, UMR 1137, INSERM - Univ<br />

Paris Diderot, Sorbonne Paris Cité, Paris, France; 2 Laboratoire Paul<br />

Painlevé UMR CNRS 8524, UFR de Mathématiques, Université des<br />

Sciences et Technologies Lille 1, Villeneuve d’Ascq, France; 3 LIR-<br />

IC-UMR995, INSERM - Université de Lille, Lille, France; 4 INSERM<br />

U988/UMR CNRS8211/Université Paris Descartes, Ecole des<br />

Hautes Etudes en Sciences Sociales, CERMES3: Centre de Recherche<br />

Médecine, Sciences, Santé, Santé Mentale et Société, Paris,<br />

France; 5 Institut de Veille Sanitaire, Saint-Maurice, France; 6 LAGA,<br />

CNRS, UMR 7539, Université Paris 13, Sorbonne Paris Cité, Villetaneuse,<br />

France; 7 Hôpital Bichat Claude Bernard, Service des<br />

Maladies Infectieuses et Tropicales, Paris, France<br />

We estimated the cost-effectiveness of strategies designed to<br />

improve harm reduction interventions and/or the cascade of<br />

care in PWID in the context of the incoming direct-acting antivirals<br />

(DAAs) in France. We used a dynamic model to simulate<br />

life expectancy in discounted quality adjusted life years<br />

(QALYs), direct lifetime discounted costs, incremental cost-effectiveness<br />

ratio (ICER) and the number of first generation new<br />

HCV infections for each strategy among PWID in Paris metropolitan<br />

area from 2015 until death: S1: base case=current<br />

practice. Time before access to needle and syringe programs<br />

(NSP) after injection initiation=1y, time before access to opioid<br />

substitution therapies (OST) when in NSP=0.5y, time to<br />

diagnosis after infection=1.25/1.45y, time to linkage to care<br />

after diagnosis=2.6y, loss to follow-up (LTFU) rate=14%/y,<br />

SVR rate=95%, treatment initiation: fibrosis ≥F2 S2: improved<br />

risk reduction interventions. Access to NSP after injection initiation=0.25y,<br />

time before access to OST when in NSP=0.25y<br />

S3: treatment initiation: fibrosis ≥F0 S4: improved cascade of<br />

care. Time to diagnosis=0.5y, time to linkage to care=0.5y,<br />

LTFU rate=5%/y, SVR rate=95% S5: S3&S4 S6: S2&S3&S4<br />

Results are presented in Table. Compared with the base case,<br />

improved cascade of care (S4) increased the average life<br />

expectancy with an ICER of 8,373€/QALY gained (


1110A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(HCV-2). Phylogenetic analysis revealed that all HCV-2 isolates<br />

from Suriname were part of 6 robust phylogenetic clusters containing<br />

HCV-2 sequences obtained from Surinamese living in<br />

The Netherlands. Conclusion HCV prevalence among the Surinamese<br />

population attending the ED was 1% and was highest<br />

among those of Javanese ethnicity. Older patients were more<br />

often infected, suggesting a long-standing infection acquired<br />

in the past. HCV-2 is the circulating genotype in Suriname and<br />

was likely introduced in Suriname by slave trade from western<br />

Africa in the past centuries. Targeted screening with linkage to<br />

care and prevention guidelines for identified at-risk groups are<br />

required, as well as implementation of stricter hygiene regulations<br />

for permanent make-up salons. Peginterferon and ribavirin<br />

are not yet routinely given as HCV treatment in Suriname,<br />

however given the predominance of favorable HCV genotype<br />

2, would be strongly advised.<br />

Disclosures:<br />

Richard Molenkamp - Independent Contractor: Roche Diagnostics<br />

Maria Prins - Speaking and Teaching: msd, roche<br />

The following authors have nothing to disclose: Sigrid MacDonald - Ottevanger,<br />

Stephen G. Vreden, Jannie van der Helm, Thijs J. van de Laar, Els Dams, Jimmy<br />

Roosblad, John F. Codrington<br />

1847<br />

Hepatocellular carcinoma (HCC) scoring system for<br />

the individualized prediction of liver cancer in 1080<br />

HCV-related compensated cirrhosis included in the<br />

French multicenter prospective cohort ANRS CO12 Cir-<br />

Vir.<br />

Nathalie Ganne 1 , Valerie Bourcier 1 , Richard Layese 2 , Nabila Talmat<br />

1 , Ventzislava Petrov-Sanchez 3 , Patrick Marcellin 4 , Dominique<br />

Guyader 5 , Stanislas Pol 6 , Dominique G. Larrey 7 , Victor de Ledinghen<br />

8 , Denis Ouzan 9 , Fabien Zoulim 10 , Pierre Nahon 1 , Françoise<br />

Roudot-Thoraval 2 ; 1 Hepatology, Jean Verdier hospital, Bondy,<br />

France; 2 Henri Mondor hospital, Creteil, France; 3 ANRS, Paris,<br />

France; 4 Beaujon hospital, Clichy, France; 5 CHU, Rennes, France;<br />

6 Cochin hospital, Paris, France; 7 CHU, Montpellier, France; 8 CHU,<br />

Bordeaux, France; 9 none, Saint Laurent du Var, France; 10 CHU,<br />

Lyon, France<br />

Background and aims: This study aimed to develop and validate<br />

a simple scoring system to refine individualized prediction<br />

of HCC risk in patients with HCV-related compensated<br />

cirrhosis included in the French prospective multicentre ANRS<br />

CO12 CirVir cohort (1) . Methods: Among 1323 patients with<br />

HCV-cirrhosis enrolled in the ANRS CO12 CirVir, 720 and<br />

360 patients were randomly assigned to the training and validation<br />

sets, respectively. Cox multivariate proportional hazards<br />

model was used to predict HCC occurrence in the population,<br />

then nomograms were computed to assess individualized risk.<br />

Results: During a mean follow-up of 51 months, 103 (14%) and<br />

39 (11%) patients developed HCC, respectively in the training<br />

and validation cohorts. According to Cox regression analysis,<br />

5 variables were independently associated with the occurrence<br />

of HCC in the training cohort (table 1). A 13-point risk score<br />

was derived in the training cohort using these 5 variables, the<br />

score of each variable being obtained by linear transformation<br />

of the coefficients and rounding. The population was stratified<br />

according to this scoring system into four groups, with HCC<br />

occurrence gradually increasing from 0% to 20% at 3 years,<br />

and 0% to 37% at 5 years, for patients with the lowest (score<br />

≤ 2) and highest (score ≥ 11) HCC risk, respectively and validated.<br />

The AUC for the 1- and 3-year prediction were 0.75<br />

[0.64; 0.85] and 0.75 [0.71;0.80] respectively. Derived from<br />

this score, a nomogram was also built, enabling individualized<br />

prediction of HCC occurrence at 1, 3 or 5 years. Conclusion:<br />

An HCC score constructed using 4 baseline variables<br />

(age, past excessive alcohol consumption, platelets count, GGT<br />

serum level) and qualitative HCV RNA at end-point is accurate<br />

to predict HCC at the individual level in French patients<br />

with HCV-cirrhosis, with or without SVR. External validation in<br />

non-French populations is warranted. (1) Trinchet et al. Complications<br />

and competing risks of death in compensated viral<br />

cirrhosis (ANRS CO12 CirVir prospective cohort). Hepatology.<br />

2015 Feb 11. doi: 10.1002/hep.27743.<br />

table 1<br />

Disclosures:<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Dominique Guyader - Advisory Committees or Review Panels: ROCHE, GILEAD,<br />

IRIS, ABBVIE; Board Membership: MERCK; Grant/Research Support: JANSSEN;<br />

Speaking and Teaching: BMS<br />

Stanislas Pol - Board Membership: Sanofi, Bristol-Myers-Squibb, Boehringer Ingelheim,<br />

Tibotec Janssen Cilag, Gilead, Glaxo Smith Kline, Roche, MSD, Novartis;<br />

Grant/Research Support: Glaxo Smith Kline, Gilead, Roche, MSD; Speaking and<br />

Teaching: Sanofi, Bristol-Myers-Squibb, Boehringer Ingelheim, Tibotec Janssen<br />

Cilag, Gilead, Glaxo Smith Kline, Roche, MSD, Novartis<br />

Dominique G. Larrey - Advisory Committees or Review Panels: BAYER, SANOFI,<br />

PFIZER, SERVIER-BIG, AEGERION, MMV, BIAL-QUINTILES, TEVA, ORION, NEG-<br />

MA-LERADS, ASTELLAS, ASTRAZENECA, DNDI, GSK, J AND J; Board Membership:<br />

BMS, GILEAD, ABBVIE, BMS, GILEAD, ITREAS, MMS; Grant/Research<br />

Support: GILEAD, MSD, BMS, ABBVIE, TIBOTEC/JANSSEN, BMS<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

Fabien Zoulim - Advisory Committees or Review Panels: Janssen, Gilead, Novira,<br />

Abbvie, Tekmyra, Transgene; Consulting: Roche; Grant/Research Support:<br />

Novartis, Gilead, Scynexis, Roche, Novira, Assemblypharm; Speaking and<br />

Teaching: Bristol Myers Squibb, Gilead<br />

Françoise Roudot-Thoraval - Advisory Committees or Review Panels: Roche; Consulting:<br />

LFB biomedicaments; Speaking and Teaching: gilead, Janssen, BMS,<br />

Roche<br />

The following authors have nothing to disclose: Nathalie Ganne, Valerie<br />

Bourcier, Richard Layese, Nabila Talmat, Ventzislava Petrov-Sanchez, Denis<br />

Ouzan, Pierre Nahon<br />

1848<br />

Differential Progression to Cirrhosis and Hepatic Decompensation<br />

in Asian Americans (AA) and Non-Asian (NA)<br />

Patients with Chronic Hepatitis C (CHC)<br />

An K. Le 1 , Nghia H. Nguyen 1,2 , Changqing Zhao 1,3 , Joseph K.<br />

Hoang 1 , Christine Y. Chang 1 , Mingjuan Jin 1,4 , Pauline Nguyen 1 ,<br />

Peter T. Nguyen 1,5 , Richard H. Le 1 , Michelle Q. Jin 1 , Linda<br />

Nguyen 6 , Lee Ann Yasukawa 7 , Jian Q. Zhang 8 , Susan C. Weber 7 ,<br />

Mindie H. Nguyen 1 ; 1 Division of Gastroenterology and Hepatology,<br />

Stanford University Medical Center, Palo Alto, CA; 2 Department<br />

of Medicine, University of California San Diego, San Diego,<br />

CA; 3 Department of Cirrhosis, Institute of Liver Disease, Shuguang<br />

Hospital, Shanghai, China; 4 Department of Epidemiology and Biostatistics,<br />

School of Public Health, Zhejiang University, Hangzhou,<br />

China; 5 School of Medicine, University of Texas Medical Branch,<br />

Galveston, TX; 6 Stanford University, Palo Alto, CA; 7 Center for<br />

Clinical Informatics, Stanford University School of Medicine, Palo<br />

Alto, CA; 8 Chinese Hospital, San Francisco, CA<br />

BACKGROUND: Little is known about the natural history of<br />

CHC in AA .Our goal is to evaluate the rate of progression<br />

to cirrhosis and hepatic decompensation in a large ethnically


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1111A<br />

diverse CHC cohort. METHODS: Using ICD-9 query and manual<br />

chart review for all identified patients, 5,942 consecutive<br />

adults with CHC and known ethnicity from 3 community clinics<br />

and one large U.S. university center were confirmed and<br />

included. RESULTS: AA were older (mean age=60.3±13.4 vs.<br />

55±10.2, p


1112A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1850<br />

Prevalence of hepatitis B core antibody (anti-HBc) and<br />

association with hepatocellular carcinoma (HCC) in<br />

patients with chronic hepatitis C virus (HCV) infection in<br />

the United States (US) and China<br />

Ming Yang 1 , Elizabeth Wu 2 , Sherry Fu 2 , Huiying Rao 1 , Bo Feng 1 ,<br />

Andy Lin 3 , Ran Fei 1 , Neehar D. Parikh 2 , Robert J. Fontana 2 , Anna<br />

S. Lok 2 , Lai Wei 1 ; 1 Peking University People’s Hospital, Peking<br />

University Hepatology Institute, Peking University Health Science<br />

Center, Beijing, China; 2 Division of Gastroenterology, University<br />

of Michigan Health System, Ann Arbor, MI; 3 The Molecular and<br />

Behavioral Neuroscience Institute, University of Michigan, Ann<br />

Arbor, MI<br />

Background: Evidence of previous hepatitis B virus (HBV) infection<br />

[i.e. detectable anti-HBc with undetectable hepatitis B<br />

surface antigen (HBsAg)] has been associated with increased<br />

risk of HCC in patients with chronic HCV infection in many<br />

Asian <strong>studies</strong> but this was not confirmed in some <strong>studies</strong> in the<br />

US. Aims: To compare the prevalence of anti-HBc between US<br />

and Chinese patients with chronic HCV and identify clinical<br />

correlates of anti-HBc positivity. Methods: Prospective study<br />

of 2 cohorts of chronic HCV patients at University of Michigan<br />

in Ann Arbor, US and Peking University People’s Hospital<br />

in urban Beijing and Gu’an and Kuancheng Clinics in rural<br />

Hebei, China. Results: A total of 1,833 patients were analyzed<br />

(963 US; 870 Chinese). 32.3% of US and 46.6% of<br />

Chinese HCV patients were anti-HBc+ and HBsAg- (p0.001).<br />

The prevalence of anti-HBc among patients with chronic hepatitis,<br />

cirrhosis, and HCC was 30.5%, 32.3% and 38.5% in<br />

the US (p=0.250) and 44.5%, 53.8% and 66.7% in China<br />

(p=0.020), respectively. Anti-HBc+ patients were older, had<br />

longer estimated duration of infection and were more likely to<br />

have a history of alcohol and tobacco use in both cohorts. US<br />

anti-HBc+ patients were more likely to be men and to report<br />

a history of injection drug use as the source of infection. In<br />

contrast, Chinese anti-HBc+ patients were more likely to report<br />

medical procedures as the source of infection. Multivariate<br />

analysis showed that there was no association between anti-<br />

HBc positivity and HCC in the US and Chinese cohorts and the<br />

combined cohort. Conclusions: Evidence of previous HBV infection<br />

among patients with chronic HCV is lower in the US cohort<br />

compared to the Chinese cohort. Although the prevalence of<br />

anti-HBc paralleled the severity of HCV-related liver disease,<br />

anti-HBc positivity was not an independent factor associated<br />

with risk of HCC in patients with chronic HCV in both US and<br />

Chinese cohorts.<br />

Characteristics of Patients with and without anti-HBc in the US<br />

and China<br />

Anna S. Lok - Advisory Committees or Review Panels: Gilead, MYR, Tekmira;<br />

Consulting: GSK, Merck; Grant/Research Support: AbbVie, BMS, Gilead, Idenix<br />

Lai Wei - Advisory Committees or Review Panels: Gilead, AbbVie; Grant/<br />

Research Support: BMS<br />

The following authors have nothing to disclose: Ming Yang, Elizabeth Wu, Sherry<br />

Fu, Huiying Rao, Bo Feng, Andy Lin, Ran Fei, Neehar D. Parikh<br />

1851<br />

Wisteria floribunda agglutinin-positive Mac-2-binding<br />

protein predicts hepatocellular carcinoma development<br />

in chronic hepatitis C patients who achieved sustained<br />

virological response to interferon-based anti-viral therapy<br />

Shunsuke Sato, Hironori Tsuzura, Takuya Genda, Akihito Nagahara;<br />

Gastroenterology and Hepatology, Juntendo University Shizuoka<br />

Hospital, Izunokuni, Japan<br />

Objective: The achievement of a sustained virological response<br />

(SVR) after anti-viral therapy reduces the incidence of hepatocellular<br />

carcinoma (HCC) development in patients with chronic<br />

hepatitis C. In advances of direct acting antivirals against hepatitis<br />

C virus, it is expected to drastically increase the patients<br />

with achievement of SVR. However, HCC development in<br />

patients who achieved SVR is not rarely observed. Liver fibrosis<br />

is known as a predictor of HCC development in patients with<br />

achievement of SVR. Although several noninvasive methods<br />

evaluating liver fibrosis has been reported, wisteria floribunda<br />

agglutinin (WFA)-positive Mac-2-binding protein (WFA + -M2BP)<br />

is recently developed as a noninvasive glycobiomarker of liver<br />

fibrosis. The aim of this study is to evaluate WFA + -M2BP as<br />

a predictive marker of HCC development in chronic hepatitis<br />

C patients with achievement of SVR. Methods: A total 280<br />

patients who achieved SVR by interferon-based anti-viral therapy<br />

and underwent measurement of serum WFA + -M2BP level<br />

were enrolled. We compared its usefulness as a predictive<br />

marker of HCC development to other risk factors including age,<br />

gender, body mass index, fibrosis stage, aspartate aminotransferase,<br />

alanine aminotransferase, gamma-glutamyl transpeptidase,<br />

and alpha-fetoprotein (AFP) by using multivariate Cox<br />

proportional hazard analysis. Cumulative incidences of HCC<br />

development were evaluated using Kaplan-Meier plot analysis<br />

and the log-rank test. Results: Serum WFA + -M2BP level was significantly<br />

correlated with liver fibrosis stage (p=0.001). During<br />

the median follow-up time of 4.0 years, 9 of the 280 patients<br />

developed HCC. Multivariate cox proportional hazard analysis<br />

revealed that WFA + -M2BP level was an independent risk factor<br />

of HCC development (hazard ratio (HR): 1.167, 95% confidence<br />

interval (CI): 1.015-1.342, p=0.030) as well as platelet<br />

counts (HR: 0.682, 95%CI: 0.538-0.864, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1113A<br />

1852<br />

HCV in Semen of HIV-infected Men During Acute and<br />

Chronic Infection<br />

Samuel Turner 1,4 , Marcus Yip 1,5 , Wouter van Seggelen 1,6 , Davey<br />

M. Smith 2,3 , Sara Gianella 2 , Daniel S. Fierer 1 ; 1 Infectious Diseases,<br />

Mount Sinai School of Medicine, New York, NY; 2 Infectious<br />

Diseases, University of California, San Diego, San Diego, CA;<br />

3 Infectious Diseases, Veterans Affairs Medical Center, San Diego,<br />

CA; 4 James Cook University, Cairns, QLD, Australia; 5 Monash University,<br />

Melbourne, VIC, Australia; 6 Amsterdam Medical Center,<br />

Amsterdam, Netherlands<br />

Introduction It remains unclear how hepatitis C virus (HCV) is<br />

transmitted in the epidemic of sexually transmitted HCV among<br />

HIV-infected men who have sex with men (MSM). Our previous<br />

epidemiological study in New York City strongly implicated<br />

semen, and seminal HCV has been detected intermittently and<br />

at low levels during chronic HCV, but little is known about<br />

seminal HCV during acute HCV infection. Methods HIV-infected<br />

MSM with acute and chronic HCV infection were prospectively<br />

enrolled. Three paired semen and blood specimens<br />

were collected at 2-week intervals. For men with acute HCV,<br />

all specimens were collected within 6 months of detection of<br />

new HCV infection. HCV viral load (VL) was quantified using<br />

qRT-PCR platforms (Abbott m2000 with lower limit of quantification<br />

[LLOQ] 12 IU/mL for seminal plasma, and Roche<br />

COBAS AMPLICOR with LLOQ 43 IU/mL for blood). Results<br />

Paired semen and blood specimens were obtained from 33<br />

HIV-infected MSM (21 acute, 12 chronic HCV). Overall, seminal<br />

HCV was detected in 16 (27%) of 59 specimens from<br />

11 (33%) men (Table). Comparing men with acute or chronic<br />

HCV, there were no differences in either the proportion of<br />

semen specimens with HCV detected (21% vs 38%, respectively;<br />

p=0.159), or in the median seminal HCV VL (1.32 vs<br />

1.77 log 10<br />

IU/mL, respectively; p=0.163). Median blood HCV<br />

VL was higher among men with detectable compared to not<br />

detectable seminal HCV (6.36 vs 5.47 log 10<br />

IU/mL, respectively;<br />

p=0.002). Among men with low serum VL (


1114A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1854<br />

Modeling the probability of hepatitis C virus transmission<br />

in injecting drug users as a function of viral load<br />

Qingwen Cui 1 , Alexander Gufraind 2,3 , Basmattee Boodram 3 ,<br />

Scott Cotler 2 , Harel Dahari 2,4 , Marian E. Major 1 ; 1 CBER/FDA,<br />

Alexandria, VA; 2 Department of Medicine, Loyola University, Maywood,<br />

IL; 3 Epidemiology & Biostatistics, University of Illinois at<br />

Chicago, Chicago, IL; 4 Theoretical Biology and Biophysics, Los<br />

Alamos National Laboratory, Los Alamos, NM<br />

Background: The major route of hepatitis C virus (HCV) transmission<br />

in the USA is through injecting drug use. It has been<br />

proposed that vaccines would not need to achieve sterilizing<br />

immunity in injecting drug user (IDU) populations in order to<br />

reduce disease spread, but this has not been tested empirically.<br />

We aimed to simulate HCV transmission through contaminated<br />

syringe sharing, apply this to a mathematical model to determine<br />

infection probabilities relative to HCV RNA and assess<br />

the impact of measures that reduce titers (e.g. vaccines or antivirals)<br />

in IDU populations. Method: HCV RNA positive (HCVpos)<br />

plasma was drawn into an insulin syringe and expelled.<br />

Negative plasma (NEG-pre) was drawn into the same syringe<br />

and expelled (NEG-post). Aliquots of HCV-pos, NEG-pre, and<br />

NEG-post were tested for RNA titer. This was repeated employing<br />

a water rinse of the syringe after expulsion of the HCV-pos<br />

sample. A binomial model was used to estimate the probability<br />

of infection with and without syringe rinse, assuming an RNA to<br />

HCV infectivity ratio of 12:1 [range=10:1-20:1]. Results: We<br />

found that 0.5% [range 0.05%-1.16%] of an HCV-pos sample<br />

could be transferred if a syringe is not rinsed or 0.07% [range<br />

0-0.18%] when rinsing is employed. Predicted probabilities<br />

of infection from syringe sharing as a function of RNA titer<br />

are shown in Fig.1. Risk of transmission increases 11-fold as<br />

titers (log10 IU/mL) rise from 2.6 to 4.0 (unrinsed) and 3.6 to<br />

5.0 (rinsed). Conclusions: IDUs experience a risk of transmission<br />

even with syringe rinsing. However, if titers are reduced<br />

below 2 log10 IU/mL transmission via syringe sharing could<br />

be prevented. Overall, a vaccine would not need to achieve<br />

sterilizing immunity to reduce disease spread in IDUs. Similarly,<br />

antiviral treatment could impact spread before sustained<br />

response is achieved.<br />

Fig1:Grey bands: Interquartile ranges based on Monte Carlo<br />

sensitivity analysis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Qingwen Cui, Alexander<br />

Gufraind, Basmattee Boodram, Scott Cotler, Harel Dahari, Marian E. Major<br />

1855<br />

Heterogeneous IL28B/IFNL4 distribution and association<br />

of the C/TT/T haplotype with spontaneous clearance<br />

and less liver damage in Mexican patients with chronic<br />

hepatitis C virus infection<br />

Karina Gonzalez-Aldaco 1,2 , João Renato R. Pinho 3 , Sonia<br />

Roman 1,2 , Ketti G. Oliveira 3 , Nora A Fierro 1,2 , Leticia Oyakawa 3 ,<br />

Rubia A. Santana 3 , Erika Martinez-Lopez 1,2 , Roberta Sitnik 3 ,<br />

Arturo Panduro 1,2 ; 1 University of Guadalajara, Jalisco, Mexico;<br />

2 Molecular Biology in Medicine, Civil Hospital of Guadalajara,<br />

Guadalajara, Mexico; 3 Hospital Israelita Albert Einstein, Sao<br />

Paulo, Brazil<br />

Background & aims. Recently, genetic polymorphisms of interleukin-28B<br />

(IL28B) (rs12979860 C allele and rs8099917 T<br />

allele) and interferon-lambda 4 (IFNL4) (ss469415590 TT<br />

allele) have been associated with spontaneous clearance (SC)<br />

of hepatitis C virus (HCV) infection among distinct populations<br />

worldwide. However, data regarding admixed and Amerindian<br />

populations from Latin America are lacking. This study aimed<br />

to analyze the genetic admixture of the West Mexico population<br />

based on the distribution of the IL28B (rs12979860C>T,<br />

rs8099917G>T) and IFNL4 (ss469415590∆G>TT) polymorphisms,<br />

and the association between the IL28B/IFNL4 haplotypes<br />

with SC and liver damage. Methods. In a cross-sectional<br />

study, 711 unrelated individuals from West Mexico, including<br />

Amerindians (86 Nahuas/95 Huicholes) and Mestizo populations<br />

(32 from Villa Purificación (VP), 172 from Guadalajara,<br />

Jalisco and 326 from Tepic, Nayarit) were genotyped. In a<br />

case-control study, 234 treatment-naïve HCV-infected Mestizo<br />

patients (149 with chronic hepatitis C (CHC) and 85 with SC)<br />

were included for the association of haplotypes with SC and<br />

liver damage. A Real-Time PCR assay perfomed genotyping,<br />

and transitional elastography (FibroScan ) staged liver damage.<br />

Genetic relatedness was evaluated by pairwise comparisons<br />

(exact tests). Both, genetic distances (Fst) and haplotypes<br />

were inferred by using Arlequin software (version 3.1). Linkage<br />

disequilibrium (LD) was calculated by GDA software (version<br />

1.0). All subjects signed an informed consent. The study protocol<br />

conformed to the Declaration of Helsinki and was approved<br />

by the Ethical Committee. Results. Three different clusters<br />

(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1115A<br />

1856<br />

Pragmatic Non-invasive test thresholds for the selection<br />

of patients with Hepatitis C cirrhosis for treatment with<br />

directly acting antiviral agents.<br />

William M. Rosenberg 1 , Julie Parkes 1,2 ; 1 Institute for Liver and<br />

Digestive Health, University College London, London, United Kingdom;<br />

2 Public Health Sciences and Medical Statistics, University of<br />

Southampton, Southampton, United Kingdom<br />

Background: Directly acting antiviral agents can cure HCV<br />

infection in most patients. Their high cost has led some healthcare<br />

providers to restrict use to patients with established cirrhosis.<br />

Non-invasive tests (NIT) for liver fibrosis are being used to<br />

stratify patients for liver fibrosis severity. However the selection<br />

of appropriate thresholds of NIT for the diagnosis of cirrhosis<br />

in this context is contentious. Methods: We took advantage of<br />

a consensus meeting of UK Hepatitis C experts to determine<br />

thresholds for NIT to diagnose cirrhosis. An appropriate sensitivity<br />

for the detection of cirrhosis was agreed by consensus.<br />

In addition a consensus threshold for fibroscan (11.5kPa) was<br />

selected. Using this threshold we reviewed published literature<br />

to identify test thresholds of commonly used NIT yielding a<br />

sensitivity for detecting cirrhosis in line with the experts’ range.<br />

Studies identified in the recent HTA Systematic Review of NIT<br />

for liver fibrosis (Corssan et al.) were considered. SPSS was<br />

used to analyse extracted data to determine the sensitivity for<br />

cirrhosis associated with a Fibroscan of 11.5±0.4kPa. The<br />

mean sensitivity for the diagnosis of cirrhosis associated with<br />

this Fibroscan reading was then used to identify the corresponding<br />

thresholds for APRI, FIB4 and ELF using data from published<br />

<strong>studies</strong> of patients with HCV infection where liver histology was<br />

used as the reference test for cirrhosis. Results: The expert consensus<br />

was that no more than 10-15% of patients with cirrhosis<br />

should be misdiagnosed as not having cirrhosis, equivalent to<br />

a test sensitivity for cirrhosis of 85-90%. Thirty three <strong>studies</strong><br />

reported fibroscan thresholds for cirrhosis ranging from 27kPa. Five reported a threshold of 11.5±0.4 and were<br />

included in the analysis, corresponding to a mean sensitivity of<br />

88.4%, a figure that lies within the range of sensitivities considered<br />

“acceptable” by the experts. The thresholds that yielded<br />

similar sensitivities for the diagnosis of cirrhosis were then identified<br />

for APRI, FIB4 and ELF and are presented in Table 1. Conclusions:<br />

UK CHC experts selected sensitivity for the detection<br />

of cirrhosis of 88% to allocate DAA treatment. Corresponding<br />

thresholds for widely used NIT have been identified.<br />

TABLE1: Thresholds for non-invasive tests for the identification of<br />

cirrhosis in HCV infection.<br />

Disclosures:<br />

William M. Rosenberg - Advisory Committees or Review Panels: Janssen, Merk,<br />

Gilead, Merk, Gilead, GSK; Board Membership: iQur Limited, iQur Limited;<br />

Consulting: siemens; Speaking and Teaching: siemens, Roche<br />

Julie Parkes - Stock Shareholder: iQur (spouse is shareholder)<br />

1857<br />

Illegal drug use disclosure among patients with chronic<br />

hepatitis C infection<br />

Kimberly Rhodes, Sherrie M. Harrell, Victor M. Cardenas, Andres<br />

Duarte-Rojo; Division of Gastroenterology and Hepatology, University<br />

of Arkansas for Medical Sciences, Little Rock, AR<br />

Background: Illegal or recreational drug use (IDU) is a major<br />

public health problem worldwide. It impacts the prevalence<br />

of hepatitis C (HCV) infection, reinfection rate, adherence to<br />

antiviral therapy, and rate of fibrosis progression. We aimed<br />

to investigate disclosure of IDU among HCV patients before<br />

and after an intervention aiming to improve response rate.<br />

Methods: Consecutive new HCV patients were evaluated at<br />

a tertiary center. Before their visit to clinic they completed a<br />

questionnaire on the use of any (A) IDU, intravenous (IV) IDU,<br />

snorted (S) IDU, cannabis (THC) use, and other (O) IDU. During<br />

clinic time patients were educated about the importance of<br />

disclosing an accurate substance abuse history, followed by<br />

a verbal reinterrogation on IDU. McNemar test was used for<br />

statistical analysis. Results: 126 patients (50±12 years, males<br />

50%) were included, and 122 (97%) answered the questionnaire.<br />

Among the 4 that did not answer, 3 had used IDU in the<br />

past (2 IV-IDU, 1 only THC). Frequencies of IDU per category<br />

were: A-IDU 84%, IV-IDU 67%, S-IDU 55%, THC 66%, O-IDU<br />

29%. The rate of patients that changed their statement on IDU<br />

before and after education is shown in Table. Revised IDU<br />

reporting was statistically significant for S-IDU (p=0.01), and<br />

THC (p


1116A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

tion of patients harboring a pre-treatment Y93H mutation. A<br />

total of 702 serum samples of Japanese HCV genotype 1b<br />

infected patients were screened in the study. The frequency of<br />

NS5A-Y93H was also determined by ultra-deep sequencing<br />

and compared by using 55 sera of patients. Possible relationships<br />

between clinical phenotypes and Y93H strains were<br />

investigated. Results: The overall success rate was 98.9%<br />

and the lower detection limit of HCV RNA level could be estimated<br />

to be almost 1 log order. The mutant strain (Y93H)<br />

was observed in 23.6% (164/694) patients. The lower detection<br />

limit of the proportion of Y93H strains could be estimated<br />

at 1 to 2 %. A significant positive correlation was observed<br />

between the proportion of Y93H strains determined by this<br />

assay system and the proportion determined by deep sequencing<br />

(r = 0.85, P 1% of<br />

total population). While prevalence is decreasing, the burden<br />

of HCV-related morbidity and mortality continues to grow. HCV<br />

transmission among PWIDs is the leading contributor to new<br />

infections with over half of infected cases reporting a history of<br />

injection drug use. The Serviço de Intervenção nos Comportamentos<br />

Aditivos e Dependências (SICAD) estimates that there<br />

were 14 426 (12 732 – 16 101) active PWIDs in Portugal in<br />

2012, and that 88% of PWIDs were HCV infected, equivalent<br />

to 12 695 infected PWIDs. In addition, SICAD estimates that<br />

16 401 individuals were enrolled in opioid substitution therapy<br />

(OST), and 950 652 syringes were distributed among<br />

PWIDs in 2013. Understanding HCV transmission dynamics<br />

among high-risk populations requires robust epidemiological<br />

data and country-specific mathematical modeling to assess<br />

the potential impact of improved HCV treatment strategies.<br />

Methods: HCV transmission was modeled using cohorts to<br />

track HCV incidence and prevalence among PWIDs in the<br />

general population, as well as active PWIDs enrolled in OST<br />

and/or needle exchange program (NEP). Model assumptions<br />

were derived from published literature and expert consensus.<br />

The relative impact of increasing treatment among PWID was<br />

considered, including the annual number needed to treat in<br />

order to substantially reduce the HCV-infected PWID population<br />

by 2030. Results: If the current transmission paradigm<br />

continues, HCV infected PWIDs are projected to decline 25%<br />

from 12 695 infected PWIDs in 2012 to 9320 HCV infected<br />

PWIDs in 2030. Annually treating 100 HCV-infected PWIDs<br />

(approximately 1% of HCV-infected PWID population in 2012)<br />

resulted in a 7% reduction in HCV-infected PWIDs by 2030,<br />

while annual treatment of 460 PWIDs (5% of 2012 population)<br />

resulted in a reduction of over 30% by 2030. Treating<br />

1400 PWIDs annually (15% of 2012 population) resulted in<br />

a 91% reduction in HCV-infected PWIDs by 2030; reductions<br />

were even greater among NEP participants (92% reduction),<br />

OST participants (92% reduction) and participants of both programs<br />

(95% reduction). Conclusions: The results show that<br />

treating a relatively small number of PWIDs resulted in substantial<br />

decreases in the HCV infected PWID population by<br />

2030. These data support implementation of a screening and<br />

treatment strategy among PWIDs; by focusing on groups with<br />

high transmission rates, the burden of HCV-related morbidity<br />

and mortality may be decreased.<br />

Disclosures:<br />

Homie Razavi - Management Position: Center for Disease Analysis<br />

Filipe Calinas - Advisory Committees or Review Panels: Merck Sharp & Dohme,<br />

Roche Pharmaceuticals, Gilead sciences, AbbVie, Janssen; Consulting: Boehringer<br />

Ingelheim; Speaking and Teaching: Bristol Myers Squibb, Gilead Sciences,<br />

Janssen; Stock Shareholder: Merck Sharpe<br />

Chris R. Estes - Employment: Center for Disease Analysis<br />

Rui T. Marinho - Advisory Committees or Review Panels: Abbvie, MSD, Roche,<br />

BMS, Janssen, Bayer<br />

The following authors have nothing to disclose: Domingos Duran, Jorge Félix,<br />

Fernando Maltez, Luís Mendão, Graca Vilar<br />

1860<br />

Comparison of On-Treatment HCV RNA Kinetics Using<br />

Two Generations of the COBAS Taqman Assay<br />

Tania M. Welzel 1 , Stanislas Pol 2 , Patrick Marcellin 3 , Robert H.<br />

Hyland 4 , Deyuan Jiang 4 , Phillip S. Pang 4 , Diana M. Brainard 4 , Vincent<br />

Leroy 5 , Stefan Zeuzem 1 , Marc Bourlière 6 ; 1 Johann Wolfgang<br />

Goethe University Medical Center, Frankfurt am Main, Germany;<br />

2 Department of Hepatology, Université Paris-René Descartes, Paris,<br />

France; 3 Hôpital Beaujon, Clichy, France; 4 Gilead Sciences, Inc,<br />

Foster City, CA; 5 CHU de Grenoble, Grenoble, France; 6 Hôpital<br />

Saint Joseph, Marseilles, France<br />

Background: The fully automated Roche COBAS® AmpliPrep/<br />

COBAS® Taqman® (version 2.0) HCV quantification test<br />

(CAP/CTM) with a lower limit of quantification (LLOQ) of 15


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1117A<br />

IU/mL is replacing the Roche COBAS® TaqMan® HCV test<br />

(version 2.0) for use with the High Pure System (CTM/HPS,<br />

LLOQ of 25 IU/mL) in clinical practice. Here we describe,<br />

through parallel testing, the HCV RNA viral decline observed<br />

in the SIRIUS study (NCT01965535) using the two different<br />

assays. Methods: Stored, frozen samples drawn during the first<br />

8 weeks of treatment from 154 patients in the SIRIUS trial were<br />

identified and analyzed in parallel using both assay platforms.<br />

These results were compared with each other and with the<br />

results initially obtained using the CTM/HPS (LLOQ=25) system<br />

at the time of the study conduct (between October 2013 and<br />

April 2014). Results: As previously reported, 97% of patients<br />

in this study achieved SVR12. 765 paired samples drawn<br />

during the first 8 weeks of treatment were analyzed. Compared<br />

to HCV RNA measurements obtained with the CTM/HPS<br />

(LLOQ=25) assay, a higher proportion of patients had quantifiable<br />

HCV RNA at treatment weeks 1, 2, and 4 using the new<br />

CAP/CTM (LLOQ=15) assay (see Table). Using the CAP/CTM<br />

(LLOQ=15) assay, twenty six subjects were >LLOQ at Week 4.<br />

Nevertheless, 25/26 (96%) of these subjects achieved SVR12<br />

after 12 or 24 weeks of treatment. Comparison with the data<br />

obtained at the time of study conduct will be presented. Conclusions:<br />

In this analysis, with the next generation Ampliprep sample<br />

preparation with the COBAS TaqMan® HCV test (version<br />

2.0), as many as 17% of patients had detectable HCV RNA<br />

at week 4; conversely, with the older High Pure System used<br />

to prepare samples, as few as 1% of patients had detectable<br />

HCV RNA at week 4. Notably, however, with either assay, the<br />

likelihood of a sustained viral response did not correlate with<br />

early virologic response. Thus, for LDV/SOF-based therapies,<br />

determination of on-treatment HCV RNA levels is not suitable to<br />

guide treatment decisions.<br />

Disclosures:<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

Stanislas Pol - Board Membership: Sanofi, Bristol-Myers-Squibb, Boehringer Ingelheim,<br />

Tibotec Janssen Cilag, Gilead, Glaxo Smith Kline, Roche, MSD, Novartis;<br />

Grant/Research Support: Glaxo Smith Kline, Gilead, Roche, MSD; Speaking and<br />

Teaching: Sanofi, Bristol-Myers-Squibb, Boehringer Ingelheim, Tibotec Janssen<br />

Cilag, Gilead, Glaxo Smith Kline, Roche, MSD, Novartis<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Robert H. Hyland - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Deyuan Jiang - Employment: Gilead Sciences<br />

Phillip S. Pang - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Diana M. Brainard - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Vincent Leroy - Board Membership: Abbvie, BMS, Gilead; Consulting: Janssen,<br />

MSD; Speaking and Teaching: Abbvie, BMS, Gilead, Janssen, MSD<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Marc Bourlière - Advisory Committees or Review Panels: Schering-Plough,<br />

Bohringer inghelmein, Schering-Plough, Bohringer inghelmein, Transgene; Board<br />

Membership: Bristol-Myers Squibb, Gilead, Idenix; Consulting: Roche, Novartis,<br />

Tibotec, Abott, glaxo smith kline, Merck, Bristol-Myers Squibb, Novartis, Tibotec,<br />

Abott, glaxo smith kline; Speaking and Teaching: Gilead, Roche, Merck, Bristol-Myers<br />

Squibb<br />

1861<br />

Expanded eligibility to HCV treatment with novel IFNfree<br />

therapies in a large real-world cohort: The German<br />

Hepatitis C Registry<br />

Dietrich Hueppe 2,6 , Stefan Mauss 2,3 , Klaus H. Boeker 2,13 , Andreas<br />

Schober 2,5 , Gerlinde Teuber 2,4 , Stefan Christensen 2,11 , Uwe<br />

Naumann 2,12 , Claus Niederau 2,10 , Stefan Zeuzem 2,8 , Michael P.<br />

Manns 1,2 , Thomas Berg 2,9 , Peter Buggisch 2,7 , Markus Cornberg 1,2 ,<br />

Christoph Sarrazin 2,8 , Heiner Wedemeyer 2,1 ; 1 Gastroenterology,Hepatology<br />

and Endocrinology, Hannover Medical School,<br />

Hannover, Germany; 2 German Hepatitis C Registry, Hannover,<br />

Germany; 3 MVZ Düsseldorf, Düsseldorf, Germany; 4 IFS Frankfurt,<br />

Frankfurt, Germany; 5 Hepatologische Praxis Göttingen, Göttingen,<br />

Germany; 6 Hepatologische SChwerpunktpraxis Herne, Herne,<br />

Germany; 7 IFI Hamburg, Hamburg, Germany; 8 University of Frankfurt,<br />

Frankfurt, Germany; 9 University of Leipzig, Leipzig, Germany;<br />

10 Hospital Oberhausen, Oberhausen, Germany; 11 CIM Münster,<br />

Münster, Germany; 12 Praxiszentrum Kaiserdamm, Berlin, Germany;<br />

13 Hepatologische Praxis Hannover, Hannover, Germany<br />

The German Hepatitis C Registry is a non-interventional prospective<br />

cohort study aiming to recruit at least 10.000 HCV-infected<br />

patients treated with novel direct acting antivirals<br />

against hepatitis C. Second generation DAAs have been used<br />

in Germany since January 2014 and each drug was available<br />

immediately after EMA approval. The current registry follows<br />

a previous national cohort which recruited more than 37.000<br />

HCV patients since 2002. The aim of this analysis was to investigate<br />

the evolution of patient profiles considered eligible for<br />

antiviral therapy over a period of 13 years with different treatment<br />

options available. Methods: Characteristics of patients<br />

enrolled in the German Hepatitis C registry after approval of<br />

2 nd generation DAAs since February 2014 (n=2247; period<br />

III) were compared with patients recruited 2002-2007 (n=<br />

19,115; period I) and patients recruited during 2011-2012<br />

when 1 st generation PIs were widely prescribed in Germany<br />

(n=1,768; period II). Results: The mean age of patients enrolled<br />

increased over the last 12 years (43.0 period I vs. 47.0 years<br />

period II vs. 51.0 years period III) while the gender distribution<br />

remained largely unchanged (male gender 58.9% vs. 64.8%<br />

vs. 61.2%). A slight shift in HCV genotype distribution became<br />

evident with less genotype 3 infections in period III (20.5%)<br />

than in period I (24.6%) but an increase in genotype 4 (2.9%<br />

vs. 2.5% vs. 5.1%). The proportion of patients with liver cirrhosis<br />

increased from 4.5% to 24.5% in period III while less<br />

patients were treatment naïve in period III compared to period<br />

I (56.3 vs. 86.3). In 2014/15 patient characteristics differed<br />

largely between individuals treated with IFN-based regimens<br />

(n=655) vs. IFN-free DAA combination therapies (n=1,169)<br />

regarding age (46.3 years vs. 54.4 years; patients >70 years<br />

3.1% vs. 9.3%), proportion of patients with liver cirrhosis<br />

(15.4% vs. 37.3%), percentage of patients with platelet counts<br />


1118A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Gerlinde Teuber - Advisory Committees or Review Panels: MSD, Gilead Sciences,<br />

BMS, Abbvie; Speaking and Teaching: MSD, Gilead Sciences, Janssen-Cilag,<br />

BMS, Abbvie<br />

Stefan Christensen - Advisory Committees or Review Panels: BMS, Abbvie, Janssen,<br />

ViiV, Gilead, MSD; Speaking and Teaching: Gilead, MSD, Abbvie, BMS,<br />

ViiV, Reckitt Benckiser, Janssen<br />

Uwe Naumann - Speaking and Teaching: MSD, Roche, BMS, Abbott, VIIV, Janssen,<br />

Boehringer Ingelheim, Gilead<br />

Claus Niederau - Advisory Committees or Review Panels: MSD, Gilead, BMS,<br />

Janssen, Abbvie; Grant/Research Support: MSD; Speaking and Teaching: MSD,<br />

Gilead, BMS, Janssen, Abbvie, Roche<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Peter Buggisch - Advisory Committees or Review Panels: Janssen, AbbVie, BMS;<br />

Speaking and Teaching: Roche, MSD, Gilead, Merz Pharma<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

Christoph Sarrazin - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Janssen, Merck/MSD, Gilead, Roche, Abbvie, Janssen, Merck/MSD;<br />

Consulting: Merck/MSD, Merck/MSD; Grant/Research Support: Abbott, Roche,<br />

Merck/MSD, Gilead, Janssen, Abbott, Roche, Merck/MSD, Qiagen; Speaking<br />

and Teaching: Gilead, Novartis, Abbott, Roche, Merck/MSD, Janssen, Siemens,<br />

Falk, Abbvie, Bristol-Myers Squibb, Achillion, Abbott, Roche, Merck/MSD, Janssen<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

The following authors have nothing to disclose: Andreas Schober<br />

1862<br />

Evaluation of liver and spleen transient elastography in<br />

chronic hepatitis C patients.<br />

Mohamed A. Elmazaly 1 , Ayman Alsebaey 1 , Maha M. Elsabaawy 1 ,<br />

El-Sayed S. Tharwa 1 , Hanaa M. Badran 1 , Nermine Ehsan 2 , Eman<br />

A. Rewisha 1 ; 1 Hepatology department, National Liver Institute,<br />

Shebin elkom menoufiya, Egypt; 2 Pathology, National Liver Institute,<br />

Shebin Elkoom, Egypt<br />

Background: Liver fibrosis is a common consequence of chronic<br />

HCV (CHC) infection. It may evolve to cirrhosis, hepatocellular<br />

carcinoma. Aim: is to evaluate the role of liver stiffness<br />

measurement (LSM), spleen stiffness measurement (SSM) and<br />

their combination (CLSM) using FibroScan TM in assessment of<br />

liver fibrosis in CHC patients. Methods: Four hundred twenty<br />

treatment naïve CHC patients and 40 healthy controls were<br />

included. Liver, renal function tests, CBC and INR were done.<br />

Liver biopsy was done for all of them except if contraindicated.<br />

Fibrosis was graded by Metavir score. Abdominal ultrasonography<br />

was done before the FibroScan TM and the liver<br />

biopsy. Transient elastography measurement was done using<br />

FibroScan TM in the supine position after 6-8 hours fasting. The<br />

patient were classified into mild fibrosis (F1-F2, n=248) and<br />

significant fibrosis (F3-F4, n=172) group. Results: There were<br />

statistically significant difference (p=0.001) between patients<br />

with mild fibrosis (F1-F2) versus patients with significant fibrosis<br />

(F3-F4) regarding; the age (35.06±8.63 vs. 43.71±7.97<br />

years), serum bilirubin (0.73±0.25 vs. 1.26±0.73 mg/dl),<br />

serum albumin (4.42±0.32 vs. 3.84±0.51 g/dl), platelets<br />

(206.81±50.55 vs. 140.50±53.77×10 3 /mL), platelets spleen<br />

ratio (1762.20±521.26 vs. 1014.64±470.27). Furthermore a<br />

statistical significant difference (p=0.001) was detected with<br />

LSM (6.57±2.62 vs. 23.04±12.15kPa), SSM (25.56±5.36 vs.<br />

46.19±16.29 kPa), and CLSS (32.13±7.15 vs. 69.23±25.43<br />

kPa) respectively. The cutoff values of significant fibrosis were<br />

9.15 kPa by LSM (94.8% Sensitivity, 88.3% Specificity, 84.9%<br />

PPV, and 96.1.8% NPV), 27.5 kPa by SSM (94.8% Sensitivity,<br />

70.2% specificity, 68.8% PPV, and 95.1% NPV) and 40.85<br />

kPa by CLSM (92.4% Sensitivity, 91.1% specificity, 87.8%<br />

PPV, and 94.6% NPV). Conclusion: The measurement of liver,<br />

spleen stiffness by FibroScan TM or their combinations are useful<br />

for liver fibrosis assessment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mohamed A. Elmazaly, Ayman<br />

Alsebaey, Maha M. Elsabaawy, El-Sayed S. Tharwa, Hanaa M. Badran, Nermine<br />

Ehsan, Eman A. Rewisha<br />

1863<br />

Impact of comorbidities on the benefit of DAA therapy<br />

in Hepatitis C<br />

Javier Ampuero 1 , Carlota Jimeno 1 , Angela Rojas 1 , A. Ortega 2 ,<br />

Jose Miguel Rosales-Zabal 4 , Marta Maraver 3 , Raquel Millán 1 ,<br />

Blanca Fombuena 1 , Ramón Morillo 5 , J. M. Navarro 4 , Raul J.<br />

Andrade 2 , Manuel Romero-Gomez 1 ; 1 Unit for the Clinical Management<br />

of Digestive Diseases, Valme University Hospital, Sevilla,<br />

Spain; 2 Digestive Department, Hospital Virgen de la Victoria, Malaga,<br />

Spain; 3 Digestive Department, Hospital Juan Ramón Jiménez,<br />

Huelva, Spain; 4 Digestive Department, Hospital Costa del Sol,<br />

Marbella, Spain; 5 Farmacology Unit, Valme University Hospital,<br />

Sevilla, Spain<br />

Background: HCV therapy achieving sustained virological<br />

response(SVR) decreases dramatically all-cause mortality.<br />

New direct acting antivirals(DAAs) are able to achieve SVR<br />

on more than 90% of patients. However, the benefit of the<br />

treatment in patients with severe comorbidities, beyond of the<br />

eradication of the virus, remains unknown. Aim: To explore<br />

the impact of comorbidities scales on decision-making according<br />

to HCV therapy outcomes. Methods: Prospective multicenter<br />

study including 104 consecutively HCV patients(≥F2<br />

by transient elastography) treated with DAAs from January<br />

2015. Basal comorbidities were evaluated by: a) Charlson<br />

comorbidity index(CCI) (Charlson, J Chron Dis 1987); b)<br />

Modified Charlson comorbidity index(mCCI) (Berkman, Ann<br />

Intern Med 1992); c) CirCom index (Jepsen, Gastroenterol<br />

2014). Adverse events were collected from the starting date<br />

of the treatment and were classified according to healthcare<br />

requirement: a) hospital admission(>24h); b) hospital observation(


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1119A<br />

including each score, to find independent variables related to<br />

adverse events: a) CCI [HR 2.66(95%CI 1.53-4.65);p=0.001];<br />

b) mCCI [HR 2.35(95%CI 1.61-3.42);p=0.0001] and albumin<br />

[HR 0.99(95%CI 0.99-1.00);p=0.040]; c) CirCom [HR<br />

4.48(95%CI 1.98-10.14);p=0.0001] and albumin [HR<br />

0.997(95%CI 0.995-0.999);p=0.040]. Conclusion: HCV<br />

patients with higher basal comorbidities are at risk of suffering<br />

adverse events during first months of treatment. Charlson and<br />

CirCom comorbidities index appear to be important to evaluate<br />

the overall benefit of the HCV treatment, beyond of the<br />

eradication of the virus. Partients with a Charlson >3 requires<br />

individual analysis before starting DAA therapy.<br />

Disclosures:<br />

Jose Miguel Rosales-Zabal - Speaking and Teaching: janssen<br />

Manuel Romero-Gomez - Advisory Committees or Review Panels: Roche Farma,SA.,<br />

MSD, S.A., Janssen, S.A., Abbott, S.A.; Grant/Research Support: Ferrer,<br />

S.A.<br />

The following authors have nothing to disclose: Javier Ampuero, Carlota Jimeno,<br />

Angela Rojas, A. Ortega, Marta Maraver, Raquel Millán, Blanca Fombuena,<br />

Ramón Morillo, J. M. Navarro, Raul J. Andrade<br />

1864<br />

Perceived risk and HCV knowledge in a high risk nonbaby<br />

boomer population from a community based testing<br />

model<br />

Stephanie Tzarnas, Allison Brodsky, Hunana Chaudhry, Carla<br />

Coleman, Michelle Dougherty, Lora Magaldi, Carolyn Moy,<br />

Ta-Wanda Preston, Stacey B. Trooskin; Drexel University College<br />

of Medicine, Philadelphia, PA<br />

Background: Hepatitis C is the most common blood-borne<br />

infection in the US with a prevalence of 1.6%. CDC guidelines<br />

recommend high-risk individuals and individuals born between<br />

1945-1965 be tested for HCV. High-risk individuals may not<br />

be tested due to lack of access to primary care or because providers<br />

may not routinely assess risk factors. We aim to assess<br />

perceived risk for contracting HCV as well as HCV knowledge<br />

in a community based HCV testing program serving high risk<br />

individuals. Methods: Community based testing was conducted<br />

by two community based organizations. A brief survey was<br />

utilized to assess general demographic information, perceived<br />

risk of HCV infection, and HCV knowledge. Results: 1,587<br />

patients were tested from June-April 2015, 11.3% were antibody<br />

positive. Of the 1,587 individuals tested, 67% were born<br />

outside of the birth cohort. Of the 1,063 individuals tested<br />

outside of the birth cohort, 9.3% were HCV antibody positive.<br />

80% of individuals born outside of the birth cohort with reactive<br />

antibody test received confirmatory NAT testing. Of those<br />

individuals; 81% were chronically infected; only 3.2% had<br />

seen their PCP in the last 12 months. Notably, 97% of HCV<br />

antibody positive individuals outside of the birth cohort had<br />

histories of drug use; of this group 47% had injected drugs<br />

and 66% of those individuals admitted to sharing injection<br />

equipment. Of the HCV antibody positive individuals outside<br />

of the birth cohort, 75% had heard of HCV; 66% answered<br />

that there was a cure for HCV, and 33% believed that there<br />

was a vaccine for HCV. Of the non-baby boomer antibody<br />

positive individuals, 40% rated themselves high risk for contracting<br />

HCV, 15% rated themselves low risk, 11% rated themselves<br />

moderate risk, and 9% rated themselves not at risk for<br />

contracting HCV. Conclusions: Community based testing and<br />

linkage to care programs serving high risk individuals have<br />

the potential to engage individuals in subspecialty care, where<br />

the standard health care system may have failed to do so.<br />

Interestingly, many of the non-baby boomer antibody positive<br />

individuals rated their perceived risk for contracting HCV high,<br />

while few rated themselves not at risk. Community based testing<br />

programs are needed to ensure that high risk individuals<br />

outside of the 1945-1965 birth cohort receive proper HCV<br />

testing, treatment, and education.<br />

Disclosures:<br />

Stephanie Tzarnas - Grant/Research Support: Gilead Sciences<br />

Allison Brodsky - Grant/Research Support: Gilead<br />

Hunana Chaudhry - Grant/Research Support: Gilead Sciences<br />

Carla Coleman - Grant/Research Support: Gilead Sciences<br />

Michelle Dougherty - Employment: Opening Doors for Diverse Populations to<br />

Health Dispairites Research Under the NIH 1R25MD006792-01 The National<br />

Institute on Minority Health and Health Disparities (NIMHA) Dr. Shannon Marquez,<br />

PI; Grant/Research Support: Gilead Sciences<br />

Lora Magaldi - Grant/Research Support: Gilead Sciences<br />

Carolyn Moy - Grant/Research Support: Gilead<br />

Stacey B. Trooskin - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Gilead Sciences<br />

The following authors have nothing to disclose: Ta-Wanda Preston<br />

1865<br />

Investigation into the high prevalence of hepatitis C<br />

virus infection in a rural village in southwest China<br />

Shiqi Tao 1 , Guangyu Huang 1 , Li Wang 1,2 , Dengming He 1 , Zehui<br />

Yan 1 , Shitao Ding 1 , Yunjie Dan 1 , Cheng Xu 1 , Xianghua Zeng 1 ,<br />

Xiaohong Wang 1 , Yuming Wang 1 ; 1 Department of infectious<br />

diseases, Southwest Hospital, Third Military Medical University,<br />

Chongqing, China; 2 Medical college of Guiyang, Guiyang, China<br />

Purpose: To clarify the changing pattern of hepatitis C virus<br />

(HCV) prevalence in China, with the aim of developing appropriate<br />

and effective strategies for the prevention and treatment<br />

of this significant emerging disease. Methods: The residents of<br />

Village M, located in southwest China, voluntarily participated<br />

in this study. Blood samples were obtained and anti-HCV titers<br />

were tested to determine the HCV status of the participants.<br />

For those who were anti-HCV positive, HCV RNA levels and<br />

genotypes were subsequently tested. HBV-related factors and<br />

anti-HIV titers were also tested. In addition, the methodology<br />

historically used to sterilize needles in the medical center used<br />

by the villagers was recreated to test the effectiveness of this<br />

procedure. Results: Anti-HCV antibody testing showed that 258<br />

(50.4%) of the 512 participants were anti-HCV positive. The<br />

majority (180/258) of anti-HCV positive participants were also<br />

HCV RNA positive. Genotyping based on NS5B sequences<br />

showed that the predominant subtype of HCV was 3b (96.0%),<br />

followed by 6a (2.0%) and 1b (2.0%). The recreated medical<br />

procedure indicated that the transmission route might have<br />

been inadequately sterilized needles. The HBV infection rate<br />

among participants was 34.4% and the HBV/HCV co-infection<br />

rate was 17.6%, but none of the participants were anti-HIV<br />

positive. Conclusions: HCV infection was found to be highly<br />

prevalent in a village in southwest China. A new transmission<br />

or prevalence pattern of HCV infection was identified that<br />

involved inadequately sterilized needles. Co-infection with HBV<br />

among the anti-HCV positive individuals was also common.


1120A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

HCV molecular phylogenetic tree based on the NS5B sequences<br />

1866<br />

Hepatitis C Eradication with Sofosbuvir Leads to Significant<br />

Metabolic Changes<br />

Amilcar Morales, Manish B. Singla, Maria Sjogren, Dawn Torres;<br />

Gastroenterology, Walter Reed National Military Medical Center,<br />

Bethesda, MD<br />

Infection with hepatitis C can cause metabolic complications,<br />

including insulin resistance, hepatic steatosis and lipid abnormalities.<br />

The purpose of this study was to assess the effect of<br />

hepatitis C virus (HCV) eradication with sofosbuvir (SOF) regimens<br />

on glycemic and lipid tests. This is a retrospective analysis<br />

of HCV patients treated and cured with a SOF regimen (with<br />

ribavirin [RBV], interferon or simeprevir) from January 2014 to<br />

November 2014. Patients with HbA1c and lipid panels within<br />

6 months before and 6 months after therapy were identified. In<br />

those treated with a RBV regimen, HbA1c was obtained a minimum<br />

of 3 months post therapy. Medical history, demographics,<br />

HCV genotype, pre-therapy RNA, pre-therapy liver associated<br />

enzymes, pre and post-therapy hemoglobin and liver biopsies<br />

were included in our analysis. Patients with an increase in<br />

therapy for hyperlipidemia or diabetes during our study period<br />

were excluded. 106 patients were treated for HCV, of which<br />

27 patients met inclusion criteria. 66.7% were male, 37.0%<br />

were Caucasian, 18.5% were African American, 85.2% had<br />

genotype 1, and 33.3% had stage 3 or 4 fibrosis on biopsy.<br />

Mean age was 60.6 ± 5.2 years. Pre-treatment body-mass<br />

index (BMI) did not differ from post-treatment BMI (27.9 ± 4.8<br />

kg/m2 vs 27.4 ± 4.6 kg/m2, p = 0.662). No patients had<br />

an increase in their HbA1c. HbA1c significantly decreased<br />

with treatment of HCV (pretreatment 6.94 ± 1.05 mg/dL vs<br />

post-treatment 6.15 ± 0.77 mg/dL, p = 0.012). There was<br />

no correlation between decreases in HbA1c and changes in<br />

hemoglobin in patients treated with RBV. 12 patients had diabetes,<br />

and 2 patients had a decrease of their diabetic therapy<br />

after eradication. 20 patients had pre- and post-treatment lipid<br />

panels; there was a significant increase in LDL and total cholesterol<br />

during the period of observation (LDL: 92.9 ± 26.8 mg/<br />

dL vs 122.8 ± 36.2 mg/dL, p = 0.005; total cholesterol (TC)<br />

163.2 ± 32.0 mg/dL vs 193.3 ± 45.7 mg/dL, p = 0.021).<br />

Our study showed a significant drop in HbA1c up to 6 months<br />

after eradication of HCV with a SOF regimen, which may be<br />

related to a decrease in inflammatory cytokines and regulation<br />

of insulin activation cascades. Our study is unique as most of<br />

our patients had clinically significant HgA1C prior to HCV<br />

therapy. Increased LDL and total cholesterol are likely due to<br />

HCV infectious mechanism; lipid alterations are frequently<br />

seen post therapy. This study suggests that physicians treating<br />

HCV patients should reassess preventive medicine measures<br />

after therapy, and the benefits of eradicating HCV may extend<br />

beyond eliminating the effects of chronic liver inflammation.<br />

Disclosures:<br />

Maria Sjogren - Speaking and Teaching: Gilead, Bristol Myers<br />

The following authors have nothing to disclose: Amilcar Morales, Manish B.<br />

Singla, Dawn Torres<br />

Disclosures:<br />

The following authors have nothing to disclose: Shiqi Tao, Guangyu Huang, Li<br />

Wang, Dengming He, Zehui Yan, Shitao Ding, Yunjie Dan, Cheng Xu, Xianghua<br />

Zeng, Xiaohong Wang, Yuming Wang<br />

1867<br />

Hepatitis C and Renal Disease: Differences in Patient<br />

Characteristics and Clinical Outcomes in the United<br />

States<br />

Senaka Peter 1 , Craig Solid 2 , Tanya Bovitz 2 , Haifeng Guo 2 , Allan<br />

J. Collins 2 , Jean Marie Arduino 1 ; 1 Merck & Co., Inc, Kenilworth,<br />

NJ; 2 Chronic Disease Research Group, Minneapolis, MN<br />

Purpose: To describe characteristics and outcomes of US commercially-insured<br />

patients with hepatitis C virus (HCV) infection<br />

with and without renal impairment. Methods: In a cross-sectional<br />

analysis, we identified an HCV cohort of patients aged<br />

20 to 64 using Truven MarketScan claims database based<br />

on ICD-9 diagnosis codes in 2011. Among this cohort, we<br />

identified those with evidence of chronic kidney disease (CKD)<br />

and end-stage renal disease (ESRD) by ICD-9 and CPT codes.<br />

Adjusted logistic regression was used to estimate the odds of<br />

clinical outcomes during a 1-year follow-up period. Results:<br />

Of the 35,965 HCV patients identified, when compared with<br />

HCV patients without evidence of renal impairment, those with<br />

renal impairment were older, more often male, and had a<br />

higher prevalence of most comorbidities, including hepatitis<br />

B virus, cirrhosis, decompensated cirrhosis, and hepatocellular<br />

carcinoma. Renal disease (either non-ESRD CKD or ESRD)<br />

was associated with a significant increase in the rates of most<br />

clinical outcomes studied, including heart failure (HF), sepsis,<br />

pneumonia, stroke, AMI, and severe anemia with the need for<br />

blood transfusion (Figure). When repeated using Medicare<br />

data and HCV patients aged 65 years and older, similar results<br />

were obtained (data not shown). Conclusions: Our findings<br />

suggest that HCV patients with concurrent renal disease have a<br />

significantly higher disease burden and likelihood of negative<br />

health outcomes. These findings are of particular interest given<br />

the efficacy of new HCV treatment options. Future <strong>studies</strong> can<br />

assess how these new treatments may influence the likelihood<br />

of adverse outcomes among HCV patients with renal impairment.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1121A<br />

majority were male (HBV DC 82%, HBV HCC 86%, HCV DC<br />

76%, HCV HCC 81%). The number of incident hospitalizations<br />

for HCV DC and HCV HCC has increased over time (Figure).<br />

In contrast, incident hospitalizations for HBV DC and HBV HCC<br />

have remained stable (Figure). Over the study period, estimated<br />

uptake of antiviral therapy has increased for HBV (2% to<br />

5%), but remained low for HCV (1-2% per annum). Conclusion<br />

The burden of HCV-related hospitalizations due to ESLD has<br />

increased markedly. Planned linkage to individual-level data<br />

on HBV and HCV antiviral therapy prescriptions should provide<br />

further insights into the contrasting pattern in ESLD burden.<br />

Population level monitoring of HCV ESLD burden will be particularly<br />

crucial to evaluate the impact of interferon-free regimens.<br />

Disclosures:<br />

Senaka Peter - Employment: Merck<br />

Allan J. Collins - Advisory Committees or Review Panels: KDIGO, Amgen, IKEA-J,<br />

NKF KDPN, Hospira; Board Membership: Kidney Health Australia; Consulting:<br />

Amgen, NxStage, Keryx, Hospira, Relypsa, Bayer, ZS Pharma; Employment:<br />

Hennepin Healthcare System; Grant/Research Support: Amgen, Merck,<br />

NxStage, ZS Pharma, Keryx, Peer Kidney Care Initiative; Speaking and Teaching:<br />

Hospira, Amgen<br />

Jean Marie Arduino - Employment: Merck & Co., Inc<br />

The following authors have nothing to disclose: Craig Solid, Tanya Bovitz, Haifeng<br />

Guo<br />

1868<br />

Trends in end-stage liver disease among people notified<br />

with HBV or HCV in New South Wales, Australia: 2000-<br />

2014<br />

Reem K. Waziry 1 , Jason Grebely 1 , Janaki Amin 1 , Maryam Alavi 1 ,<br />

Behzad Hajarizadeh 1 , Jacob George 2 , Gail Matthews 1 , Matthew<br />

Law 1 , Gregory Dore 1 ; 1 The Kirby institute, University of New South<br />

Wales, Sydney, NSW, Australia; 2 Storr Liver Unit, Westmead Millennium<br />

Institute for Medical Research and Westmead Hospital,<br />

University of Sydney, Westmead, NSW, Australia, Sydney, NSW,<br />

Australia<br />

Study aims To assess trends in hospitalizations for end-stage<br />

liver disease (ESLD); (decompensated cirrhosis (DC) and hepatocellular<br />

carcinoma (HCC)) among people notified with HBV<br />

or HCV in New South Wales (NSW), Australia. Methods<br />

HBV and HCV cases notified to the NSW Health Department<br />

between 1993 and 2012 were linked to data on hospitalizations<br />

(2000-2014). Time trends in hospitalizations (incident<br />

and total) due to DC (including ascites, esophageal varices<br />

with bleeding, hepatic failure, alcoholic hepatic failure, and<br />

alcoholic liver cirrhosis) and HCC were evaluated [International<br />

Classification of Diseases (ICD 10) coding]. Results<br />

Between 1993 and 2012, a total of 151,307 individuals were<br />

notified with HBV (n=54,399), HCV (n=93,099), or HBV/<br />

HCV (n=3,809). From 2000 to 2014, there were 24,130<br />

(HBV=2,346; HCV=20,713, HBV/HCV=1,071) and 7,044<br />

(HBV=2,583; HCV=4,180, HBV/HCV =281) hospitalizations<br />

for DC and HCC, respectively. Among all hospitalizations<br />

for ESLD, the number of incident hospitalizations was<br />

6,031 (HBV=907, HCV=4,848, HBV/HCV=276) for DC and<br />

2,055 (HBV=732, HCV=1,244, HBV/HCV=79) for HCC.<br />

Median age at admission with ESLD was 57 (IQR=12) and<br />

58 (IQR=20) for HBV DC, 61 (IQR=16) for HBV HCC, and 51<br />

(IQR=11) for HCV DC and 58 (IQR=13) for HCV HCC .The<br />

Disclosures:<br />

Jason Grebely - Advisory Committees or Review Panels: Merck, Gilead; Grant/<br />

Research Support: Merck, Gilead, Abbvie, BMS<br />

Jacob George - Advisory Committees or Review Panels: Roche, BMS, MSD,<br />

Gilead, Janssen, Abbvie; Grant/Research Support: MSD<br />

Gail Matthews - Advisory Committees or Review Panels: gilead; Consulting: Viiv;<br />

Grant/Research Support: Gilead Sciences, janssen; Speaking and Teaching:<br />

BMS, MSD<br />

Matthew Law - Grant/Research Support: Boehringer Ingelhiem, Gilead Sciences,<br />

Merck Sharp & Dohme, Bristol-Myers Squibb, Janssen-Cilag, ViiV HealthCare<br />

Gregory Dore - Board Membership: Gilead, Merck, Abbvie, Bristol-Myers<br />

Squibb; Grant/Research Support: Gilead, Merck, Abbvie, Bristol-Myers Squibb;<br />

Speaking and Teaching: Gilead, Merck, Abbvie, Bristol-Myers Squibb<br />

The following authors have nothing to disclose: Reem K. Waziry, Janaki Amin,<br />

Maryam Alavi, Behzad Hajarizadeh<br />

1869<br />

The influence of Hepatitis C virus infection and its clearance<br />

on the serum lipid profile and glucose level<br />

Yinping Li, Xiaomei Wang, Hongqin Xu, Ruihong Wu, Xiumei Chi,<br />

Xiuzhu Gao, Yu Pan, Junqi Niu; Hepatology, The First Hospital of<br />

Jilin University, Changchun, China<br />

Chronic hepatitis C (CHC) is associated with a disturbance of<br />

lipid metabolism and glucose intolerance, and the clearance<br />

of HCV may be followed by a decrease in serum total cholesterol<br />

(TC), triglyceride (TG) and insulin resistance to adverse<br />

levels. The present study was designed to determine the impact<br />

of CHC and its treatment on circulating lipids and glucose<br />

levels. A total of 748 patients with HCV infection were retrospectively<br />

enrolled, Fasting serum total cholesterol (TC), triglyceride<br />

(TG) and blood glucose (FBG) levels in these patients<br />

were compared with 664 healthy individuals. Serum TC, TG<br />

and systematic glucose metabolism were measured in 183<br />

patients with chronic hepatitis C receiving interferon α-2b plus<br />

ribavirin at baseline, at the end of treatment, and at week 24<br />

post-treatment. We analysed systemic glucose metabolism in<br />

terms of the following indices: fasting insulin (FINS), C-peptide<br />

(FCP), homeostasis model assessment for insulin resistance<br />

(HOMA-IR) and beta-cell function (HOMA-β) and the insulin


1122A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

sensitivity index (ISI).Patients with HCV infection had significantly<br />

lower TC and TG than healthy controls. This difference<br />

remained significant when adjusted for sex, age and BMI.<br />

During the antiviral treatment, we found a significant reduction<br />

of TC levels independently from the outcome of treatment,<br />

but this rebounded to baseline in patients achieving a SVR<br />

(4.51-4.45mmol/l, P=0.509) and stayed below baseline in<br />

patients without SVR (4.39-4.72mmol/l, P=0.004). Serum TG<br />

levels were also significantly elevated in both groups at the end<br />

of treatment, and persisted until week 24 post-treatment only<br />

in the patients with SVR. In sustained responders, fasting insulin<br />

(P=0.007), C-peptide (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1123A<br />

guidelines during the time of this study (including prioritizing<br />

RX among those with advanced liver disease and non-genotype<br />

1). Anticipation of more efficacious therapy (1 st generation<br />

DAAs in development > 2007; and release of 2 nd generation<br />

DAAs >2013) also influenced RX trends.<br />

Disclosures:<br />

The following authors have nothing to disclose: Carla V. Rodriguez, Kevin Rubenstein,<br />

Haihong Hu, Michael A. Horberg<br />

1872<br />

Can Hepatitis C be eliminated from a high disease burden<br />

country like Pakistan?<br />

Saeed Hamid 1 , Huma Qureshi 2 , Asad A. Chaudhry 3 , Zaigham<br />

Abbas 4 , Altaf Alam 5 , Altaf Baqir 6 , Javed I. Farooqi 7 , Muhammad<br />

S. Memon 8 , Arif A. Nawaz 9 , Homie Razavi 10 , Devin M. Razavi-Shearer<br />

11 , Amjad Salamat 12 , Masood Siddiq 13 , Arif Siddiqi 14 ,<br />

Ghias Un Nabi Tayyab 15 , Muazzam Uddin 16 , Aasim Yusuf 17 ,<br />

Bader F. Zuberi 18 , Aamir G. Khan 19 , Syed M. Jafri 1 ; 1 Medicine,<br />

Aga Khan University, Karachi, Pakistan; 2 Pakistan Medical<br />

Research Council, Karachi, Pakistan; 3 Liver Foundation, Gujranwala,<br />

Pakistan; 4 Sindh Institute of Urology and Transplantation,<br />

Karachi, Pakistan; 5 Skeikh Zayed Medical Institute, Lahore, Pakistan;<br />

6 Nishtar Hospital, Multan, Pakistan; 7 Post Graduate Medical<br />

Institute, Peshawer, Pakistan; 8 AIMS, Hyderabad, Pakistan;<br />

9 Fatima Memorial Hospital, Lahore, Pakistan; 10 Center for Disease<br />

Analysis, Denver, CO; 11 Center for Disease Analysis, Denver, Pakistan;<br />

12 CMH, Rawalpindi, Pakistan; 13 Fauji Foundation Hospital,<br />

Rawalpindi, Pakistan; 14 Allama Iqbal Medcial College, Lahore,<br />

Pakistan; 15 PGMI, Lahore, Pakistan; 16 Bolan Medcial College,<br />

Quetta, Pakistan; 17 Shaukat Khanum Hospital, Lahore, Pakistan;<br />

18 Dow Medical College, Karachi, Pakistan; 19 Khyber Teaching<br />

Hospital, Peshawer, Pakistan<br />

Background: Pakistan has the highest number of individuals<br />

with active hepatitis C virus (HCV) infection of all countries<br />

except China, but unlike China, the total number of infections<br />

in Pakistan is not declining. An estimated 4.2% of the general<br />

population or 7,192, 900 persons were infected with HCV in<br />

2014. Therefore, it is critical to assess the impact of strategies<br />

that can tackle the ongoing transmission and manage the HCV<br />

disease burden. Methods: Using a Markov model, the progression<br />

of the HCV infected population was quantified and<br />

the impact of different intervention strategies (prevention, treatment<br />

and screening) was measured. The inputs were gathered<br />

through a review of local and international publications, analysis<br />

of local data, and discussions with an expert panel. Findings:<br />

Elimination of HCV in 15 years is only feasible through<br />

a combination of aggressive prevention, treatment and screening.<br />

An estimated 230,000 new infections occur annually.<br />

While 85,000 individuals are treated each year, the total number<br />

of HCV infections is projected to increase to 7,529,200<br />

by 2030. Contaminated medical equipment remains one of<br />

the main risk factors, and education campaigns combined with<br />

wide-scale availability of sterile equipment could reduce the<br />

new infections by 50%. In addition, increasing treatment to<br />

510,000 patients each year with high sustained viral response<br />

therapies would reduce the total number of HCV infections to<br />

less than 700,000 by 2030. Lower number of infections in 15<br />

years would require the treatment of elderly and youth as well.<br />

Implementation of this strategy would require an increase in<br />

screening from 100,000 newly diagnosed cases to 600,000<br />

annually. Conclusions: Elimination of HCV in Pakistan in 15<br />

years is feasible but requires wide scale education, availability<br />

of sterile equipment, screening and treatment. This strategy<br />

would avert 220,000 liverrelated deaths and 116,000 new<br />

hepatocellular carcinoma cases within 15 years.<br />

Disclosures:<br />

Homie Razavi - Management Position: Center for Disease Analysis<br />

The following authors have nothing to disclose: Saeed Hamid, Huma Qureshi,<br />

Asad A. Chaudhry, Zaigham Abbas, Altaf Alam, Altaf Baqir, Javed I. Farooqi,<br />

Muhammad S. Memon, Arif A. Nawaz, Devin M. Razavi-Shearer, Amjad<br />

Salamat, Masood Siddiq, Arif Siddiqi, Ghias Un Nabi Tayyab, Muazzam<br />

Uddin, Aasim Yusuf, Bader F. Zuberi, Aamir G. Khan, Syed M. Jafri<br />

1873<br />

Dynamic change of α-fetoprotein, platelet counts and<br />

aminotransferase-to-platelet ratio index (APRI) Predict<br />

Hepatocellular Carcinoma in chronic hepatitis C patients<br />

after antiviral therapy<br />

Cheng-Kun Wu, Kuo-Chin Chang, Po-Lin Tseng, Sheng-Nan Lu,<br />

Chien-Hung Chen, Jing-Houng Wang, Chuan-Mo Lee, Ming-Tsung<br />

Lin, Chao-Hung Hung, Yi-Hao Yen, Tsung-Hui Hu; Division of Hepato-Gastroenterology,<br />

Department of Internal Medicine, Kaohsiung<br />

Chang Gung Memorial Hospital, Taiwan, Kaohsiung City, Taiwan<br />

Background : Some patients who achieved viral eradication<br />

still developed HCC in the future. It is very important to<br />

accurately stratify the risks of HCC development in both SVR<br />

and non-SVR groups. Less is known about the role of dynamic<br />

change of serum markers on influence of HCC development.<br />

Aims: To clarify the relationship between the dynamic change<br />

of serum biomarkers and HCC risks Methods: We conducted<br />

a large-scale, long-term cohort study with 2405 CHC patients<br />

who have been treated with interferon (IFN) or pegylated interferon<br />

(peg-IFN) plus ribavirin therapies enrolled. After exclusion<br />

of patient with HCC development before treatment or<br />

within 6 months after the end of therapy, loss to follow-up and<br />

who didn’t have complete laboratory data, 1903 patients were<br />

finally assessed for analysis. The collection of pre-treatment<br />

hematological and biochemical data was completed before<br />

IFN-based therapy. Post-treatment hematological and biochemical<br />

data were collected at least 1 year after IFN-based therapy<br />

and to the date of latest follow-up if possible. AFP data were<br />

obtained more than 6 months prior to HCC development to<br />

exclude HCC-related AFP elevation. Results: HCC developed<br />

in 142 patients during follow-up. The entire cohort patients<br />

had decreased AFP after IFN-based therapy irrespective of<br />

viral eradication. Significantly decreased APRI was observed<br />

in patients with SVR and non-SVR groups. The SVR patients<br />

had increased platelet counts after antiviral therapy, whereas<br />

non-SVR patients continued to have decreased platelet counts<br />

during follow-up. On multivariate analysis, older age, liver cirrhosis,<br />

lower hemoglobin level, non-SVR status, higher pre- and<br />

post-treatment AFP (^15 ng/ml) and higher post-treatment APRI<br />

(≥0.7) were significant risk factors for HCC. Risk factors for<br />

HCC among patients with SVR were older age, liver cirrhosis<br />

and higher pre- and post-treatment AFP; among those without<br />

SVR, the risk factors were liver cirrhosis, lower hemoglobin<br />

level, lower pre-treatment platelet


1124A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1874<br />

Linkage to Care and Treatment of Hepatitis C in a Federally<br />

Qualified Health Center in New Haven, Connecticut<br />

Izona Bock 1 , Mina A. Garcia 1 , Robert Bruce 1,2 ; 1 Infectious Diseases,<br />

Cornell Scott Hill Health Center, New Haven, CT; 2 Yale<br />

Center for Clinical Investigation Scholar, Yale University, New<br />

Haven, CT<br />

Introduction: Approximately 2.7-3.9 million persons are<br />

infected with hepatitis C in the United States. The current standard<br />

therapy of hepatitis C is revolutionary given the high<br />

cure rates, short treatment duration, and relatively mild side<br />

effects. Our Federally Qualified Health Center (FQHC) serves<br />

approximately 39,000 underserved patients in New Haven<br />

County and provides treatment of substance abuse and chronic<br />

hepatitis C. The goal of this study was to develop a hepatitis<br />

C treatment cascade chart by determining the rates of linkage<br />

to care and treatment of patients with chronic hepatitis<br />

C in our center. Methods: With use of the electronic health<br />

record, we retrieved a list of patients coded with hepatitis C<br />

between 2013 and 2015. By performing a chart review, we<br />

determined the number evaluated specifically for hepatitis C<br />

treatment, those who were started on therapy, and the number<br />

who achieved virologic suppression. Results: Of the 7716<br />

patients screened for hepatitis C between 2013 and 2015,<br />

786 patients (10.2%) were identified as having hepatitis C.<br />

32% (250) of these patients were evaluated by a provider who<br />

treats hepatitis C. Ultimately, 10.4% (82) of patients started<br />

therapy and 6.8% (54) of those were found to have a hepatitis<br />

C viral load


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1125A<br />

between first and last LSM was 51 months (4-106). Median<br />

basal LSM was similar in treated and nontreated or between<br />

SVR and N-SVR. Fibrosis regression was observed in treated -<br />

49,2% vs NT-36,8% (p=0,30), SVR – 71% vs N-SVR- 26,7%<br />

(p=0.002). Analyzing the final LSM, the SVR group had a<br />

significant increase of the number of patients with stage < F2<br />

(plus 35,6%, p =0.001), in comparison to N-SVR (minus 6.4%,<br />

p ns). During follow-up, a significant decrease of median LSM<br />

was observed in SVR (from 6.55 to 5.55 kPa, p


1126A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

nificantly affects the long-term natural history of this process<br />

even in low fibrosis groups. Median monitoring intervals of 3<br />

years is suggested to allocate fibrosis and is crucial in non-SVR<br />

& non-Tx patients with F0-F2.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shaul Yaari, Orit Pappo, Saleh<br />

N. Daher, Ariel Benson, Nadav Ilani, Tova Goldberg-Clein, Muhammad Massarwa,<br />

Johnny Amer, Rifaat Safadi<br />

1879<br />

Evaluation of liver and spleen transient elastography in<br />

the diagnosis of esophageal varices<br />

Ayman Alsebaey 1 , Mohamed A. Elmazaly 1 , Maha M. Elsabaawy 1 ,<br />

El-Sayed S. Tharwa 1 , Hanaa M. Badran 1 , Nermine Ehsan 2 , Eman<br />

A. Rewisha 1 ; 1 Hepatology department, National Liver Institute,<br />

Shebin elkom menoufiya, Egypt; 2 Pathology, National Liver Institute,<br />

Shebin Elkoom, Egypt<br />

Background: Portal hypertension with subsequent esophageal<br />

varices (EVs) development is a common complication of HCV<br />

related cirrhosis. Aim: is to evaluate the role of liver stiffness<br />

measurement (LSM), spleen stiffness measurement (SSM) and<br />

their combination (CLSM) using FibroScan TM in diagnosis of<br />

EVs. Methods: One hundred sixty five HCV related F3-4 Metavir<br />

score fibrosis were included. Liver, renal function tests, CBC,<br />

INR and abdominal ultrasonography were done before the<br />

FibroScan TM . Transient elastography measurement was done<br />

using FibroScan TM in the supine position after 6-8 hours fasting<br />

followed by diagnostic esophageogastroscopy. Varices<br />

were classified into none (n= 110), small (n =30) and large (n<br />

=25). Results: Patients with varices had higher serum bilirubin<br />

(1.68±0.82 vs. 1.00±0.55 mg/dl) and lower platelet count<br />

(105.09±31.34vs. 161.21±52.97 ×10 3 /mL) that patients<br />

without varices (p=0.001). The patients with varices had statistically<br />

significant (p=0.001) higher platelets spleen ratio<br />

(671.14±258.89 vs 1215.41±445.58), LSM (31.93±13.29<br />

vs. 17.55±6.53 kPa), SSM (62.85±12.71 vs. 36.94±8.83<br />

kPa) and CLSS (94.78±20.98 vs. 54.49±12.84 kPa) than<br />

patients without varices. In patients with small and large varices<br />

LSM was comparable (30.84±12.69 vs 33.64±13.97 kPa;<br />

p=0.391) but a statistically significant difference was detected<br />

with SSM (59.92±13.47 vs. 66.98±8.67 kPa; p=0.031) and<br />

CLSS (90.76±21.76 vs 100.62±19.00 kPa; p=0.043). With<br />

a cutoff of 20.4 kPa LSM (81.0% sensitivity, 71.8% specificity,<br />

52.3% PPV, 90.8% NPV, 74.3% accuracy), 43.2 kPa SSM<br />

(92.9% sensitivity, 84% specificity, 69.6% PPV, 96.9% NPV,<br />

86.8% accuracy) and 59.3 kPa CLSS (95.2% sensitivity, 70%<br />

specificity, 54.8% PPV, 97.5% NPV, 76.9% accuracy) esophageal<br />

varices can be detected. Conclusion: The measurement of<br />

liver, spleen stiffness by FibroScan TM or their combinations are<br />

useful for esophageal varices diagnosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ayman Alsebaey, Mohamed A.<br />

Elmazaly, Maha M. Elsabaawy, El-Sayed S. Tharwa, Hanaa M. Badran, Nermine<br />

Ehsan, Eman A. Rewisha<br />

1880<br />

Self-reported Alcohol Use at Time of HCV Antibody<br />

Screening for Individuals Tested in Community Settings<br />

Michelle Dougherty, Allison Brodsky, Hunana Chaudhry, Carla<br />

Coleman, Lora Magaldi, Carolyn Moy, Ta-Wanda Preston, Stephanie<br />

Tzarnas, Stacey B. Trooskin; Drexel University College of<br />

Medicine, Philadelphia, PA<br />

Background: According to AASLD/IDSA, individuals with HCV<br />

should avoid alcohol. Although there is no safe level of alcohol<br />

consumption, those who consume three or more drinks daily<br />

are at a higher risk for accelerated liver damage. We aim<br />

to assess alcohol use at time of HCV antibody screening for<br />

those tested in community based settings. Methods: A survey<br />

was administered to individuals receiving point of care HCV<br />

testing in community based settings. Data was collected from<br />

June, 2014 to April, 2015. Results: Information about alcohol<br />

consumption was collected from 1,254 individuals who were<br />

tested in community based settings. Of these 1,254 individuals,<br />

11.80% (n=148) tested positive for HCV antibody. 29%<br />

of those who tested positive for HCV antibody reported current<br />

alcohol consumption, compared to 35.7% of individuals who<br />

had no antibody detected. Of the anti-HCV positive individuals<br />

who answered follow up questions regarding alcohol consumption<br />

(n=41), 36.6% reported drinking four or more times per<br />

week, compared to 14.6% of anti-HCV negative individuals<br />

who answered the same follow up questions regarding alcohol<br />

consumption (n=417) [p=0.0003]. Additionally, of the anti-<br />

HCV positive individuals who answered the follow up questions,<br />

17.1% reported consuming seven or more drinks when<br />

they drink, compared to 5.3% of anti-HCV negative individuals<br />

[p=0.0031]. Conclusion: Although individuals who tested<br />

negative for HCV in the community were more likely to report<br />

that they drank alcohol than those who tested positive for antibody<br />

to HCV, the quantity and frequency of alcohol use was<br />

higher among anti-HCV positive individuals. Because alcohol<br />

use is associated with accelerated liver damage, these findings<br />

underscore the need for community outreach, education and<br />

HCV testing programs.<br />

Disclosures:<br />

Michelle Dougherty - Employment: Opening Doors for Diverse Populations to<br />

Health Dispairites Research Under the NIH 1R25MD006792-01 The National<br />

Institute on Minority Health and Health Disparities (NIMHA) Dr. Shannon Marquez,<br />

PI; Grant/Research Support: Gilead Sciences<br />

Allison Brodsky - Grant/Research Support: Gilead<br />

Hunana Chaudhry - Grant/Research Support: Gilead Sciences<br />

Carla Coleman - Grant/Research Support: Gilead Sciences<br />

Lora Magaldi - Grant/Research Support: Gilead Sciences<br />

Carolyn Moy - Grant/Research Support: Gilead<br />

Ta-Wanda Preston - Grant/Research Support: Gilead Sciences<br />

Stephanie Tzarnas - Grant/Research Support: Gilead Sciences<br />

Stacey B. Trooskin - Advisory Committees or Review Panels: Gilead Sciences;<br />

Grant/Research Support: Gilead Sciences<br />

1881<br />

Rate and Predictors of Serum HCV-RNA >6 Million IU/<br />

mL in Chronic Hepatitis C Patients<br />

Pablo Barreiro 2 , Pablo Labarga 1 , Jose V. Fernandez-Montero 4 ,<br />

Carmen de Mendoza 3 , Laura Benitez 3 , Jose M Peña 2 , Vincent<br />

Soriano 2 ; 1 La Luz Clinic, Madrid, Spain; 2 La Paz University Hospital<br />

& IdiPAZ, Madrid, Spain; 3 Internal Medicine, Puerta de Hierro<br />

Research Institute, Majadahonda, Spain; 4 University Hospital<br />

Crosshouse, Kilmarnock, United Kingdom<br />

Background: Baseline serum HCV-RNA predicts sustained<br />

virological response in chronic hepatitis C patients treated<br />

with both peginterferon-ribavirin and more recently with some<br />

direct-acting antiviral regimens. Whereas thresholds around<br />

800,000 IU/mL split out good and poor responders to interferon-based<br />

regimens, a greater threshold at 6,000,000 IU/<br />

mL has been propose to discriminate treatment outcomes on<br />

sofosbuvir-ledipasvir. In comparison with the general population,<br />

immunosuppressed individuals exhibit greater viral load<br />

values. Methods: Serum HCV-RNA values recorded from all<br />

chronic hepatitis C patients consecutively attended at our clinic<br />

during the last decade were analyzed. The COBAS TaqMan<br />

HCV test v2.0 with the high pure system was used, with a


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1127A<br />

lower limit of quantification of 25 IU/mL. Results: A total of<br />

816 individuals with detectable serum HCV-RNA were identified.<br />

The main characteristics of this population were as follows:<br />

mean age: 48.6±6.9 years-old; male gender 73.4%;<br />

mean ALT: 82.6±70.9 IU/L; mean HCV-RNA: 6.02±0.75 log<br />

IU/mL; HCV genotypes 1 or 4: 80.6%; mean liver stiffness:<br />

11.3±10.4 kPa; advanced liver fibrosis (>9.5 kPa): 34.9%;<br />

IL28B-CC: 35.4%. HIV coinfection was found in 78.7% (mean<br />

CD4 count: 496±295 cells/uL; HAART: 91%). Overall, 127<br />

(15.6%) had serum HCV-RNA values >6 million IU/mL. This<br />

high viremia was found in 18.2% of HIV-positive versus 5.7%<br />

of HIV-negative (p6 million IU/mL were as follows: mean age<br />

48.2±5.2 years-old; male gender 78.7%, HCV genotypes 1-4:<br />

87.6%; advanced liver fibrosis 33.6%; IL28B-CC 41.1%; and<br />

HIV pos 92.1%. In multivariate analysis, serum HCV-RNA >6<br />

million IU/mL was only significantly associated with HIV coinfection<br />

(OR: 4.03; 95% CI: 1.98-8.19, p6 million IU/mL is overall seen<br />

in 6% of chronic hepatitis C patients, being significantly more<br />

frequent in patients with HIV coinfection (18%) and/or with<br />

HCV genotypes 1/4 (10%).<br />

Disclosures:<br />

The following authors have nothing to disclose: Pablo Barreiro, Pablo Labarga,<br />

Jose V. Fernandez-Montero, Carmen de Mendoza, Laura Benitez, Jose M Peña,<br />

Vincent Soriano<br />

1882<br />

Acute hepatitis and reactivation of hepatitis C virus in<br />

cancer patients receiving systemic chemotherapy<br />

Hae Lim Lee, Si Hyun Bae, Hyun Yang, Jeong Won Jang, Jong<br />

Young Choi, Seung Kew Yoon; the catholic university of Korea,<br />

Seoul, Korea (the Republic of)<br />

Introduction : Reactivation of hepatitis C virus (HCV) is less<br />

common and only a few cases with severe consequences have<br />

been reported compared to HBV infection when it comes to<br />

receiving systemic chemotherapy. Many reported <strong>studies</strong> suggested<br />

that HCV reactivation was associated with chemotherapy<br />

containing rituximab and/or steroids. This retrospective<br />

study was aim to characterize acute exacerbation of HCV<br />

infection in cancer patients receiving systemic chemotherapy.<br />

Patients and methods : A total of 102 patients reviewed from<br />

January 1, 2008 to March 1, 2015 at Seoul St. Mary’s hospital,<br />

positive for anti-HCV antibody and receiving systemic chemotherapy<br />

for cancer except hepatocellular carcinoma, were<br />

included. 7 patients who lack HCV RNA titers were excluded.<br />

Hepatitis was defined as over three fold increase in ALT level.<br />

HCV reactivation was defined as reappearance of RNA which<br />

was previously undetected or resolved or RNA increasd over<br />

10 fold compared with the baseline level. Results : Among the<br />

95 patients with HCV infection and cancer, 53 patients (56%)<br />

showed hepatitis of all cause. 23(43%) of the 53 patients had<br />

hematological tumors and these were DLBCL(6, 11%), acute<br />

leukemia(n=11, 20%) and others(n=6, 11%). 14(33%) of the<br />

42 patients without hepatitis had hematological tumor. Neither<br />

severe hepatitis nor hepatic decompensation was noted. 32<br />

patients had RNA level determination before and after chemotherapy.<br />

HCV reactivation was noted in 10 patients (31%).<br />

Among them, no patient (0/10) with acute leukemia regardless<br />

of stem cell transplantation was identified to have HCV<br />

reactivation. In univariate analysis for HCV reactivation, no<br />

statistically significant factor was found. In 4 of 10 patients,<br />

hepatitis was documented in the period of HCV reactivation. 2<br />

patients were diagnosed as DLBCL and had received R-CHOP<br />

for chemotherapy regimen. Others were breast cancer and<br />

malignant melanoma. Conclusion : Cancer patients with HCV<br />

infection can be generally treated with systemic chemotherapy.<br />

Although systemic chemotherapy can induce hepatitis and HCV<br />

reactivation, no severe hepatitis or hepatic decompensation<br />

was documented, thus it might have less clinical significance<br />

compared to HBV infection.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hae Lim Lee, Si Hyun Bae, Hyun<br />

Yang, Jeong Won Jang, Jong Young Choi, Seung Kew Yoon<br />

1883<br />

Healthcare Utilization and Costs among Hepatitis C<br />

Patients with and without Impaired Renal Function in<br />

the United States<br />

Senaka Peter 1 , Craig Solid 2 , Tanya Bovitz 2 , Haifeng Guo 2 , Allan<br />

J. Collins 2 , Jean Marie Arduino 1 ; 1 Merck & Co., Inc., Kenilworth,<br />

NJ; 2 Chronic Disease Research Group, Minneapolis, MN<br />

Purpose: To compare rates of healthcare utilization and costs<br />

among US commercially-insured patients with hepatitis C virus<br />

(HCV) infection with and without renal impairment. Methods:<br />

This cross-sectional analysis used Truven MarketScan claims<br />

database to identify an HCV cohort of patients aged 20 to<br />

64 based on ICD-9 diagnosis codes in 2011. Within the<br />

cohort, we identified those with evidence of chronic kidney<br />

disease (CKD) or end-stage renal disease (ESRD) by ICD-9<br />

codes. All-cause healthcare utilization (HCU) and medical<br />

costs for all health encounters excluding pharmacy during<br />

a 1-year follow-up period were evaluated. The subset of allcause<br />

HCU and costs related to HCV were identified as those<br />

claims with a principal diagnosis code for HCV or HCV-related<br />

liver disease. Generalized linear models, adjusted for<br />

patient characteristics and baseline comorbidities, were used<br />

to test for differences in costs. Results: A total of 35,965 HCV<br />

patients (mean age= 53.2 years; 63% male) were identified<br />

in 2011. All-cause and HCV-related HCU in the inpatient and<br />

outpatient settings were substantially higher for HCV patients<br />

with non-ESRD CKD or ESRD compared to HCV patients with<br />

no evidence of renal impairment during the 2012 follow-up<br />

year (Table). Compared to patients without evidence of renal<br />

impairment, all-cause and HCV-related medical costs (Table)<br />

were also significantly higher for those with renal disease, as<br />

demonstrated by adjusted rate ratios of 1.74 and 2.53 (both<br />

p


1128A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Senaka Peter - Employment: Merck<br />

Allan J. Collins - Advisory Committees or Review Panels: KDIGO, Amgen, IKEA-J,<br />

NKF KDPN, Hospira; Board Membership: Kidney Health Australia; Consulting:<br />

Amgen, NxStage, Keryx, Hospira, Relypsa, Bayer, ZS Pharma; Employment:<br />

Hennepin Healthcare System; Grant/Research Support: Amgen, Merck,<br />

NxStage, ZS Pharma, Keryx, Peer Kidney Care Initiative; Speaking and Teaching:<br />

Hospira, Amgen<br />

Jean Marie Arduino - Employment: Merck & Co., Inc<br />

The following authors have nothing to disclose: Craig Solid, Tanya Bovitz, Haifeng<br />

Guo<br />

1884<br />

Does pregnancy protect against fibrosis or promote SVR<br />

in HCV infected women?<br />

Christine Pocha 1 , Pascal Benkert 2 , Salome A. Keller 1 , Beat Müllhaupt<br />

3 , David Semela 4 , Markus H. Heim 5 , Darius Moradpour 6 ,<br />

Andreas Cerny 7 , Raffaele Malinverni 8 , Francesco Negro 9 ,<br />

Jean-Francois Dufour 1 ; 1 Hepatology, University Clinic for visceral<br />

Surgery and Medicine,, University Hospital Bern, Bern, Switzerland;<br />

2 Clinical Trial Unit, University of Basel, Basel, Switzerland;<br />

3 Gastroenterology and Hepatology, University of Zürich, Zürich,<br />

Switzerland; 4 Gastroenterology and Hepatology, Kantonsspital<br />

St. Gallen, St. Gallen, Switzerland; 5 Gastroenterology and Hepatology,<br />

University Hospital Basel, Basel, Switzerland; 6 Gastroenterology<br />

and Hepatology, University of Lausanne, Lausanne,<br />

Switzerland; 7 Internal Medicine, Clinica Luganese Moncucco,<br />

Lugano, Switzerland; 8 Internal Medicine, Hôpital Neuchâtelois,<br />

Neuchâtel, Switzerland; 9 Gastroenterology, Hepatology and Clinical<br />

pathology, University Hospital, Geneva, Switzerland<br />

BACKGROUND: Sex specific differences in liver disease<br />

account for less hepatic decompensation in female cirrhotics,<br />

higher spontaneous HCV clearance and better chronic hepatitis<br />

C treatment response. Preclinical evidence suggests that<br />

relaxin, a pregnancy specific hormone, shows antifibrotic<br />

features. We aimed to evaluate the influence of past pregnancies<br />

on the rate of liver fibrosis progression and sustained<br />

virologic response (SVR) in women. METHODS: We extracted<br />

baseline and follow up demographic and clinical data on<br />

all women enrolled in Swiss Hepatitis C Cohort Study since<br />

2001. Differences between groups (women with/without history<br />

of pregnancy) were investigated using Fisher’s exact test<br />

for categorical variables and t-test for continuous variables.<br />

To reduce potential confounders, propensity score techniques<br />

were used. Cox proportional hazards model was used to analyze<br />

associations between predictors and cirrhosis taking into<br />

account the time to event (diagnosis of HCV infection to cirrhosis<br />

diagnosis). RESULTS: Among 1669 women included in<br />

the cohort, 1121 reported at least one pregnancy. Baseline<br />

characteristics were similar although women with pregnancy<br />

were significantly older at baseline, at time of first documented<br />

HCV serology and had longer follow-up time while non-pregnant<br />

women were more often co-infected with HIV. Overall,<br />

19.7% of patients were cirrhotic at a mean time from first documented<br />

HCV serology at 9.5 years, which was similar in both<br />

groups. Associated with cirrhosis were HCV therapy (HR 2.86,<br />

CI 2,39-3.41, p < 0.001), co-infection with HIV (HR 3.61,<br />

CI 2.68-4.86, p < 0.001), HCC (HR 2.47, CI 1.87-3.27,<br />

p < 0.001), diabetes (HR 2.50, CI 2.01-3.11, p < 0.001),<br />

and alcohol abuse (HR 2.08, CI 1.69-2.57, p < 0.001). Pregnancy<br />

was associated with a 18% risk reduction of cirrhosis<br />

(HR 0.82, CI 0.70-0.96, p = 0.015). Overall 39% of women<br />

received at least one HCV treatment with improvement or no<br />

change in fibrosis score. However, influence of pregnancy was<br />

non-conclusive. A history of pregnancy was associated with<br />

higher likelihood of achieving SVR with interferon-alpha-based<br />

antiviral therapy (OR 1.31, CI 0.95-1.81, p = 0.10). CONCLU-<br />

SIONS: Our data suggest that pregnancy is associated with a<br />

reduced risk of progression to cirrhosis as well as increased<br />

response to interferon-alpha- based HCV therapy. To further<br />

confirm a positive association between past pregnancies and<br />

lower rate of fibrosis progression we will integrate different<br />

fibrosis tests in further analyses.<br />

Disclosures:<br />

Francesco Negro - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

MSD, Gilead, AbbVie; Grant/Research Support: Gilead<br />

Jean-Francois Dufour - Advisory Committees or Review Panels: Bayer, BMS, Gilead,<br />

AbbVie, Novartis, Sillagen, Genfit<br />

The following authors have nothing to disclose: Christine Pocha, Pascal Benkert,<br />

Salome A. Keller, Beat Müllhaupt, David Semela, Markus H. Heim, Darius<br />

Moradpour, Andreas Cerny, Raffaele Malinverni<br />

1885<br />

Safety and Efficacy of Interferon/Ribavirin-Free Therapy<br />

in Septuagenerians and Octogenerians with Chronic<br />

Hepatitis C<br />

Rafael Stern 1 , Karin Kozbial 1 , Clarissa Freissmuth 1 , Stephanie<br />

Hametner 2 , Ramona Al Zoairy 3 , Stephan Moser 4 , Asia<br />

Karpi 4 , Hermann Laferl 5 , Rudolf E. Stauber 6 , Heinz M. Zoller 3 ,<br />

Andreas Maieron 2 , Wolfgang Vogel 3 , Ivo Graziadei 7 , Michael<br />

Gschwantler 4 , Harald Hofer 1 , Peter Ferenci 1 ; 1 Department of Internal<br />

Medicine III, Medical University of Vienna, Vienna, Austria;<br />

2 Department of Gastroenterology, Elisabethinen Hospital, Linz,<br />

Austria; 3 Department of Medicine II, Medical University of Innsbruck,<br />

Innsbruck, Austria; 4 Department of Medicine IV, Wilheminen-Hospital,<br />

Vienna, Austria; 5 Department of Internal Medicine,<br />

Kaiser-Franz-Josef-Hospital, Vienna, Austria; 6 Department of Internal<br />

Medicine, Medical University of Graz, Graz, Austria; 7 Department<br />

of Internal Medicine, Krankenhaus Hall, Hall, Austria<br />

Specific objective: Direct acting antivirals (DAA) have revolutionized<br />

treatment of chronic hepatitis C. Septua- and octogenarians<br />

were not included in clinical <strong>studies</strong> and therefore<br />

data on efficacy and safety of IFN-free treatments in the elderly<br />

population are missing. Elderly patients may have a decreased<br />

renal and hepatic function. Furthermore many are at risk for<br />

drug-drug interactions of DAA with drugs required for longterm<br />

treatment of other medical conditions. We investigated<br />

safety and efficacy of interferon (IFN) and ribavirin (RBV)-free<br />

treatments in patients >70 years with advanced liver disease<br />

due to chronic hepatitis C. Methods: 44 otherwise healthy<br />

patients >70 years (mean age: 74.6, range 70-88, m/f:<br />

19/26, cirrhosis: 32, F3:12, median HCV RNA 1.18x10 6<br />

IU/ml, HCV-genotype: GT-1a:6; 1b:34; 3:1; 4:2; missing:<br />

1; treatment experienced: 25; median platelet count: 137


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1129A<br />

G/l; median albumin: 39.5 g/l; median bilirubin: 0.86 mg/<br />

dl; median GFR using Cockroft-Gault formula: 66.6 ml/min)<br />

were enrolled. The patients were treated with sofosbuvir (SOF)<br />

400mg/day combined either with daclatasvir (DCV) 60mg/<br />

day, simeprevir (SMV) 150mg/day, or ledipasvir (LPV) 90<br />

mg/day. HCV RNA quantification was carried out with Abbott<br />

RealTime HCV (ART) assay (LLOQ


1130A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Table 1.<br />

CI: 95% confidence intervals; AUC: Area under Receiver Operating<br />

Curve; LR-likelihood ratio<br />

Disclosures:<br />

Keyur Patel - Advisory Committees or Review Panels: Merck; Consulting: Gilead<br />

Sciences, Santaris, Akros, Nitto Denko; Grant/Research Support: Bristol Myers<br />

Squibb<br />

The following authors have nothing to disclose: Bassem Matta, Julius M. Wilder,<br />

Tzu-Hao Lee<br />

1888<br />

The use of hepatitis C core antigen assay in the monitoring<br />

of treatment with direct antiviral agents in patients<br />

with chronic hepatitis C<br />

Elisabetta Loggi 1 , Carmela Cursaro 1 , Alessandra Scuteri 1 , Ranka<br />

Vukotic 1 , Martina Pirillo 1 , Marzia Margotti 1 , Valeria Guarneri 1 ,<br />

Silvia Galli 2 , Giuliano Furlini 2 , Claudio Galli 3 , Maria Paola Landini<br />

2 , Pietro Andreone 1 ; 1 Medical and Surgical Sciences, University<br />

of Bologna, Bologna, Italy; 2 OU Microbiology and Virology,<br />

Azienda Ospedaliera di Bologna, Bologna, Italy; 3 Scientific<br />

Affairs, Abbott Diagnostics, Rome, Italy<br />

Background and Aim: Treatment of chronic hepatitis C (CHC)<br />

with direct antiviral antigens (DAAs) achieve high rates of sustained<br />

virological response (SVR), accompanied by few side<br />

effects, short duration, and a simplified route of administration.<br />

The implementation of these therapies have to face the problems<br />

of high costs and their management, that includes the efficacy<br />

monitoring. While HCV-RNA measured by sensitive molecular<br />

amplification techniques represents the gold standard, hepatitis<br />

C core antigen (HCV-Ag) is emerging as an interesting new<br />

tool in diagnosis and treatment monitoring for CHC. However,<br />

its role in DAAs therapies has not been still evaluated. The aim<br />

of our ongoing study is to evaluate the accuracy of HCV-Ag<br />

measurement in CHC patients treated with different schedules<br />

of DAAs. Methods: To date, 27 patients (21 male, median<br />

age 55, range 25-78) have started interferon-free (IFN-f) DAAs<br />

treatment for 12 or 24 weeks, and 11 completed the therapy<br />

course. Sero-virological testing was carried out at baseline,<br />

after 1-2 weeks and every 4 weeks thereafter. HCV-RNA was<br />

quantified by real time PCR (Amplicor, Roche; limit of detection<br />

15 IU/mL); HCV-Ag was assessed by an automated chemiluminescent<br />

two-step immunoassay (Abbott HCV core antigen<br />

assay; limit of detection 3 fmol/L). Results: HCV-RNA and<br />

HCV-Ag were highly correlated, considering both only baseline<br />

values (Spearman r =0.93, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1131A<br />

mortality. Therefore it is crucial to early make the differential<br />

diagnosis between the two forms in order to plan the antiviral<br />

treatment.<br />

Disclosures:<br />

Xavier Forns - Consulting: Jansen, Abbvie; Grant/Research Support: Jansen,<br />

Gilead<br />

The following authors have nothing to disclose: Salvatore Stefano Sciarrone,<br />

Alberto Zanetto, Martina Gambato, Maria Alba Diaz, Alberto Ferrarese, Massimo<br />

Rugge, Marco Senzolo, Giacomo Germani, Francesco P. Russo, Patricia<br />

Llovet, Patrizia Burra<br />

1890<br />

Evaluation of the role of hepatic progenitor cells and<br />

cells resistant to apoptosis in prediction of disease<br />

progression and treatment response in patients with<br />

chronic hepatitis C<br />

Maha El Sabaawy 2 , Eman Abdelsameea 2 , Ayat Abdallah 3 , Ahmed<br />

El Refaie 1 , Mervat M. Soltan 1 , Nermine Ehsan 1 ; 1 Pathology, Natioanl<br />

Liver Institute, Menoufia, Egypt; 2 Hepatology, Natioanl Liver<br />

Institute, Menoufia, Egypt; 3 Community Medicine, Natioanl Liver<br />

Institute, Menoufia, Egypt<br />

Background: Recent <strong>studies</strong> have shown increase expression<br />

of hepatic progenitor cells and cells resistant to apoptosis in<br />

chronic HCV infection. However their role in predicting disease<br />

progression or treatment response was not clearly elucidated.<br />

Aim of this study: was to investigate the role of hepatic progenitor<br />

cells and cells resistant to apoptosis in disease severity and<br />

progression along with HCV treatment response by immunohistochemistry<br />

(IHC). Methods: This retrospective case control<br />

study was conducted on 10 healthy control subjects (donors<br />

for liver transplantation) and 91 chronic HCV patients who<br />

had completed combination treatment (interferon plus ribavirin).<br />

Demographic and clinical characteristics were collected<br />

from the data registry. Patients were categorized according<br />

to their SVR (sustained virological response) into two groups;<br />

responders and non-responders. Paraffin blocks from liver<br />

biopsies of control and patients included in this study were<br />

retrieved. Semi-thin sections were prepared on positive charge<br />

slides for immunostaining with CK7, Ki67 and bcl2 antibodies.<br />

Results: No specific age or sex prediction was reported<br />

in patients’ study group concerning any of the investigated<br />

markers. Control healthy subjects showed negative immunoreaction<br />

to hepatic progenitor cells either isolated or ductular,<br />

Ki67 and bcl2 lymphocyte associated portal tracts (LPT). One<br />

case of control group that revealed 15% steatosis had positive<br />

immunoreaction to bcl2 lymphocyte associated hepatic parenchyma<br />

(LAH). Laboratory data including transaminases (AST<br />

and ALT), platelets and prothrombin time exhibited significant<br />

relation with Ki67, CK7 both isolated and ductular and bcl2<br />

both LPT and LAH. CK7 ductular showed significant association<br />

with fibrosis and necroinflammatory activity according to Ishak<br />

score (P< 0.05) while non-significant relation was noticed in<br />

the CK7 isolated form and Ki67 (P> 0.05). Moreover bcl2<br />

both (LPT) and (LAH) demonstrated significant association with<br />

fibrosis and necroinflammatory activity (P< 0.05). Pertaining<br />

to SVR significant relation was confirmed with Ki67, CK7 both<br />

isolated and ductular and bcl2 (PT) and (LAH) (P< 0.05). Multivariate<br />

analysis revealed their role as predictors of HCV treatment<br />

response. Conclusion: Hepatic progenitor cells and cells<br />

resistant to apoptosis are significantly related to disease progression<br />

and could predict response of combination treatment<br />

in chronic HCV.<br />

Disclosures:<br />

The following authors have nothing to disclose: Maha El Sabaawy, Eman Abdelsameea,<br />

Ayat Abdallah, Ahmed El Refaie, Mervat M. Soltan, Nermine Ehsan<br />

1891<br />

Cascade of Care of HCV & HIV Infected Patients Identified<br />

Through Community Pop-Up Clinics (CPCs)<br />

Syune Hakobyan; Vancouver ID Research and Care Centre Society,<br />

Vancouver, BC, Canada<br />

AUTHORS: S. Hakobyan, S. Sharma, A. King, F. Zahedieh,<br />

H. Tossonian, B. Conway. BACKGROUND: The prevalence<br />

of Hepatitis C Virus (HCV) and Human Immunodeficiency<br />

Virus (HIV) infections is very high among the 18,000 men and<br />

women of Vancouver’s Downtown East Side (DTES). Despite<br />

the widespread availability of medical and non-medical services,<br />

there is little information regarding the cascade of care<br />

for those diagnosed with HIV and/or HCV in this neighborhood.<br />

The aim of this study was to document the cascade of<br />

HIV and HCV care in the DTES and identify obstacles to its<br />

improvement. METHODS: Participants were recruited at CPCs<br />

held at different community-based centers in the DTES frequented<br />

by people who inject drugs (PWID). During the CPCs,<br />

OraQuick HIV and HCV Rapid Antibody point of care testing<br />

was offered and participants were then asked to complete a<br />

targeted questionnaire while they waited for test results. Care<br />

for HIV immediately offered to those who were infected, while<br />

the questionnaire answers were used to identify parameters of<br />

engagement around HCV infection. RESULTS: From 03/13 to<br />

05/15, 1410 individuals were tested for HIV and HCV. Of<br />

these, 448 (31.8%) were found to be infected with HCV and<br />

29 (2.1%) co-infected with HIV. Within a linked multidisciplinary<br />

clinic, 18 HIV-infected individuals (62.1%) were linked<br />

to care with 9 (50%) achieving an undetectable viral load on<br />

treatment. In evaluating the feasibility of expanding the program<br />

to address HCV infection, 910 participants completed a<br />

questionnaire (22.9% female, 56.5% Caucasian, 28.5% First<br />

Nations; mean age: 46.3). Amongst all survey participants, 24<br />

were co-infected with HIV and HCV. In this population, 41.7%<br />

were Aboriginal, 75.0% had injected drugs, 29.2% shared<br />

needles and other injection equipment, 62.5% were previously<br />

incarcerated, and 75.0% were aware of a HCV cure, and<br />

87.5% stated they would consider treatment if they had HCV,<br />

although 54.2% did not believe they needed it. CONCLU-<br />

SIONS: Despite extensive prior testing for HIV on the DTES, we<br />

identified 29 individuals who were unaware or unengaged in<br />

care, and could arguably be considered as “core transmitters”<br />

of HIV and HCV. Their engagement in treatment for HIV has<br />

been successful in our model of care. There is significant interest<br />

in considering HCV treatment were it offered. CPCs are an<br />

important tool in initially engaging these individuals in medical<br />

care and moving towards increased HCV treatment uptake in<br />

this population.<br />

Disclosures:<br />

The following authors have nothing to disclose: Syune Hakobyan<br />

1892<br />

Correlates of Successful HCV Treatment in HIV Co-Infected<br />

Vulnerable Populations<br />

Brian Conway; Vancouver ID Research and Care Centre Society,<br />

Vancouver, BC, Canada<br />

AUTHORS: B. Conway, S. Hakobyan, A. King, F. Zahedieh,<br />

H. Tossonian, S. Sharma INTRODUCTION: Vulnerable populations,<br />

including people who inject drugs (PWID), are over-represented<br />

among HIV/HCV co-infected populations. Although<br />

clinical trial results suggest that many treatment regimens for<br />

HCV infection are equally effective in the setting of HIV co-infection,<br />

this has not been clearly established in populations<br />

consisting mainly of PWID and members of related vulnerable


1132A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

inner city groups. METHODS: We have established a multi-disciplinary<br />

outreach program to recruit and retain HIV/HCV-infected<br />

PWID in care. The program includes access to specialty<br />

medical care, extended social support services, complimentary<br />

over-the-counter medications and vitamins, daily snacks/beverages,<br />

and weekly meals and HIV/HCV support groups. HIV<br />

and HCV treatment is offered to all who qualify for it on medical<br />

grounds. We have conducted a retrospective analysis of<br />

all HIV co-infected patients treated for HCV infection within our<br />

program and for whom a definite HCV treatment outcome was<br />

ascertained. This analysis correlates SVR with a range of baseline<br />

demographic and clinical variables, including housing and<br />

active recreational drug use. RESULTS: Of 512 HIV-infected<br />

individuals in our program, 245 (47.9%) are co-infected with<br />

HCV, including 172 (70.2%) active (current/recent) PWID.<br />

In total, 75 (30.6%) have completed HCV treatment (5 on alloral<br />

regimens), including 55 (73.3%) active PWID, and 46<br />

(61.3%) with genotype 1 infection. The mean age was 51, 70<br />

(93.3%) male, 22 (26.7%) on opiate substitution, 72 (96.0%)<br />

on HIV treatment (62/72 with full virologic suppression), 20<br />

(27.8%) homeless, and 32 (42.7%) attending weekly HCV<br />

support groups. The SVR rate was 43.1% (31/72), 80% (4/5)<br />

in genotype 1 patients on all-oral regimens. SVR was higher<br />

in non-genotype 1 (54.5% vs. 43.1%) patients. Success rates<br />

were no higher in subjects on methadone 12 (54.5%), and no<br />

lower in the homeless 10 (50.0%) or active PWID 25 (45.5%).<br />

SVR was over 56% in active support group participants. An<br />

additional 9 co-infected patients have been started on all-oral<br />

regimens, and SVR data will be presented. DISCUSSION: In our<br />

centre, >35% of HIV/HCV co-infected patients (mostly PWID)<br />

have been treated for HCV infection. Many of the usual negative<br />

predictive factors do not affect treatment success, a fact<br />

that may be linked to the low-barrier multi-disciplinary nature<br />

of our program. There is a trend towards increased success in<br />

patients more actively engaged in our program. Our model will<br />

provide an opportunity for vulnerable populations to benefit<br />

from treatment for both HIV and HCV infections.<br />

Disclosures:<br />

Brian Conway - Advisory Committees or Review Panels: Vertex Pharmaceuticals,<br />

Merck, Boehringer Ingelheim, Jannsen Pharmaceuticals; Grant/Research Support:<br />

Vertex Pharmaceuticals, Merck, Boehringer Ingelheim, Jannsen Pharmaceuticals,<br />

AbbVie, Gilead Sciences, Gilead Sciences<br />

time intervals until postoperative recurrence (within 2 years,<br />

after 2 years or later, and no recurrence). (1) The maximum<br />

standardized uptake value (SUV max) of tumor was assessed<br />

and the best cut-off value of SUV max for predicting MVI was<br />

obtained using receiver operating characteristic (ROC) curve.<br />

Independent predictors for MVI were identified by multivariate<br />

analysis. (2) Using the best cut-off value of SUV max for predicting<br />

MVI, the associations between SUV max and recurrence<br />

patterns and intervals were analyzed. Results: (1) 31 of 79<br />

(39.2%) patients showed positive preoperative PET-scans and<br />

the mean SUV max was 3.15 ± 1.28 (range, 2−8.8). MVI<br />

was present in 13 of 79 patients (16.5%) on pathologic examination.<br />

SUV max ≥5.0 was suggested to be the best cut-off<br />

value for predicting MVI by ROC curve (area under the curve<br />

[AUC] =0.712, sensitivity=46.1%, specificity=90.6%). Multivariate<br />

analysis showed that SUV max ≥5.0 (p=0.05), AFP-<br />

L3 ≥15% (p=0.02), and non-simple nodular as macroscopic<br />

classification (p=0.039) were independent predictors for MVI.<br />

(2) SUV max in the recurrence within the MC (3.25±0.94)<br />

and no recurrence group (2.92±0.86) was significantly lower<br />

than that in the recurrence beyond the MC group (4.98±2.79).<br />

Multivariate analysis showed that SUV max ≥5.0 (p=0.002)<br />

was an independent predictor for recurrence beyond the MC.<br />

SUV max in the group with recurrence after 2 years or later<br />

(3.45±1.56) and no recurrence (2.92±0.86) was significantly<br />

lower than that in the within 2 years group (4.06±2.14). Multivariate<br />

analysis showed that SUV max ≥5.0 (p=0.002) and<br />

tumor size ≥30mm (p=0.026) were independent predictors for<br />

recurrence within 2 years. Conclusions: Preoperative 18 F-FDG<br />

PET-CT predicts MVI, recurrence beyond the MC, and recurrence<br />

within 2 years after curative HR for early-stage HCC.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Tomoki Kobayashi, Hiroshi<br />

Aikata, Fumi Shinohara, Norihito Nakano, Yuki Nakamura, Masahiro Hatooka,<br />

Kei Morio, Reona Morio, Takayuki Fukuhara, Yuuko Nagaoki, Tomokazu Kawaoka,<br />

Akira Hiramatsu, Michio Imamura, Yoshiiku Kawakami<br />

1893<br />

High 18 F-FDG uptake on preoperative PET-CT correlates<br />

with microvascular invasion and predicts early and<br />

massive recurrence after curative resection for early-stage<br />

hepatocellular carcinoma<br />

Tomoki Kobayashi, Hiroshi Aikata, Fumi Shinohara, Norihito<br />

Nakano, Yuki Nakamura, Masahiro Hatooka, Kei Morio, Reona<br />

Morio, Takayuki Fukuhara, Yuuko Nagaoki, Tomokazu Kawaoka,<br />

Akira Hiramatsu, Michio Imamura, Yoshiiku Kawakami, Kazuaki<br />

Chayama; Department of Gastroenterology and Metabolism, Hiroshima<br />

University Hospital, Hiroshima, Japan<br />

BackgroundAim: Microvascular invasion (MVI) predicts recurrence<br />

beyond the Milan criteria (MC) and early recurrence<br />

after hepatic resection (HR) for hepatocellular carcinoma<br />

(HCC). Recently, several <strong>studies</strong> have shown that 18 F-FDG<br />

uptake may be a surrogate marker for MVI. This study aimed to<br />

assess the value of preoperative 18 F-FDG PET-CT for predicting<br />

recurrence patterns and intervals after HR in early-stage HCC.<br />

Methods: 79 consecutive patients who were preoperatively<br />

examined with 18 F-FDG PET-CT and underwent curative HR for<br />

HCC within the MC were enrolled in this retrospective cohort<br />

study. They were classied according to the recurrence patterns<br />

(beyond the MC and within the MC or no recurrence) and the<br />

1894<br />

Reduced sensitivity of ultrasound in detecting hepatocellular<br />

carcinoma in obese patients, using explants<br />

pathology as gold standard<br />

Jamak Modaresi Esfeh 2 , Kaveh Hajifathalian 1 , Kianoush Ansari-Gilani<br />

3 , Ahyoung Kim 1 , Carlos J. Romero-Marrero 2 ; 1 Internal Medicine,<br />

Cleveland clinic, Beachwood, OH; 2 Gastroenterology and<br />

Hepatology, Cleveland Clinic, Cleveland, OH; 3 Diagnostic Radiology,<br />

University Hospitals, Case Medical Center, Cleveland, OH<br />

Background: American Association for the Study of Liver<br />

Diseases (AASLD) recommends screening for hepatocellular<br />

carcinoma (HCC) among patients with cirrhosis to be done<br />

with ultrasound (US) every six months, regardless of their BMI.<br />

We examined the sensitivity of US for diagnosis of HCC in<br />

obese patients. Methods: We used the prospective data from<br />

liver transplant patients between January 2005 and December<br />

2010, who were diagnosed with HCC in the pathologic examination<br />

of their explants. All patients had US <strong>studies</strong> of their<br />

liver within six months of diagnosis of HCC. Number and size<br />

of HCC lesions were extracted from radiologic and pathologic<br />

reports. Obesity was defined based on BMI≥30 kg/m 2 . Results:<br />

99 patients, 22% female, with mean BMI of 31 kg/m 2 (range<br />

20-43) were included. Most common underlying liver disease


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1133A<br />

was HCV (52%). At diagnosis, median number of HCC lesions<br />

was 1 (IQR 1-2) and median size of the largest lesion was 2 cm<br />

(IQR 1.5-3.1). Overall sensitivity of an US study for detection of<br />

HCC was 38% (95% CI 29-48%). Sensitivity was 77% (95% CI<br />

62-93%) in patients with BMI


1134A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1897<br />

Diagnostic Performance of Contrast-enhanced Imaging<br />

Techniques for Small Hepatocellular Carcinoma Measuring<br />

less than 2 cm<br />

Seong Hee Kang 1 , Jong Eun Yeon 1 , Young-Sun Lee 1 , Tae Suk<br />

Kim 1 , Yang Jae Yoo 1 , Sang Jun Suh 1 , Young Kul Jung 1 , Ji Hoon<br />

Kim 1 , Hyung Joon Yim 1 , Chang Hee Lee 2 , Kwan Soo Byun 1 ; 1 Gastroenterology<br />

& Hepatology, Korea University Medical Center,<br />

Seoul, Korea (the Republic of); 2 Radiology, Korea University Medical<br />

Center, Seoul, Korea (the Republic of)<br />

BACKGROUND: Recent guidelines recommend ultrasound follow-up<br />

for nodules (< 1 cm) discovered via ultrasonography<br />

surveillance (US) in liver cirrhosis patients. Nodules larger than<br />

1cm should be followed the recall policy with either dynamic<br />

contrast 4-phase CT scan or dynamic enhanced contrast MRI.<br />

There is no specified recall policy of nodule less than 1 cm<br />

found on US. We evaluated CT and MRI efficacy in hepatocellular<br />

carcinoma (HCC) diagnosis in nodules lesser than<br />

1cm and 1~2 cm. METHODS: We retrospectively analyzed<br />

130 ultrasonography-detected liver nodules


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1135A<br />

1899<br />

Efficacy of Delta-AFP in Predicting the Development of<br />

Hepatocellular Carcinoma in Cirrhotic Patients<br />

Ross Mund 1 , Yanhong Deng 2 , Maria Ciarleglio 2 , Tamar H. Taddei<br />

1,3 ; 1 Yale University School Of Medicine, New Haven, CT;<br />

2 Yale Center for Analytical Sciences, New Haven, CT; 3 VA Connecticut<br />

Healthcare System, West Haven, CT<br />

Background: Current AASLD practice guidelines recommend<br />

abdominal ultrasound every six months for hepatocellular carcinoma<br />

(HCC) screening in at risk individuals. Serum AFP was<br />

removed from guidelines in 2011 due to poor sensitivity. This<br />

study evaluated the change in AFP over time (delta AFP) as<br />

a predictor of HCC development in cirrhotic patients. Methodology:<br />

This retrospective cohort study included all cirrhotic<br />

patients without HCC from the VA Connecticut Healthcare System<br />

who had a baseline AFP and two subsequent AFP measurements<br />

during 2006-2011. AFP was measured at baseline, first<br />

serial AFP at an average of 320.88 days (std dev 103.13),<br />

and last serial AFP at an average of 605.75 days (std dev<br />

122.21) from the baseline. The absence of HCC was defined<br />

by absence of a focal hepatic mass consistent with HCC on<br />

dynamic imaging from the time of baseline AFP to the last measurement.<br />

Subjects were classified according to whether they<br />

had hepatitis C (HCV) cirrhosis or not. Longitudinal analysis<br />

with patient level random effects were used to model change<br />

in AFP between those who developed and did not develop<br />

HCC. Logistic regression analysis was used to examine the link<br />

between delta AFP and HCC. Results: 169 subjects met criteria<br />

for the study. N=91 for HCV cirrhosis (53.85%). 17 subjects<br />

developed HCC (10.06%). Delta AFP was assessed between<br />

the two groups. There was significant difference in delta AFP<br />

from baseline to the first serial measurement between HCC-N<br />

and HCC-Y groups (+0.2 and +8.75, respectively, p=0.0362,<br />

95% CI: 0.56-16.55), and from the baseline to the last AFP<br />

measurement (+1.72 and +85.73, respectively, p +20 were 5.342 times more likely<br />

to develop HCC as compared to those with delta AFP < +20<br />

(p= 0.0067; 95% CI: 1.593-17.919). HCV as a moderator<br />

could not be assessed due to limited sample size. Conclusion:<br />

This study revealed statistically significant differences in delta<br />

AFP in cirrhotic subjects who developed HCC versus those who<br />

did not develop HCC. Importantly, this difference in delta AFP<br />

was seen at the first serial AFP value, an average of 284.87<br />

days prior to the development of HCC. Additionally, the risk of<br />

developing HCC for subjects with a delta AFP > +20 over a<br />

2 year span was significantly increased. Due to small sample<br />

size and wide confidence intervals, this study should be validated<br />

in a larger, prospective cohort. This study supports the<br />

hypothesis that AFP may be a cost effective, readily available<br />

and early biomarker for the diagnosis of HCC, once we better<br />

understand thresholds and changes that may increase its sensitivity.<br />

Disclosures:<br />

Tamar H. Taddei - Consulting: Onyx Pharmaceuticals; Grant/Research Support:<br />

Bayer HealthCare<br />

The following authors have nothing to disclose: Ross Mund, Yanhong Deng,<br />

Maria Ciarleglio<br />

1900<br />

A mass spectrometry based serum test for the detection<br />

of hepatocellular carcinoma (HCC) in high risk patients<br />

Devalingam Mahalingam 1 , Julio A. Gutierrez 1 , W. Kenneth Washburn<br />

1 , Glenn A. Halff 1 , Leonidas Chelis 2 , Stylianos Kakolyris 2 ,<br />

Stylianos Vradelis 3 , Julia Grigorieva 4 , Carlos Oliveira 4 , Heinrich<br />

Roder 4 , Joanna Roder 4 ; 1 Cancer Therapy and Research Center,<br />

University of Texas Health Sciences Center San Antonio, San Antonio,<br />

TX; 2 Department of Medical Oncology, Democritus University<br />

of Thrace, Alexandroupolis, Greece; 3 2nd Department of Internal<br />

Medicine, Democritus University of Thrace, Alexandroupolis,<br />

Greece; 4 Biodesix, Boulder, CO<br />

Improved screening protocols (SP) for patients at high risk of<br />

developing HCC could lead to improved patient outcome and<br />

possibly cure if detected early. The current SPs, ultrasound with<br />

the possible addition of alpha-fetoprotein (AFP) measurement,<br />

suffer from insufficient sensitivity and specificity, and less than<br />

30% of patients are diagnosed early enough to be suitable<br />

candidates for resection or transplantation. A better biomarker<br />

for the early detection of HCC is clearly needed. We used<br />

mass spectrometry analysis of serum samples combined with<br />

specialized deep learning techniques to train and validate such<br />

a test. Samples were available from two patient cohorts. The<br />

first was of patients undergoing transplant or resection for HCC<br />

(N=48; median MELD = 14), and patients with advanced cirrhosis<br />

undergoing liver transplant (N =52; median MELD =<br />

25); underlying cause of liver disease was 49% hepatitis C<br />

infection (HCV). The second cohort consisted of patients with<br />

advanced HCC (N= 110; 72/27/11 Child-Pugh A/B/C) and<br />

liver disease but no HCC (N= 83; 74/7/2 Child-Pugh A/B/C),<br />

with 68% of patients having hepatitis B infection (HBV). The<br />

two cohorts were combined to create a more representative set<br />

of patients with and at risk of HCC, which was split into development<br />

and validation sets, stratified by cause and severity<br />

of liver disease and stage of HCC. The classifier was trained<br />

especially to avoid confounding by liver function. The resulting<br />

test showed a sensitivity/specificity of 83%/84% in cross<br />

validation in the development set, with 73% of patients with<br />

HCC who could be treated by resection or transplant (BCLC<br />

A) correctly identified. Performance was good across all represented<br />

causes of liver disease. Validation set performance was<br />

similar with sensitivity/specificity of 81%/79%, an accuracy of<br />

64% for patients with BCLC A HCC, and little dependence on<br />

cause of liver disease (overall accuracy: 91% HBV, 79% HCV).<br />

Within the validation set, when compared with AFP alone,<br />

the test had 22% greater specificity at fixed sensitivity and<br />

12% greater sensitivity at fixed specificity. Results from further<br />

ongoing validation of the test on independent patient cohorts<br />

will also be presented. The developed test shows promise in<br />

the early detection of HCC as a screening tool in high risk<br />

patients independent of the cause of underlying liver disease.<br />

While further validation will be necessary to conclusively prove<br />

its clinical validity, we believe such a test could substantially<br />

improve early detection of HCC.<br />

Disclosures:<br />

Glenn A. Halff - Board Membership: texas organ sharing alliance, texas organ<br />

sharing alliance, texas organ sharing alliance, texas organ sharing alliance;<br />

Stock Shareholder: xenex, xenex, xenex, xenex<br />

Julia Grigorieva - Stock Shareholder: Biodesix<br />

Carlos Oliveira - Employment: Biodesix; Patent Held/Filed: Biodesix<br />

Heinrich Roder - Employment: Biodesix<br />

Joanna Roder - Employment: Biodesix; Patent Held/Filed: Biodesix; Stock Shareholder:<br />

Biodesix<br />

The following authors have nothing to disclose: Devalingam Mahalingam, Julio<br />

A. Gutierrez, W. Kenneth Washburn, Leonidas Chelis, Stylianos Kakolyris, Stylianos<br />

Vradelis


1136A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1901<br />

ARID1A mutations may represent a novel pathway of<br />

carcinogenesis in biliary carcinomas<br />

Motoko Sasaki 1 , Takeo Nitta 2 , Yasunori Sato 1 , Yasuni<br />

Nakanuma 3 ; 1 Human Pathology, Kanazawa University Graduate<br />

School of Medicine, Kanazawa, Japan; 2 Gastroenterological<br />

Surgery II, Hokkaido University Graduate School of Medicine,<br />

Sapporo, Japan; 3 Division of Pathology, Shizuoka Cancer Center,<br />

Shizuoka, Japan<br />

Backgrounds/Aims. We have reported that some cholangiocarcinoma<br />

arise following a stepwise pathway of carcinogenesis<br />

through flat and papillary premalignant lesions, biliary intraepithelial<br />

neoplasia (BilIN) and intraductal papillary neoplasm of<br />

the bile duct (IPNB) and KRAS mutations and GNAS mutations<br />

may be involved in this pathway. Recent <strong>studies</strong> using exome<br />

sequencing and next generation sequencing revealed frequent<br />

alterations in several new genes including the three chromatin-remodeling<br />

genes (BAP1, ARID1A or PBRM) in cholangiocarcinomas.<br />

Given frequent inactivating mutations in ARID1A<br />

gene coding BAF250a protein in intrahepatic cholangiocarcinoma<br />

and a reported usefulness of ARID1A/BAF250a immunostaining<br />

to detect ARID1A mutation, this study investigates<br />

immunohistochemically the clinicopathological significance of<br />

ARID1A mutations in biliary carcinomas. Methods. We examined<br />

the status of inactivating mutations in ARID1A by immunohistochemistry<br />

and the relationship with clinicopathological<br />

features in 13 patients with combined hepatocellular-cholangiocarcinoma<br />

(cHC-CC, M/F=9/4), 49 patients with intrahepatic<br />

cholangiocarcinoma (ICC, M/F=22/27), 17 with IPNB (M/<br />

F=11/6), 72 with extrahepatic cholangiocarcinoma (EHCC,<br />

M/F=51/21) and 43 with gallbladder carcinoma (GBC, M/<br />

F=18/25). Fifteen BilIN lesions were also examined. Results.<br />

Complete loss or clonal loss of ARID1A immunoreactivity indicating<br />

the inactivating mutations of ARID1A was detected in<br />

one (7.7%) of cHC-CC, 9 (18.4%) of ICC, none of IPNB, 11<br />

(15.3%) of EHCC and 4 (9.1%) of GBC. The patients with<br />

biliary carcinoma with ARID1A mutations were significantly<br />

female-predominant (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1137A<br />

cirrhosis etiology in cirrhotic patients undergoing screening<br />

for HCC. Methods: This was a case-control study conducted<br />

on patients with cirrhosis with and without HCC identified via<br />

ICD-9 diagnosis codes retrospectively from 1990-2014 in a<br />

large state-wide database who underwent surveillance imaging.<br />

HCC was diagnosed by imaging and/or biopsy. The cirrhosis<br />

cohort without HCC was determined by absence of HCC<br />

on last imaging study. AFP levels within 90 days of imaging<br />

were abstracted. The sensitivity and specificity of AFP levels<br />

was determined for HCC diagnosis in HCV cirrhosis, non-viral<br />

cirrhosis and the entire cohort. AFP thresholds were explored<br />

based on previous literature by ROC analysis and group comparisons<br />

were performed by Chi-square tests. Results: 10,481<br />

patients with cirrhosis were identified, 6166 patients excluded<br />

(no AFP levels, no AFP levels within 90 days, or AFP >500).<br />

1641 patients with HCC were identified, 402 excluded as<br />

HCC diagnosed before first imaging study. 3371 patients with<br />

cirrhosis and 551 HCC patients were analyzed. Patients were<br />

predominantly white (87 % without HCC, 86 % with HCC),<br />

and male (59% cirrhotic cohort, 72% HCC cohort). Patients<br />

with cirrhosis were younger (54 vs 57 years p


1138A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1905<br />

WITHDRAWN<br />

1906<br />

LI-RADS hepatocellular carcinoma diagnostic classification<br />

system: utilization by community radiologists, and<br />

results of second opinion reading by staff radiologists at<br />

a transplant center<br />

Jesse M. Civan 1 , Aaron Martin 2 , Raza Hasan 2 , Flavius Guglielmo<br />

3 , Sandeep Deshmukh 3 , Christopher G. Roth 3 , Donald G.<br />

Mitchell 3 ; 1 Gastroenterology & Hepatology, Thomas Jefferson University<br />

Hospital, Philadelphia, PA; 2 Internal Medicine, Thomas Jefferson<br />

University, Philadelphia, PA; 3 Radiology, Thomas Jefferson<br />

University, Philadelphia, PA<br />

Background: LI-RADS (Liver Imaging Reporting and Data System)<br />

was created to clarify radiographic degree of concern<br />

for hepatocellular carcinoma (HCC). In 2013, UNOS adopted<br />

criteria analogous to LI-RADS category 5 (definite HCC) for<br />

lesions to qualify for HCC MELD exception. We evaluated<br />

LI-RADS utilization by community radiologists, inter-operator<br />

agreement of LI-RADS scoring, and prediction of LI-RADS classification<br />

by descriptive language. Methods: We reviewed 150<br />

outside abdominal MRI and CT <strong>studies</strong> that were overread<br />

(formal consultation) by radiology specialists at a transplant<br />

center. Studies performed at other academic centers were<br />

excluded. We abstracted qualitative phrases used by outside<br />

radiologists in lieu of LI-RADS to convey degree of concern<br />

for HCC. Statistical analysis was performed using SPSS v23.<br />

Results: LI-RADS classification was provided in 13% of cases by<br />

outside radiologists and 88% of internal re-reads. For outside<br />

reports utilizing LI-RADS, correlation with our LI-RADS classification<br />

was poor, Spearman 0.00 (p=0.99) for combined<br />

LI-RADS 4/5, Spearman 0.07 (p=0.82) for LI-RADS 5. For outside<br />

reports not utilizing LI-RADS, 16 distinct qualitative phrases<br />

were used to convey degree of concern for HCC, many of<br />

which did not map clearly to a LI-RADS category (e.g. “in<br />

keeping with HCC”, “could represent HCC”). Overall, descriptive<br />

terms did not predict a combined outcome of the presence<br />

of LI-RADS 4 or 5 lesions (Cramer’s V=0.48, p=0.27),<br />

or presence of LI-RADS 5 lesions alone (Cramer’s V=0.56,<br />

p=0.06). Correlation between individual descriptive phrases<br />

and LI-RADS classification is depicted in figure 1. Conclusion:<br />

The LI-RADS classification system is under-utilized in the community<br />

setting. Qualitative language used in outside radiology<br />

reports poorly predicted diagnosis of probable or definite HCC<br />

by LI-RADS criteria.<br />

Disclosures:<br />

Jesse M. Civan - Consulting: Merck<br />

The following authors have nothing to disclose: Aaron Martin, Raza Hasan,<br />

Flavius Guglielmo, Sandeep Deshmukh, Christopher G. Roth, Donald G. Mitchell<br />

1907<br />

A Model to Estimate Survival in Ambulatory Patients<br />

with Hepatocellular Carcinoma: Can It Predict the Natural<br />

Course of Hepatocellular Carcinoma?<br />

Won-Mook Choi 1 , Su Jong Yu 2 , Young Youn Cho 2 , Eun Ju Cho 2 ,<br />

Jeong-Hoon Lee 2 , Yoon Jun Kim 2 , Jung-Hwan Yoon 2 , June Sung<br />

Lee 3 ; 1 Graduate School of Medical Science and Engineering,<br />

Korea Advanced Institute of Science and Technology(KAIST), Daejeon,<br />

Korea (the Republic of); 2 Department of Internal Medicine<br />

and Liver Research Institute, Seoul National University College of<br />

Medicine, Seoul, Korea (the Republic of); 3 Department of Internal<br />

Medicine, Ilsan Paik Hospital, Inje University College of Medicine,<br />

Goyang, Korea (the Republic of)<br />

Objectives: Several hepatocellular carcinoma (HCC) staging<br />

systems are available; however, whether these staging systems<br />

could predict the natural course of HCC is largely unknown.<br />

This study aimed to investigate whether the various HCC staging<br />

systems could reflect the natural course of HCC. Methods:<br />

1116 patients with history of HCC treatment and 111 patients<br />

without any history of treatment till death or last follow-up at a<br />

single tertiary hospital were included. To minimize selection<br />

bias, patients with treatment were matched using propensity-score<br />

matching at a 1:1 ratio with patients without treatment.<br />

The model’s performance was assessed and compared to other<br />

staging systems using C-statistics and Akaike information criterion<br />

(AIC). Results: In the group of treated patients before<br />

propensity score matching, the Model to Estimate Survival<br />

in Ambulatory HCC patients (MESIAH) score showed higher<br />

predictive performance, with a C-statistic of 0.832 (95% confidence<br />

interval [CI], 0.808-0.856), when compared to the<br />

Barcelona Clinic Liver Cancer staging system and the seventh<br />

edition of the American Joint Committee on Cancer TNM staging<br />

system. However, in the group of untreated patients, all<br />

staging systems including the MESIAH score failed to predict<br />

survival. After propensity score matching, although the MESIAH<br />

score was balanced between the two groups (5.89±1.35 and<br />

5.94±1.45, P=0.60), the treated group survived longer than<br />

the untreated group, which suggests that treatment by itself did<br />

prolong survival. Moreover, even the MESIAH score failed to<br />

predict outcome of not only the untreated group but also the<br />

treated group after propensity score matching. Conclusions:<br />

Although the MESIAH score provided better prognostic stratification<br />

than other staging systems, it also was not helpful<br />

in predicting the natural course of HCC. Since the preferred<br />

treatment modality is varied by multiple factors, it is necessary<br />

to develop a new staging system that reflects the natural course<br />

of HCC regardless of treatment.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1139A<br />

Fig 1. Observed and predicted overall survival curves stratified<br />

by the MESIAH score (A) with treated groups, (B) with untreated<br />

groups before propensity score matching<br />

seem to be a good option for the subset of patients with symptomatic<br />

severe portal hypertension.<br />

Disclosures:<br />

The following authors have nothing to disclose: Vincent Soriano, Pablo Labarga,<br />

Jose V. Fernandez-Montero, Carmen de Mendoza, Pablo Barreiro<br />

Disclosures:<br />

Yoon Jun Kim - Grant/Research Support: Bristol-Myers Squibb, Roche, JW Creagene,<br />

Bukwang Pharmaceuticals, Handok Pharmaceuticals, Hanmi Pharmaceuticals,<br />

Yuhan Pharmaceuticals; Speaking and Teaching: Bayer HealthCare<br />

Pharmaceuticals, Gilead Science, MSD Korea, Yuhan Pharmaceuticals, Samil<br />

Pharmaceuticals, CJ Pharmaceuticals, Bukwang Pharmaceuticals, Handok Pharmaceuticals<br />

The following authors have nothing to disclose: Won-Mook Choi, Su Jong Yu,<br />

Young Youn Cho, Eun Ju Cho, Jeong-Hoon Lee, Jung-Hwan Yoon, June Sung Lee<br />

1908<br />

Long-term Outcome of HIV+ Patients with Non-Cirrhotic<br />

Portal Hypertension Upon Discontinuation of Didanosine<br />

Vincent Soriano 1 , Pablo Labarga 4 , Jose V. Fernandez-Montero 2 ,<br />

Carmen de Mendoza 3 , Pablo Barreiro 1 ; 1 La Paz University Hospital<br />

& IdiPAZ, Madrid, Spain; 2 University Hospital Crosshouse,<br />

Kilmarnock, United Kingdom; 3 Puerta de Hierro Research Institute,<br />

Majadahonda, Spain; 4 La Luz Clinic, Madrid, Spain<br />

Background: Exposure to the antiretroviral drug didanosine has<br />

been associated to hepatic vascular damage and non-cirrhotic<br />

portal hypertension (NCPH) in HIV+ individuals. Life-threatening<br />

episodes of variceal bleeding and acute portal thrombosis<br />

are the most feared complications. Herein, we describe the<br />

long-term outcome of a series of patients that developed HIV-associated<br />

NCPH. Methods: All HIV+ patients followed at one<br />

single HIV reference large clinic in Madrid were examined.<br />

Diagnosis of NCPH was made based on recognition of portal<br />

hypertension in the absence of cirrhosis (confirmed either by<br />

biopsy and/or elastometry), and exclusion of any known etiology<br />

other than didanosine exposure. Following removal of<br />

the drug, patients have been on regular follow-up using other<br />

antiretroviral drugs. Results: From a total of 3,200 HIV+ individuals<br />

attended during the last decade, a total of 21 (0.6%)<br />

were diagnosed with didanosine-associated NCPH. Characteristics<br />

features of small portal venulopathy were found in the<br />

liver biopsy of all 12 patients with histologically available specimens.<br />

At diagnosis, mean age was 43-years and 17 were<br />

male. No deaths due to liver-related deaths have occurred<br />

after an average follow-up of 8.3 years. However, two thirds<br />

developed serious liver-related complications early on, including<br />

portal thrombosis (7), ascites (8), variceal bleeding (6) and<br />

encephalopathy (2). Two subjects underwent TIPPS and one<br />

surgical porto-cava shunt due to severe portal hypertension.<br />

Interestingly, none of these 3 patients developed encephalopathy<br />

thereafter, most likely due to their relative well-preserved<br />

hepatic function. No further liver-related complications occurred<br />

beyond 2 years of didanosine removal in our series except for<br />

one patient that suffered variceal re-bleeding 7 years later.<br />

Conclusions: NCPH in HIV+ patients is a rare but potentially<br />

life-threatening condition linked to didanosine exposure that<br />

tends to ameliorate over time after stopping didanosine. TIPPS<br />

1909<br />

AMPK activation prevents and reverses drug-induced<br />

mitochondrial damage in hepatocytes by regulating<br />

mitochondrial quality control<br />

Dong Fu 1 , Ghada Haydar 1 , Sun Woo Sophie Kang 1 , Geoffrey C.<br />

Farrell 2 , Jennifer Lippincott-Schwartz 3 , Irwin M. Arias 3 ; 1 Faculty<br />

of Pharmacy, The University of Sydney, Sydney, NSW, Australia;<br />

2 Liver Research Group, National University of Australia Medical<br />

School, Canberra, ACT, Australia; 3 NICHD, National Institutes of<br />

Health, Bethesda, MD<br />

Mitochondrial damage plays a central role in drug-induced<br />

liver injury (DILI) which is responsible for many cases of acute<br />

liver failure, and pre- or post-market withdrawal of drugs. Mitochondria<br />

maintain function through quality control that includes<br />

regulated fusion/fission dynamics and autophagy-mediated<br />

degradation that eliminates damaged mitochondria. Therefore,<br />

enhancement of mitochondrial quality control is a potential<br />

approach for treatment of DILI. AMP activated kinase (AMPK)<br />

is a cellular energy sensor which, when activated by phosphorylation,<br />

promotes mitochondrial biogenesis and activates<br />

autophagy. Using collagen sandwich cultures of rat and human<br />

hepatocytes, we investigated the role of AMPK in mitochondrial<br />

quality control, and prevention and reversal of drug-induced<br />

mitochondrial and hepatocellular damage. Results: Western<br />

blot and immunofluorescence results demonstrated that hepatotoxic<br />

drugs, acetaminophen (for intrinsic DILI) and diclofenac<br />

(for idiosyncratic DILI), significantly decreased expression<br />

of mitochondrial fusion protein Mfn1; acetaminophen also<br />

decreased expression of fusion proteins Mfn2 and Opa1. Both<br />

drugs caused mitochondrial fragmentation, and decreased ATP<br />

levels and mitochondrial membrane potential in human and<br />

rat hepatocytes, revealing that both drugs cause mitochondrial<br />

damage. Moreover, both drugs significantly decreased cell viability<br />

and caused depolarization in human and rat hepatocytes.<br />

Activation of AMPK, by simultaneous administration of hepatotoxic<br />

drugs and AICAR, a specific AMPK activator, or addition<br />

of AICAR at 5hr post exposure of both drugs when hepatocellular<br />

damage occurred, restored mitochondrial function and<br />

fusion, hepatocyte polarization and cell viability, revealing<br />

that activation of AMPK prevents and reverses drug-induced<br />

mitochondrial and hepatocellular damage. AMPK activation<br />

prevented drug-mediated decreases in Mfn1, 2 and Opa1,<br />

indicating AMPK induced mitochondrial fusion by regulating<br />

fusion machinery. AMPK activation also increased autophagy<br />

in the presence of acetaminophen but not diclofenac, suggesting<br />

that the autophagic effect of AMPK plays a differential role<br />

in prevention of drug-induced hepatotoxicity. Conclusion: Acetaminophen<br />

and diclofenac fragment mitochondria through inhibition<br />

of the fusion machinery, and cause mitochondrial and<br />

hepatocellular damage. Through promotion of mitochondrial<br />

fusion or together with activation of autophagy, two important<br />

processes of mitochondrial quality control, activation of AMPK<br />

prevents and reverses drug-induced mitochondrial and hepatocellular<br />

damage. Activation of AMPK may provide a novel<br />

strategy for treatment of DILI.<br />

Disclosures:<br />

The following authors have nothing to disclose: Dong Fu, Ghada Haydar, Sun<br />

Woo Sophie Kang, Geoffrey C. Farrell, Jennifer Lippincott-Schwartz, Irwin M.<br />

Arias


1140A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1910<br />

The role of CD36 in acetaminophen-induced hepatotoxicity<br />

in mice<br />

chen zhang 1 , Xianmin Mu 1 , Che Xu 1 , Shi Hu 1 , Yuanfang Lu 1 ,<br />

Qiang You 1,2 ; 1 Department of Biotherapy, Second Affiliated Hospital,<br />

Nanjing Medical University, Nanjing, China; 2 Department of<br />

Immunology, Nanjing Medical University, Nanjing, China<br />

Acetaminophen (APAP) overdose cause hepatotoxicity, and<br />

even acute liver failure, which involves a series of critical events<br />

including APAP metabolite protein adduct formation, mitochondrial<br />

dysfunction, oxidant stress, peroxynitrite formation and<br />

nuclear DNA fragmentation. Recent <strong>studies</strong> indicate that sterile<br />

inflammation and innate immune cells may play important roles<br />

in APAP-induced liver injury and repair. CD36 is a member of<br />

the class B scavenger receptor family of cell surface proteins. It<br />

binds a variety of ligands including collagen, thrombospondin,<br />

erythrocytes parasitized with Plasmodium falciparum, oxidized<br />

low density lipoprotein, native lipoproteins, oxidized phospholipids,<br />

and long-chain fatty acids. CD36 also assembles<br />

with Toll-like receptor 4 and 6 to promote sterile inflammation.<br />

The present study aims to investigate the role of CD36<br />

in APAP induced murine liver injury. Methods: WT C57BL/6J<br />

and CD36 -/- mice on same background were intraperitoneally<br />

injected with APAP (300mg/kg) after fasting for 16h. The<br />

blood and liver tissues were collected at 8h and 24h after<br />

treatment. Liver injury was evaluated by serum ALT level and<br />

liver tissue H&E staining. Liver inflammatory factors expression<br />

was determined by real time PCR. The metabolite of APAP,<br />

N-acetyl-p-benzoquinone imine (NAPQI) protein adducts, and<br />

Cytochrome P450 2E1 (CYP2E1) level were measured by<br />

Western blot. Liver infiltrating macrophages and neutrophils<br />

were characterized by flow cytometry. Results: Compared with<br />

WT mice, APAP treated CD36 -/- mice show less liver injury indicated<br />

by serum ALT level and liver tissue H&E staining. There<br />

was no significant difference in NAPQI-protein adducts and<br />

CYP2E1 expression between these two strains. However, less<br />

IL-1β and CXCL1 mRNA expression were observed in APAP<br />

treated CD36 -/- mice as well as infiltrating macrophages and<br />

neutrophils. Conclusion: APAP treated CD36 -/- mice have less<br />

liver damage than WT mice which correlates with the extent of<br />

inflammation. CD36 could be a new target to reduce APAP-induced<br />

liver injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: chen zhang, Xianmin Mu, Che<br />

Xu, Shi Hu, Yuanfang Lu, Qiang You<br />

1911<br />

Withaferin-A Reduces Acetaminophen-Induced Liver<br />

Injury in Mice<br />

Ravirajsinh Jadeja 1 , Nathalie H. Urrunaga 2 , Suchismita Dash 2 ,<br />

Neeraj K. Saxena 2 , Sandeep Khurana 1 ; 1 Gastroenterology and<br />

Hepatology, Georgia Regents University, Augusta, MD; 2 Gastroenterology<br />

and Hepatology, University of Maryland School of Medicine,<br />

Baltimore, MD<br />

Acetaminophen (APAP) overdose induces oxidative stress that<br />

leads to hepatocyte necrosis. Withaferin-A (WA), a lactone,<br />

has anti-oxidant activities however, its therapeutic potential<br />

in APAP hepatotoxicity is unknown. The aim of our study was<br />

to assess the therapeutic potential of WA in a mouse model<br />

that mimics APAP-induced liver injury (AILI) in humans. Overnight<br />

fasted C57BL/6NTac male mice (56 wk. old) received<br />

200 mg/kg APAP intraperitoneally (i.p.). An hour later mice<br />

were treated with 40 mg/kg WA or vehicle i.p., and euthanized<br />

at 4 and 16 h; their livers were harvested and serum collected<br />

for analysis. At 4 h, compared to vehicle-treated mice,<br />

WA-treated mice had reduced serum ALT levels (P < 0.05),<br />

hepatocyte necrosis (P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1141A<br />

bumin (HNA). Out of the 354 subjects, 68 were analyzed for<br />

HSA functional changes. We examined the radical scavenging<br />

activity of purified HSA by monitoring the absorbance of<br />

diphenylpicrylhydrazyl (DPPH) radicals and the affinity of the<br />

purified HSA using ketoprofen. Analysis of variance (ANOVA)<br />

followed by the Tukey’s method, Jonckheere-Terpstra (JT) test,<br />

receiver operating characteristic curves (ROC), and Pearson’s<br />

correlation were used for statistical analysis. Results: The radical<br />

scavenging ability and binding capacity of HSA significantly<br />

deteriorated with the severity of the disease (P


1142A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

*Median [Min, Max]; ** P-values from Fisher’s Exact for categorical<br />

variables, Wilcoxon-Mann-Whitney for continuous variables<br />

Disclosures:<br />

Dean W. Roberts - Management Position: Acetaminophen Toxicity Diagnostics,<br />

LLC<br />

William M. Lee - Consulting: Eli Lilly, Sanofi; Grant/Research Support: Gilead,<br />

BMS, Vertex, Merck<br />

Laura James - Grant/Research Support: NIDDK; Management Position: Acetaminophen<br />

Toxicity Diagnostics, LLC<br />

The following authors have nothing to disclose: Jack A. Hinson, Shasha Bai, R.<br />

Todd Stravitz, Adrian Reuben, Christopher J. Swearingen, Pippa Simpson<br />

1915<br />

Understanding the Relationship between Systemic and<br />

Hepatic Exposure of Obeticholic Acid for the Treatment<br />

of Liver Disease in Patients with Cirrhosis<br />

Jeffrey Edwards 1 , Carl LaCerte 1 , Nathalie H. Gosselin 2 , Thomas<br />

Peyret 2 , J.F. Marier 2 , Alan F. Hofmann 3 , David Shapiro 1 ; 1 Intercept<br />

Pharmaceuticals, San Diego, CA; 2 Pharsight, a Certara Company,<br />

Princeton, NJ; 3 Department of Medicine, University of California<br />

San Diego, San Diego, CA<br />

Obeticholic acid (OCA) is a selective and potent farnesoid<br />

X receptor (FXR) agonist in development for the treatment of<br />

primary biliary cirrhosis (PBC) and other liver diseases. OCA<br />

is structurally similar to chenodeoxycholic acid (CDCA) and<br />

has similar pharmacokinetics (PK), including conjugation and<br />

enterohepatic circulation. The amidates of OCA activate FXR<br />

in the intestine and liver, leading to a reduced hepatic bile<br />

acid pool. Patients with cirrhosis have higher systemic exposure<br />

of bile acids (x18) while hepatic exposure of bile acids is<br />

only ~2-fold higher (Fischer et al. Clinica Chimica Acta 1996<br />

251:173). In OCA-treated patients, systemic OCA exposure<br />

was significantly higher in those with moderate (x4) and severe<br />

(x17) hepatic impairment, similar to increases in systemic exposure<br />

of endogenous bile acids. FXR activation (measured by<br />

fibroblast growth factor-19) was similar in patients with hepatic<br />

impairment versus healthy volunteers, consistent with similar<br />

liver and intestinal OCA exposure. Thus systemic exposure of<br />

OCA is not reflective of hepatic concentrations of OCA, the<br />

primary site of action for safety and efficacy. To understand the<br />

relationship between systemic and hepatic exposure of OCA<br />

and its pharmacologically active conjugates in patients with<br />

and without hepatic (cirrhosis) impairment, a physiologic PK<br />

model of OCA was developed. The PK model for OCA was<br />

based on a multi-compartment PK model for CDCA (Molino et<br />

al. Eur J Clin Invest 1986 16:397) and consisted of systemic,<br />

hepatobiliary, and intestinal systems. The model accounted<br />

for the pathophysiology associated with cirrhosis including<br />

decreased hepatic uptake of OCA, portal-systemic shunting,<br />

decreased physiologic/functional liver volume, and differences<br />

in amino acid conjugation. The PK model was used to simulate<br />

the overall mean concentration-time profile of total OCA in<br />

plasma and liver. There was good agreement between predicted<br />

and observed increases in systemic exposure of total<br />

OCA (the sum of OCA and its conjugates) after a single 10 mg<br />

dose in subjects with mild (x1.4), moderate (x8), and severe<br />

(x13) hepatic impairment relative to healthy subjects. The predicted<br />

increases in liver exposure for subjects with mild, moderate,<br />

and severe hepatic impairment were only 1.1-, 1.5-,<br />

and 1.7-fold, respectively, relative to healthy participants and<br />

are similar to the hepatic exposure of bile acids observed in<br />

patients with cirrhosis. Model-predicted hepatic exposure of<br />

OCA, consistent with clinical study results, support the safety<br />

and efficacy of OCA in PBC patients with cirrhosis at the therapeutic<br />

doses used in non-cirrhotic patients.<br />

Disclosures:<br />

Jeffrey Edwards - Employment: Intercept Pharmaceuticals; Stock Shareholder:<br />

Intercept Pharmaceuticals<br />

Carl LaCerte - Employment: Amylin Pharmaceuticals; Stock Shareholder: Amylin<br />

Pharmaceuticals<br />

Nathalie H. Gosselin - Employment: Pharsight<br />

Thomas Peyret - Employment: Pharsight<br />

Alan F. Hofmann - Board Membership: Vital Therapies, Vital Therapies; Consulting:<br />

Albireo Pharma, Shire, Intercept Pharma, Relypsa; Stock Shareholder:<br />

Intercept Pharma<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: J.F. Marier<br />

1916<br />

Transcriptional Profiling Of Primary Human Hepatocytes<br />

Under Hemodynamics And Transport Differentiates Distinct<br />

Mechanisms Of Drug-Induced Liver Injury.<br />

Robert Figler 1 , Brett R. Blackman 1 , Charles Qualls 2 , Svetlana<br />

Marukian 1 , Sol Collado 1 , Mark Lawson 1 , Aaron J. Mackey 1 ,<br />

David Manka 1 , Brian R. Wamhoff 1 , Ajit Dash 1 ; 1 HemoShear Therapeutics,<br />

Charlottesville, VA; 2 Qualls Preclinical Solutions LLC,<br />

Thousand Oaks, CA<br />

Drug induced liver injury (DILI) is a major cause of liver failure,<br />

drug withdrawal and non-approval. Predicting DILI is challenging<br />

since drug concentrations used in in vitro hepatocyte<br />

cultures are often significantly higher than therapeutic plasma<br />

concentrations, making mechanistic pathway changes difficult<br />

to interpret. We have described a system using liver-derived<br />

blood flow parameters to restore primary human hepatocyte<br />

biology, allowing for culture at close to physiological insulin/<br />

glucose conditions and eliciting drug responses at clinically-relevant<br />

concentrations. We used this system to assess three DILI<br />

causing drugs, Acetaminophen (APAP), Trovafloxacin (TVX)<br />

and Chlorpromazine (CPZ), to differentiate their potential<br />

for toxicity and transcriptomic signatures, at concentrations<br />

approximating both therapeutic and toxic levels. For instance,<br />

the toxic concentration of APAP used (1323 μM) was based on<br />

serum levels that would initiate treatment for acute toxicity using<br />

the Rumack-Matthew nomogram, and is significantly lower than<br />

concentrations typically reported in conventional in vitro culture<br />

systems. Primary hepatocytes from 5 human donors plated in<br />

the system were exposed to the drugs for 48 hours. Whole<br />

genome transcriptomics was performed using RNAseq, along<br />

with functional assays for CYP activity and function. Distinctive<br />

responses distinguishing the toxicity potential of the drugs at<br />

a signaling and phenotypic level underscored the differences<br />

in underlying mechanisms. Cholestasis-related pathways and<br />

genes were perturbed by CPZ and TVX but not APAP, consistent<br />

with lack of cholestatic phenotype in clinical reports of<br />

APAP toxicity. Fibrotic potential, reflected by increased TGF-β<br />

signaling, and inflammatory changes evidenced by NFκB activation<br />

were only seen with TVX. Coagulation and complement<br />

activation was noted only with APAP while CPZ exhibited activation<br />

of unfolded protein response. Novel findings including<br />

suppression of TGF-β with APAP, are currently being explored.<br />

The correlation of the cellular responses to clinical effects<br />

reported with the drugs, demonstrates the system’s ability to


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1143A<br />

predict toxicity mechanisms at clinically relevant concentrations<br />

and differentiate distinct mechanisms of DILI.<br />

Disclosures:<br />

Robert Figler - Employment: HemoShear Therapeutics, LLC<br />

Brett R. Blackman - Board Membership: HemoShear LLC; Management Position:<br />

HemoShear LLC; Patent Held/Filed: HemoShear LLC; Stock Shareholder:<br />

HemoShear LLC<br />

Svetlana Marukian - Employment: HemoShear LLC<br />

Mark Lawson - Employment: Hemoshear<br />

Aaron J. Mackey - Employment: HemoShear, LLC<br />

David Manka - Employment: HemoShear Therapeutics; Stock Shareholder:<br />

HemoShear Therapeutics<br />

Brian R. Wamhoff - Stock Shareholder: HemoShear, LLC<br />

Ajit Dash - Employment: HemoShear LLC<br />

The following authors have nothing to disclose: Charles Qualls, Sol Collado<br />

1917<br />

Caffeine Confers Hepatoprotection by Preconditioning<br />

Hepatocytes with Superior Ability to Withstand DNA<br />

Damage through Regulation of ATM Signaling Pathways<br />

Priya Gupta 1 , Yogeshwar Sharma 1 , Sanjeev Gupta 2 , Preeti Viswanathan<br />

2 ; 1 Medicine and Pathology, Marion Bessin Liver Centre,<br />

Albert Einstein College of Medicine, Bronx, NY; 2 Division of<br />

Pediatric Gastroenterology, Hepatology and Nutrition, Children’s<br />

Hospital at Montefiore Medical Center, Albert Einstein College of<br />

Medicine, Bronx, NY<br />

Background: Epidemiological <strong>studies</strong> indicate that regular<br />

consumption of caffeine decreases severity of chronic hepatitis,<br />

cirrhosis, etc. However, the molecular basis for this organ<br />

protective effect of caffeine is unknown. To develop needed<br />

paradigms for defining hepatoprotective basis of caffeine we<br />

hypothesized that ATM signaling pathways, which protect cells<br />

from DNA damage, will be involved since caffeine inhibits<br />

ATM kinase activity. Methods: We used HuH-7 human hepatocellular<br />

carcinoma cells for model development, including<br />

IC50 cytotoxicity with acetaminophen (APAP) using MTT cell<br />

viability assays, DNA damage by γH2AX staining and Comet<br />

formation, and a cell subline expressing hATM promoter-driven<br />

tdTomato reporter for evaluating global pathway-specific<br />

changes. Cells were studied with or without caffeine preconditioning<br />

over 5 d. Flow cytometry was performed to analyze cell<br />

cycling with PI staining and Side Population (SP) was analyzed<br />

by Hoechst dyes with or without verapamil pretreatment to<br />

inhibit dye efflux through MDR activity. Results: APAP toxicity<br />

in HuH-7 cells led to dose-dependent increases in γH2AX<br />

expression, Comet formation and greater ATM promoter activity.<br />

When cells were exposed to APAP plus 10 mM caffeine<br />

overnight, cytotoxicity was worsened, along with further significant<br />

increases in γH2AX expression, Comet formation and<br />

ATM promoter activity, along with depletion of cells in S and<br />

G2/M, p


1144A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

for the treatment or prevention of drug-induced fulminant liver<br />

failure.<br />

Disclosures:<br />

Etsuro Hatano - Speaking and Teaching: Bayer<br />

Masa Asagiri - Grant/Research Support: Astellas Pharma Inc.<br />

The following authors have nothing to disclose: Kenji Takemoto, Keiko Iwaisako,<br />

Masatoshi Takeiri, Naruto Noma, Saori Ohmae, Kan Toriguchi, Kazutaka<br />

Tanabe, Keigo Machida, Kojiro Taura, Shinji Uemoto<br />

1919<br />

Statins prevent liver tumor promoting effect of dioxin-like<br />

environmental toxins in different biological models<br />

Alvarez Laura; Facultad de medicina UBA, Argentina, Argentina<br />

Hepatocellular carcinoma (HCC) is the third leading cause of<br />

cancer death worldwide. Hormonal imbalance plays a key role<br />

in the development of neoplasms. There is an open relationship<br />

between HCC and environmental toxins. Hexachlorobenzene<br />

(HCB) is an environmental pollutant, associated with a broad<br />

spectrum of harmful effects on human health as alterations in<br />

thyroid metabolism, neurotoxicity, developmental and carcinogenic<br />

effects in human and experimental animals. Statins<br />

reduce the incidence of various tumors. Their anti-tumor activity<br />

has been related to their pro-apoptotic and anti-angiogenic<br />

effect and the prevention of metastasis. However, the exact<br />

mechanism mediating the in vivo anti-tumor effect of statins has<br />

not been yet fully clarified. The objective of this study was to<br />

determine: key molecules involved in HCB promotion of liver<br />

preneoplastic foci in an initiation-promotion model in rats [diethylnitrosamine<br />

(DEN) (100 mg / kg bw) and HCB (100 mg / kg<br />

bw)]. 1- We evaluated in rat liver: a) Proliferating cell nuclear<br />

antigen levels (PCNA) in focal and non-focal areas, (Western<br />

blot); b) thyroid hormones (TH) concentration, c) deiodinase<br />

types I (DI) and III (DIII) mRNA levels; c) 3-Hydroxy-3-methylglutaryl-coenzyme<br />

A reductase (HMGCoAR) (RIA and RT-PCR),<br />

d) Thioredoxin 1 levels (TRX1) and d) serum cholesterol. 2-In<br />

Hep-G2 cells, we analyzed the ability of atorvastatin (AT) and<br />

simvastatin (SM) to reverse the effect of HCB on key molecules<br />

involved in the proliferative effect of HCB. Hep-G2 cells were<br />

treated with HCB (5 uM), in the presence and absence of AT<br />

(10, 20 and 30 mM) and SM (5, 10, 20 uM). We analyzed<br />

protein levels of: a) PCNA, b) pERK1/2, c) cyclin D1 (CD1); b)<br />

TRX1; d) TGF-β1 and HMGCoAR mRNA expression (RT-PCR).<br />

Results: In vivo: HCB increased (60% p≤0,001) PCNA positive<br />

cells in focal areas (DEN + HCB) vs. DEN. HMGCoAR<br />

increased 31% (p≤0,01); T 4<br />

increased 38% (p≤0.01) and T 3<br />

decreased 37% (p≤0,01). DIII increased 30% (p≤0.01) and<br />

DI declined 41% (p≤0.01). TRX1 increased 29% (p≤0.01) in<br />

(DEN + HCB) (p≤0,01). Cholesterol increased 28% (p ≤ 0.05).<br />

In vitro, the proliferative effect of HCB decreased with AT or<br />

SM. HMGCoAR decreased 29% and 38% with AT (20 and<br />

30 mM), and 20% and 31% with SM (10 to 20 mM) respectively.<br />

TRX 1 and TGF-β1 protein levels decreased with AT (30<br />

mM) and SM (20 mM, 34%), p ≤ 0.05. Conclusion: AT and<br />

SM reduced HCB-induced cell proliferation in Hep-G2 cells.<br />

HMGCoAR, TGF-β1, TRX1 and HT, may be potential target<br />

molecules for statins mechanism of action on HCC caused by<br />

environmental toxins.<br />

Disclosures:<br />

The following authors have nothing to disclose: Alvarez Laura<br />

1920<br />

Drug-Induced Liver Injury Associated With Vemurafenib<br />

Treatment<br />

Marie Lou Gacusan Munson, Liat Schwartz Sagi, Roland Morley,<br />

Wassim Aldairy, Mason Shih, Edwin Tucker; Genentech, South<br />

San Francisco, CA<br />

Background: Vemurafenib (VEM) is the first-in-class selective<br />

oncogenic BRAF V600 kinase inhibitor approved for adults<br />

with unresectable or metastatic melanoma. In preclinical and<br />

clinical development programs, VEM has been identified to<br />

cause liver laboratory abnormalities. The general incidence<br />

of drug-induced liver injury (DILI) is reported to be 1:10,000<br />

to 1:100,000 (all severity grades) and is usually identified in<br />

the postmarketing setting after considerable patient (pt) exposure.<br />

We describe the incidence, characteristics, risk factors,<br />

specific clinical signature, and clinical course of hepatotoxicity<br />

with VEM. Methods: The Genentech company safety database<br />

was searched for medically confirmed, serious cases of hepatic<br />

adverse events. The definition of DILI was based on the clinical<br />

chemistry criteria threshold recommended by the International<br />

DILI Expert Working Group (EWG) (Aithal et al. Clinical Pharmacol<br />

Ther. 2011;89:806-815). The WHO Global Introspection<br />

method was used in the assessment of causality. Clinical<br />

pattern of liver injury was categorized using the R value; severity<br />

grading was based on the EWG DILI severity index. Results:<br />

63 cases of DILI attributed to VEM were identified: 46 from clinical<br />

trials and 17 from spontaneous reports. Median age was<br />

58 years (range, 28-83); 17 pts were elderly (≥65 years old).<br />

27 pts were male and 36 were female. The R value was provided<br />

for 43 pts; the pattern of injury was cholestatic/mixed in<br />

28 pts and hepatocellular in 15. Median latency was 43 days<br />

(range, 1-265) and was similar between cholestatic/mixed<br />

and hepatocellular patterns. The distribution of clinical pattern<br />

of injury was similar for elderly and younger pts. DILI severity<br />

was assessed in 40 pts and was mild, moderate, and severe in<br />

11, 27, and 2 pts, respectively. No cases had fatal outcomes,<br />

and no cases required liver transplantation. VEM treatment<br />

interruption was necessary in 46 of 63 pts; dose modifications<br />

according to guidance in the product label were necessary in<br />

25 of those 46. The event did not recur in 20 of the 25 pts.<br />

In 5 cases, treatment interruption/dose modification required<br />

permanent discontinuation. Conclusion: The crude reporting<br />

rate of DILI for VEM is 5.13 per 1000 pt-years (95% CI, 3.8-<br />

6.4) based on estimated 12,262 pt-years’ exposure. A median<br />

latency of 43 days with a tendency toward a cholestatic/mixed<br />

pattern was observed, and 2 cases were assessed as severe.<br />

Risk factors and populations at risk were not identified. The prescribed<br />

dose modification in the label was helpful in managing<br />

cases. To our knowledge, this is the first publication describing<br />

DILI with a BRAF inhibitor.<br />

Disclosures:<br />

Marie Lou Gacusan Munson - Employment: Genentech<br />

Liat Schwartz Sagi - Employment: Genentech-Roche<br />

Roland Morley - Employment: Roche Genentech; Stock Shareholder: Bristol Myers<br />

Squibb<br />

The following authors have nothing to disclose: Wassim Aldairy, Mason Shih,<br />

Edwin Tucker


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1145A<br />

1921<br />

3D culture strategies for improved MSCs-derived<br />

hepatocyte-like cells: potential toxicological and clinical<br />

applications<br />

Madalena Z. Cipriano 1 , Nora Freyer 2 , Fanny Knoespel 2 , Rita Barcia<br />

3 , Pedro E. Cruz 3 , Helder Cruz 3 , Nuno G. Oliveira 1 , Jorge M.<br />

Santos 3 , Katrin Zeilinger 2 , Joana P. Miranda 1 ; 1 CBT, Research<br />

Institute for Medicines (iMed.ULisboa), Faculty of Pharmacy, Universidade<br />

de Lisboa. Av. Prof. Gama Pinto, 1649-003 Lisbon,<br />

Portugal, Lisboa, Portugal; 2 Bioreactor Group, Berlin Brandenburg<br />

Center for Regenerative Therapies (BCRT), Charité Universitätsmedizin<br />

Berlin, 13353 Berlin, Germany, Berlin, Germany; 3 ECBio<br />

S.A., Rua Henrique Paiva Couceiro, N 27, 2700-451-Amadora,<br />

Portugal, Amadora, Portugal<br />

Background: Three-dimensional (3D) cultures have emerged<br />

as promising alternative models for maintaining hepatocyte<br />

phenotype in short and long-term in vitro cultures by better<br />

resembling the in vivo environment of the liver. However,<br />

human primary hepatocytes have limited availability and<br />

high inter-individual differences. As such, the differentiation<br />

of human stem cells into hepatocyte-like cells (HLCs) has been<br />

suggested. Human neonatal mesenchymal stem cells (hnMSCs),<br />

derived from the umbilical cord tissue, are particularly interesting<br />

given their proliferation potential, low immunogenic<br />

and immuno-modulatory properties along with a demonstrated<br />

HLCs differentiation in monolayer cultures. However a mature<br />

hepatic phenotype was not yet achieved. In this study, 3D cell<br />

cultures, namely a spheroid (SC) and a 3D multi-compartment<br />

membrane bioreactor (BR) system, were tested as an approach<br />

for improving HLCs maturation in vitro. Methods: A 27-day<br />

differentiation protocol consisting of the sequential addition<br />

of cytokines/growth factors was optimized. Firstly, hnMSCs<br />

(UCX ® ) were differentiated into hepatoblast-like cells in 2D for<br />

17 days and afterwards cultured under 3D conditions (SC and<br />

BR) or maintained in 2D as control for maturation over 10<br />

days. Results: All systems enabled hnMSC differentiation into<br />

HLCs as shown by positive immunostaining of hepatic markers,<br />

such as cytoskeleton proteins (CK-18), the transcription<br />

factor HNF-4α, albumin, the hepatic transporters OATP-C and<br />

MRP-2 and CYP1A2 and CYP3A4. Additionally, the three culture<br />

models displayed relevant glucose metabolism, tested by<br />

the capacity to store glycogen and to release glucose into the<br />

medium, and the ability to convert ammonia into urea, to produce<br />

albumin and to perform phase I metabolism. However,<br />

urea production (5.8 and 8.3 fold induction for the BR and the<br />

SC, respectively) and EROD activity (4.8 and 3.8 fold induction<br />

in the BR and the SC, respectively) were both increased in<br />

3D relative to 2D. Albumin production was also induced 2.9<br />

fold in the BR when compared to 2D. Finally, cells differentiated<br />

within BR presented relevant CYP1A2 and 3A4 activities<br />

when compared with 2D, whereas HLCs from spheroids and<br />

2D presented similar CYP3A4, 1A1 and 2C9 activity on day<br />

27. Conclusions: Overall both the SC and BR models improved<br />

HLCs maturation suggesting its potential for in vitro and clinical<br />

applications, namely drug toxicity prediction, regenerative<br />

medicine or external liver support. Additionally, differential<br />

HLCs phenotypes were observed for each of the 3D models,<br />

further supporting that the optimal culture system should be<br />

selected depending on the scientific purpose.<br />

Disclosures:<br />

Rita Barcia - Management Position: ECBio<br />

Pedro E. Cruz - Board Membership: ECBio, S.A.<br />

Helder Cruz - Board Membership: ECBio<br />

Jorge M. Santos - Employment: ECBio<br />

The following authors have nothing to disclose: Madalena Z. Cipriano, Nora<br />

Freyer, Fanny Knoespel, Nuno G. Oliveira, Katrin Zeilinger, Joana P. Miranda<br />

1922<br />

Death and Liver Transplantations Occurring Within Two<br />

Years of Drug-Induced Liver Injury<br />

Paul H. Hayashi 1 , Don C. Rockey 2 , Robert J. Fontana 3 , Hans L.<br />

Tillmann 4 , Neil Kaplowitz 5 , Huiman X. Barnhart 6 , Jiezhun Gu 6 ,<br />

Averell H. Sherker 7 , Naga P. Chalasani 9 , Victor J. Navarro 8 ,<br />

Jawad Ahmad 10 , Jay H. Hoofnagle 7 ; 1 Division of Gastroenterology<br />

& Hepatology, University of North Carolina, Chapel Hill, NC;<br />

2 Medical University of South Carolina, Charleston, SC; 3 University<br />

of Michigan, Ann Arbor, MI; 4 East Carolina University, Greenville,<br />

NC; 5 University of Southern California, Los Angeles, CA; 6 Duke<br />

University, Durham, NC; 7 National Institutes of Health, Bethesda,<br />

MD; 8 Einstein Medical Center, Philadelphia, PA; 9 Indiana University,<br />

Indianapolis, IN; 10 Mount Sinai Medical Center, New York,<br />

NY<br />

Background: While idiosyncratic drug-induced liver injury (DILI)<br />

is usually reversible, it can lead to liver transplantation (LT)<br />

and/or death. But the overall mortality rate and risk factors for<br />

death or LT are not well defined. Aim: To characterize the role<br />

and clinical features of DILI in deaths and LT cases. Methods:<br />

All patients in the Drug Induced Liver Injury Network (DILIN)<br />

who died or underwent LT within 2 years of follow-up were<br />

identified. Each case was reviewed independently by 2 hepatologists<br />

who judged whether DILI had a primary, contributory<br />

or no role in the death or LT. Cases of DILI having a primary<br />

role were categorized as acute liver failure (ALF: 6 months) or other.<br />

Cases of DILI having a contributory role were assigned another<br />

cause of death (e.g. cancer, sepsis). Discrepancies between<br />

reviewers were resolved by conference call. R-values (ALT/ULN<br />

÷ ALP/ULN) at DILI onset were used to classify injury patterns<br />

as hepatocellular (R>5), mixed (R 2-5) or cholestatic (R


1146A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Paul H. Hayashi, Neil Kaplowitz,<br />

Huiman X. Barnhart, Jiezhun Gu, Averell H. Sherker, Victor J. Navarro, Jawad<br />

Ahmad, Jay H. Hoofnagle<br />

1923<br />

Severe and prolonged jaundice associated with body<br />

building supplements (BBS): review of 44 cases from the<br />

Drug-Induced Liver Injury Network (DILIN)<br />

Andrew Stolz 1 , Victor J. Navarro 2 , Paul H. Hayashi 3 , Naga P.<br />

Chalasani 4 , Robert J. Fontana 5 , Herbert L. Bonkovsky 6 , Sunil<br />

Hwang 6 , Jay H. Hoofnagle 7 , David E. Kleiner 8 , Huiman X. Barnhart<br />

9 ; 1 Medicine, University of Southern California, Los Angeles,<br />

CA; 2 Medicine, Einstein Medical Center, Philadelphia, PA;<br />

3 Medicine, University of North Carolina, Chapel HIll, NC; 4 Medicine,<br />

Indiana University School of Medicine, Indianapolis, IN;<br />

5 Medicine, University of Michigan, Ann Arbor, MI; 6 Medicine<br />

and Research, Carolinas Medical Center, Charlotte, NC; 7 Liver<br />

Disease Research Branch, Division of Digestive Diseases and<br />

Nutrition, National Institute of Diabetes and Digestive and Kidney<br />

Diseases, Bethesda, MD; 8 Laboratory of Pathology, National Cancer<br />

Institute, Bethesda, MD; 9 Duke Clinical Research Institute, Duke<br />

University, Durham, NC<br />

Background & Aim: The use of body building supplements<br />

(BBS) containing anabolic steroids (AS) continues to rise<br />

despite efforts to curtail their availability. We sought to characterize<br />

the clinical features and outcomes of presumed AS-induced<br />

jaundice in patients taking BBS prospectively enrolled<br />

in the DILIN. Methods: 44 (5%) out of 847 cases of liver injury<br />

attributed to drug or herbal dietary supplements enrolled<br />

between Sept 2004 and March 2013 were attributed solely to<br />

BBS. The presence of various natural and synthetic androgens<br />

in consumed products was quantified by LC/MS compared<br />

with 18 known standards with an average assay sensitivity<br />

[limit of quantitation] ~ 38 fmoles/mg of product. Liver biopsies<br />

were interpreted by the study hepatopathologist without<br />

clinical information. Results: All 44 were men, ages 21 to 59<br />

(mean = 32) years with 81% self-identified as Caucasian and<br />

the remainder African-American. 2 subjects had chronic HCV<br />

infection. The BBS products were used for body building and<br />

purported to contain various AS with different chemical structures<br />

and commercial names. All were were taken orally in otherwise<br />

healthy males. Median latency to onset was 1.8 (range<br />

0.4 to 14) months, but the exact dates of starting and stopping<br />

were not always available. All 44 cases were jaundiced<br />

and symptomatic at presentation and 43 patients had pruritus<br />

that was frequently severe, prolonged and debilitating. Initial<br />

median bilirubin was 9.8 mg/dL, ALT 168 U/L, Alk P 111<br />

U/L and INR 1. Subsequently, median ALT fell steadily, while<br />

Alk P and bilirubin levels increased 2-3 fold before falling.<br />

Peak median bilirubin in first 6 months was 25.8 (range 7.3-<br />

63). 64% were hospitalized and 18% had evidence of hepatic<br />

dysfunction (e.g. elevated INR). Liver biopsy in 22 patients<br />

reviewed revealed canalicular cholestasis often with only<br />

mild portal and lobular inflammation, in contrast to the classic<br />

“bland cholestasis” often described in association with AS.<br />

Symptomatic hepatic failure was not observed and no patient<br />

died or required liver transplantation. Seven of the 44 patients<br />

(16%) had creatinine elevation > 1.5 mg/dL during their episode<br />

which trended down as the bilirubin improved. Of 14<br />

BBS products available for analysis, 71% had identifiable AS<br />

along with other non-defined steroid containing compounds.<br />

Conclusion: Jaundice associated with BBS containing AS is a<br />

distinct clinical syndrome marked by severe and prolonged<br />

cholestasis in previously healthy men but which did not lead<br />

to hepatic failure or chronic liver injury in this study. AS were<br />

often identified in BBS despite efforts to limit their availability.<br />

Disclosures:<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

Robert J. Fontana - Consulting: GlaxoSmithKline, CLDF; Grant/Research Support:<br />

Gilead, vertex, BMS, Jansen, Gilead<br />

Herbert L. Bonkovsky - Advisory Committees or Review Panels: Recordati Rare<br />

Chemicals, Clinuvel, Inc.; Consulting: Alnylam, Inc, Clinuvel, Inc., Clinuvel, Inc.;<br />

Grant/Research Support: Gilead Sciences<br />

The following authors have nothing to disclose: Andrew Stolz, Victor J. Navarro,<br />

Paul H. Hayashi, Sunil Hwang, Jay H. Hoofnagle, David E. Kleiner, Huiman X.<br />

Barnhart<br />

1924<br />

Combined Activities of JNK1 and JNK2 in Hepatocytes<br />

Protect<br />

Against Toxin-induced Liver Injury<br />

Francisco Javier Cubero 1 , Wei Hu 1 , Gang Zhao 1 , Miguel Eugenio<br />

Zoubek 1 , Jin Peng 1 , Yulia A. Nevzorova 1 , Malika Al Masaoudi 1 ,<br />

Lars Bechmann 4 , Mark V. Boekschoten 3 , Michael Muller 3 , Christian<br />

Preisinger 2 , Nikolaus Gassler 5 , Ali Canbay 4 , Tom Luedde 1 ,<br />

Roger J. Davis 6 , Christian Liedtke 1 , Christian Trautwein 1 ; 1 Department<br />

of Medicine III, University Hospital Aachen, RWTH Aachen,<br />

Aachen, Germany; 2 Proteomics Facility, University Hospital,<br />

RWTH Aachen, Germany, Aachen, Germany; 3 Nutrition, Metabolism<br />

& Genomics group Nutrition, Wageningen, The Netherlands,<br />

Wageningen, Netherlands; 4 Department of Gastroenterology and<br />

Hepatology, University Hospital Duisburg-Essen, Essen, Germany;<br />

5 Institute of Pathology, Aachen, Germany; 6 Howard Hughes Medical<br />

Institute and University of Massachusetts Medical School,<br />

Worcester, Germany<br />

Background& Aims: c-Jun N-terminal kinase (JNK)1 and JNK2<br />

are expressed in hepatocytes and have overlapping and<br />

distinct functions. JNK proteins are activated, via phosphorylation,<br />

in response to acetaminophen- or CCl 4<br />

-induced liver<br />

damage - the level of activation correlates with the degree<br />

of injury. SP600125, a JNK inhibitor, has been reported to<br />

block acetaminophen-induced liver injury. We investigated<br />

the role of JNK in drug-induced liver injury (DILI) in liver tissues<br />

from patients and in mice with genetic deletion of JNK in<br />

hepatocytes. Methods: We studied liver sections from patients<br />

with DILI (due to phenprocoumon, non-steroidal anti-inflammatory<br />

drugs, or acetaminophen) or autoimmune hepatitis, or<br />

patients without acute liver failure (controls), collected from a<br />

DILI Biobank in Germany. Levels of total and activated (phosphorylated)<br />

JNK were measured by immunohistochemistry and<br />

immunoblot assays. Mice with hepatocyte-specific deletion of<br />

Jnk1 (Jnk1 Δhepa ) or combination of Jnk1 and Jnk2 (Jnk Δhepa ),<br />

as well as Jnk1-floxed C57BL/6 (control) mice, were given<br />

repetitive CCl 4<br />

injections to induce fibrosis or acetaminophen<br />

to trigger toxic liver injury. We performed gene expression<br />

microarray, and phosphoproteomic analyses to determine<br />

mechanisms of JNK activity in hepatocytes. Results: Liver samples<br />

from patients with DILI contained more activated JNK, predominantly<br />

in hepatocytes, than healthy tissue. Administration<br />

of acetaminophen to Jnk Δhepa mice produced a greater level<br />

of liver injury than that observed in Jnk1 Δhepa or control mice,<br />

based on levels of serum markers and microscopic and histologic<br />

analysis of liver tissues. Administration of CCl 4<br />

induced<br />

significantly stronger hepatic injury in Jnk Δhep mice, based on<br />

increased inflammation, cell proliferation, and fibrosis progression,<br />

compared to Jnk1 Δhepa or control mice. Hepatocytes from<br />

Jnk Δhepa mice given acetaminophen had an increased oxidative<br />

stress response, leading to decreased activation of AMPK and<br />

JunD and subsequent necrosis. Administration of SP600125<br />

before or with acetaminophen protected Jnk Δhepa and control<br />

mice from liver injury. Conclusions: In hepatocytes, JNK1 and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1147A<br />

JNK2 have combined effects in protecting mice from CCl 4<br />

- and<br />

acetaminophen-induced liver injury. It is important to study the<br />

functions of both proteins, rather than just JNK1, to define their<br />

role during toxic liver injury. JNK inhibition with SP600125<br />

has off-target effects.<br />

Disclosures:<br />

Christian Trautwein - Grant/Research Support: BMS, Novartis, BMS, Novartis;<br />

Speaking and Teaching: Roche, BMS, Roche, BMS<br />

The following authors have nothing to disclose: Francisco Javier Cubero, Wei<br />

Hu, Gang Zhao, Miguel Eugenio Zoubek, Jin Peng, Yulia A. Nevzorova, Malika<br />

Al Masaoudi, Lars Bechmann, Mark V. Boekschoten, Michael Muller, Christian<br />

Preisinger, Nikolaus Gassler, Ali Canbay, Tom Luedde, Roger J. Davis, Christian<br />

Liedtke<br />

1925<br />

Translocation of Iron from Lysosomes to Mitochondria<br />

during Acetaminophen-Induced Toxicity in Mouse<br />

Hepatocytes<br />

Jiangting Hu 1 , Andaleb Kholmukhamedov 1 , Hartmut Jaeschke 2 ,<br />

John J. Lemasters 1 ; 1 Medical University of South Carolina, Charleston,<br />

SC; 2 University of Kansas Medical Center, Kansas city, KS<br />

Background: Acetaminophen (APAP) overdose causes hepatotoxicity<br />

involving mitochondrial dysfunction and the mitochondrial<br />

permeability transition (MPT). Reactive oxygen species<br />

(ROS) play an important role in APAP-induced hepatotoxicity.<br />

Iron is a critical catalyst for ROS formation. Previous <strong>studies</strong><br />

show that disrupted lysosomes release ferrous iron (Fe 2+ )<br />

into the cytosol during APAP hepatotoxicity, which triggers<br />

the MPT and cell killing. Here, our Aim was to investigate<br />

the role of lysosomal iron mobilization into mitochondria<br />

during APAP-induced toxicity to mouse hepatocytes. Methods:<br />

Hepatocytes were isolated from fasted male C57BL/6 mice.<br />

Necrotic cell killing was assessed by propidium iodide fluorimetry.<br />

Mitochondrial membrane potential was visualized by<br />

confocal microscopy of rhodamine 123 (Rh123) or tetramethylrhodamine<br />

methylester (TMRM). Chelatable Fe 2+ was monitored<br />

by quenching of calcein (cytosol) and mitoferrofluor (MFF,<br />

mitochondria). Results: Starch-desferal (1 mM, a lysosomally<br />

targeted iron chelator) and Ru360 (100 nM, an inhibitor of the<br />

mitochondrial calcium uniporter [MCU]) decreased cell killing<br />

after APAP (10 mM) by 50% and 25% at 10 h, respectively.<br />

Progressive quenching of calcein and MFF began after ~4 h,<br />

signifying increased cytosolic and mitochondrial chelatable<br />

Fe 2+ . Mitochondria then depolarized after ~10 h. Dipyridyl, a<br />

membrane-permeable iron chelator, dequenched calcein and<br />

MFF fluorescence. Starch-desferal, but not Ru360, suppressed<br />

cytosolic calcein quenching, whereas both starch-desferal and<br />

Ru360 suppressed mitochondrial MFF quenching and mitochondrial<br />

depolarization. Conclusion: Release of chelatable<br />

Fe 2+ from lysosomes followed by uptake into mitochondria via<br />

MCU occurs during APAP hepatotoxicity, which triggers the<br />

MPT and cell killing. Ru360 prevents the increase of mitochondrial<br />

but not cytosolic Fe 2+ after APAP but is not as protective as<br />

starch-desferal, which prevents Fe 2+ increases in both compartments.<br />

Thus, increased Fe 2+ in the cytosol may also contribute<br />

to APAP hepatotoxicity.<br />

Disclosures:<br />

John J. Lemasters - Consulting: Novo Nordisk<br />

The following authors have nothing to disclose: Jiangting Hu, Andaleb<br />

Kholmukhamedov, Hartmut Jaeschke<br />

1926<br />

Azathioprine And 6-Mercaptopurine Induced Liver<br />

Injury<br />

Bjornsson S. Einar 1,2 , Jiezhun Gu 3 , David E. Kleiner 2 , Naga P.<br />

Chalasani 4 , Paul H. Hayashi 5 , Jay H. Hoofnagle 6 ; 1 The NAtional<br />

University Hospital of Iceland, Reykjavik, Iceland; 2 NIH, Bethesda,<br />

MD; 3 3Duke Clinical Research Institute, Durham, NC; 4 Indiana<br />

University School of Medicine, Indinapolis, IN; 5 University of<br />

North Carolina, Chapel Hill, NC; 6 NIDDK, NIH, Bethesda, MD<br />

Background: Most information of regarding the drug induced<br />

liver injury (DILI) from azathioprine (Aza) and 6-mercaptopurine<br />

(6-MP) comes from isolated case reports and the clinical<br />

features and frequency of outcomes is not well defined. The<br />

aims of this study were to better define the clinical, biochemical<br />

and histologic features of Aza and 6-MP induced liver<br />

injury. Methods: Patients with thiopurine hepatotoxicity from<br />

the DILIN Prospective Study were identified. Patients enrolled<br />

between 2004 and 2014 underwent structured assessment of<br />

potential competing etiologies and analysis of clinical features.<br />

Results: 22 patients (12 attributed to Aza, 10 to 6-MP, none<br />

to thioguanine) were identified who had probable (n=8), very<br />

likely (n=12) or definite (n=2) causality scores. The median<br />

age was 55 years, and 68% were female. Inflammatory bowel<br />

disease was the indication in 55%, and median dose was 150<br />

mg daily (range 25-300 mg). The duration of therapy before<br />

onset varied widely (median 75, range 3 to 2584 days). However,<br />

injury often arose after a dose escalation (59%) and the<br />

median time of onset after a last dose increase was 44 days<br />

(range 3 to 254 days). Overall 73% of cases were icteric. The<br />

median initial ALT was 210 U/L, Alk P 151 U/L and bilirubin<br />

7.4 mg/dL. Median peak bilirubin levels were 13.4 mg/<br />

dL and INR 1.3. Icteric and anicteric cases differed in their<br />

typical clinical features. Cases without jaundiced had minimal<br />

symptoms and a hepatocellular pattern of enzyme elevations,<br />

whereas jaundiced cases had a self-limited cholestatic<br />

hepatitis with minimal or modest elevations in enzyme levels.<br />

Available biopsies (7 cases) showed cholestatic hepatitis; 3<br />

also showed evidence of nodular regeneration. There was little<br />

fibrosis except in the patient with pre-existing cirrhosis. There<br />

were no major differences between Aza and 6-MP cases. One<br />

The patient with pre-existing cirrhosis underwent emergency<br />

liver transplantation, all others resolved the injury clinically,<br />

although one patient had moderate Alk P elevations 2 years<br />

after onset. Conclusions: Nearly three-quarters of thiopurine-induced<br />

liver injury presents with self-limited cholestatic hepatitis.<br />

Injury often arises within a few months of dose increase,<br />

highlighting the importance of monitoring liver tests after dose<br />

escalaction. The prognosis of Aza and 6-MP related liver injury<br />

is favorable except in patients with pre-existing cirrhosis.<br />

Disclosures:<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

The following authors have nothing to disclose: Bjornsson S. Einar, Jiezhun Gu,<br />

David E. Kleiner, Paul H. Hayashi, Jay H. Hoofnagle<br />

1927<br />

Gender-specific glucuronidation of acetaminophen in<br />

htgUGT1A WT mice<br />

Anja Winkler, Sandra Kalthoff, Christian P. Strassburg; Clinics and<br />

Policlinic I, University Hospital Bonn, Bonn, Germany<br />

Aims: Acetaminophen (APAP) is a commonly used over-thecounter<br />

drug with analgesic as well as antipyretic properties<br />

and at low doses is primarily metabolized through sulfation<br />

and glucuronidation, while only a small fraction is oxidized to<br />

the reactive metabolite N-acetyl-p-benzoquinone (NAPQI) by


1148A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

cytochrome P450 enzymes. At supratherapeutic doses saturation<br />

of the sulfation and glucuronidation pathways lead to an<br />

increase of NAPQI and therefore to APAP-mediated hepatotoxicity<br />

potentially causing acute liver failure (ALF). Previous <strong>studies</strong><br />

have already reported gender-related differences in APAP<br />

metabolism with females displaying relative resistance towards<br />

APAP-induced hepatotoxicity. However, the exact molecular<br />

mechanisms involved are not fully understood. Aim of this study<br />

was therefore to evaluate possible gender-related differences<br />

in glucuronidation of APAP in a humanized tgUGT1A mouse<br />

model. Methods: Male and female htgUGT1A WT mice were<br />

intraperitoneally (i.p.) treated with acetaminophen (APAP, 5<br />

mg/kg) or vehicle for three consecutive days. Expression of<br />

UGT1A genes and transcription factors (TF) HNF-4α, PXR and<br />

ESR were analyzed by Taqman-PCR. Results: Assessment of<br />

hepatic UGT1A mRNA expression in APAP-treated animals<br />

revealed a 10-, 36- and 11-fold increase of UGT1A1, UGT1A6<br />

and UGT1A9 in female mice contrasting an only 7-, 9- and<br />

10-fold induction of the aforementioned isoforms in males.<br />

Additionally, female mice exhibited significant increases of<br />

UGT1A1 (3.2-, 7.4-fold) and UGT1A6 (5.6-, 4.8-fold) in both<br />

jejunum and colon, whereas in males upregulation of intestinal<br />

UGT1A genes was only observed in colon (UGT1A1: 6.3-fold;<br />

UGT1A6: 7.2-fold). Interestingly, investigation of TFs in APAPtreated<br />

htgUGT1A WT mice revealed upregulation of ESR and<br />

PXR mRNA in livers of females, which was reduced in male<br />

mice. Moreover, HNF4-a expression was downregulated in<br />

males. Conclusions: Analyses of hepatic as well as intestinal<br />

UGT1A genes relevant for APAP glucuronidation uncovers significant<br />

differences in UGT1A mRNA expression between male<br />

and female mice following treatment with acetaminophen. Particularly<br />

APAP-mediated upregulation of hepatic UGT1A6 was<br />

profoundly reduced in male htgUGT1A WT mice in comparison<br />

to their female counterparts. These data suggest a possible<br />

role of glucuronidation for gender-related distinctions in<br />

susceptibility towards APAP-mediated hepatotoxicity and clinically<br />

observed sex-related differences in acetaminophen pharmacokinetics<br />

and –dynamics. Furthermore, reduced hepatic<br />

expression of the TFs ESR, PXR and HNF4-a represents a possible<br />

molecular mechanism for differential UGT1A regulation<br />

in males<br />

Disclosures:<br />

Christian P. Strassburg - Advisory Committees or Review Panels: Novartis, Roche;<br />

Speaking and Teaching: Novartis, Merz, MSD, Falk Pharma, BMS, Abbvie<br />

The following authors have nothing to disclose: Anja Winkler, Sandra Kalthoff<br />

1928<br />

Drug Induced Liver Injury (DILI) in the East<br />

Sarah E. Low 1 , Kieron B. Lim 1 , Seng Gee Lim 1,2 ; 1 Gastroenterology<br />

& Hepatology, National University Health System, Singapore,<br />

Singapore; 2 Yong Loo Lin School of Medicine, National University<br />

of Singapore, Singapore, Singapore<br />

Background Is Drug Induced Liver Injury (DILI) different in the<br />

East compared with the West? There are only a handful of<br />

population <strong>studies</strong> worldwide estimating crude incidence rates<br />

of 3-19 of 100,000 inhabitants. Asians have different pharmacogenetic<br />

profiles with different disease patterns, susceptibility<br />

and hence handle drugs differently. There is widespread use<br />

of herbal and alternative medicines (HM), the safety of which<br />

is poorly defined. Aims To review existing literature on prevalence<br />

of DILI in the East compared to the West. We also<br />

seek to examine differences in frequency of causative agents<br />

of DILI in the East and postulate why. Methods A Medline<br />

search with “Drug-Induced Liver Injury[Mesh] OR Hepatotoxicity<br />

AND” was undertaken. Countries included:Singapore,<br />

Malaysia, Indonesia, Thailand, India, Sri Lanka,<br />

Vietnam, China, Korea, Taiwan, Pakistan, Cambodia, Japan,<br />

Hong Kong, Nepal, Mongolia, Laos, Philippines. Results Anti<br />

tuberculosis (TB) drugs (57% of DILI in India) and HM were most<br />

commonly implicated. Of 290 articles, 43.5% were on TB DILI.<br />

Although TB is more prevalent in Asia, Asians appear more<br />

susceptibile to TB DILI, with twice as much prevalence, 6.1%<br />

versus 3.1%, likely due to polymorphisms of drug metabolizing<br />

genes (NAT2, CYP2D1, GSTM1, GSTT1). Despite under<br />

reporting, proportion of DILI from HM ranged from 18%-40%<br />

with incidence of DILI in HM consumers to be 0.7%-1.5%. Hepatotoxicity<br />

from Amoxicillin Clavulanate was rarely reported in<br />

the East. Severity of intrinsic hepatotoxicity of Acetaminophen<br />

is attenuated in the East, with virtually no cases of liver failure<br />

or death reported. Conclusion The profile of DILI appears to<br />

be different in the Asia compared to the West, with primarily<br />

TB DILI and Herbal DILI, and rarely Amoxicillin Clavulanate or<br />

Acetaminophen in contrast to the West<br />

Disclosures:<br />

Kieron B. Lim - Advisory Committees or Review Panels: Gilead, Abbvie, Sirtex;<br />

Consulting: Novartis; Grant/Research Support: Astellas, Bayer<br />

Seng Gee Lim - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals, Gilead<br />

Pharmaceuticals, Roche Pharmaceuticals; Speaking and Teaching: Bristol-Myers<br />

Squibb, Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals,<br />

Gilead Pharmaceuticals<br />

The following authors have nothing to disclose: Sarah E. Low<br />

1929<br />

Application of the Drug-Induced Liver Injury Expert<br />

Working Group Criteria in the Assessment of Hepatotoxicity<br />

With the BRAF Inhibitor Vemurafenib<br />

Marie Lou Gacusan Munson, Liat Schwartz Sagi, Mason Shih,<br />

Edwin Tucker; Genentech, South San Francisco, CA<br />

Background: Vemurafenib (VEM) is the first BRAF inhibitor<br />

approved by the US Food and Drug Administration (FDA) and<br />

the European Medicines Agency (EMA) for patients with metastatic<br />

melanoma. At the time of its approval, based on a safety<br />

database of ~500 patients (EMA. Public Assessment Report<br />

on Vemurafenib. 2011), VEM was known to cause liver laboratory<br />

abnormalities. After 2 y of commercial use, Hy’s law<br />

cases were identified, prompting a comprehensive hepatotoxicity<br />

risk review of VEM. The international Drug-Induced Liver<br />

Injury (DILI) Expert Working Group (EWG) DILI criteria (Aithal<br />

et al. Clinical Pharmacol Ther. 2011;89:806-815), with an<br />

EBM 2B level of evidence, were used to evaluate VEM. The


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1149A<br />

aims were to (1) assess the diagnostic yield of the EWG clinical<br />

chemistry criteria for DILI when applied to individual case<br />

safety reports of VEM, (2) describe the limitations of the EWG<br />

criteria in real-world application and (3) present the modifications<br />

implemented. Methods: We conducted a retrospective<br />

review and analysis of the EWG algorithm and diagnostic criteria<br />

for identifying and evaluating DILI cases for VEM. Results:<br />

228 serious, medically confirmed cases with hepatic adverse<br />

events were included; 81 met the DILI EWG criteria threshold:<br />

a diagnostic yield of 35.5%. The EWG algorithm was modified<br />

at 3 levels: (1) The CTCAE severity grading was substituted<br />

for actual laboratory values. (2) Cases that did not meet the<br />

DILI criteria but were reported as hepatic failure or had fatal<br />

outcomes were included during causality assessment so clinically<br />

significant cases were not missed. (3) To allow flexibility<br />

with missing data, the WHO global introspection method for<br />

causality assessment was used instead of the CIOMS Roussel<br />

Uclaf Causality Assessment Method (RUCAM). Conclusions:<br />

The DILI EWG clinical chemistry criteria and algorithm provide<br />

an objective and systematic approach to the assessment of DILI<br />

within the postmarketing setting. The modifications proposed<br />

herein address the inadequate case information from spontaneous<br />

sources.<br />

Disclosures:<br />

Marie Lou Gacusan Munson - Employment: Genentech<br />

Liat Schwartz Sagi - Employment: Genentech-Roche<br />

The following authors have nothing to disclose: Mason Shih, Edwin Tucker<br />

1930<br />

Minocycline hepatotoxicity: Clinical characterization and<br />

identification of HLA- B*35:02 as a risk factor<br />

Thomas J. Urban 1,2 , Paola Nicoletti 3 , Naga P. Chalasani 4 , Jose<br />

Serrano 5 , Andrew Stolz 6 , Ann K. Daly 7 , Guruprasad P. Aithal 8 ,<br />

John F. Dillon 9 , Huiman X. Barnhart 10 , Paul B. Watkins 2 , Robert<br />

J. Fontana 11 ; 1 Eshelman School of Pharmacy, University of North<br />

Carolina at Chapel Hill, Chapel Hill, NC; 2 Hamner Institute for<br />

Drug Safety Sciences, Durham, NC; 3 Center for Computational<br />

Biology and Bioinformatics, Columbia University, New York,<br />

NY; 4 Indiana University School of Medicine, Indianapolis, IN;<br />

5 National Institute of Diabetes Digestive and Kidney Diseases,<br />

Bethesda, MD; 6 University of Southern California, Los Angeles,<br />

CA; 7 University of Newcastle, Newcastle, United Kingdom; 8 Nottingham<br />

University, Nottingham, United Kingdom; 9 University of<br />

Dundee, Dundee, United Kingdom; 10 Duke University, Durham,<br />

NC; 11 University of Michigan, Ann Arbor, MI<br />

Prolonged minocycline use has been associated with rare<br />

instances of hepatotoxicity that can present with prominent<br />

autoimmune features in previously healthy individuals. Prior<br />

genome-wide association <strong>studies</strong> (GWAS) have demonstrated<br />

human leukocyte antigen (HLA) polymorphisms that are associated<br />

with DILI susceptibility to various agents. The aim of this<br />

study was to identify potential genetic determinants of minocycline<br />

DILI in a well-phenotyped cohort of patients. METHODS: A<br />

total of 25 Caucasian patients (22 from DILIN, 3 from iDILIC)<br />

with DILI due to minocycline underwent genome-wide genotyping<br />

and were compared to population controls. Coding<br />

sequences of HLA alleles were imputed based on SNP genotypes<br />

in the HLA region. Verification of HLA carrier status was<br />

undertaken using sequence-based HLA typing and included an<br />

additional 2 DILIN cases that were not included in the GWAS.<br />

RESULTS: Amongst the 26 cases from DILIN with complete clinical<br />

information, 73.1% were female with a median age of<br />

19.5 years (range: 15 to 61). The median latency from drug<br />

start to DILI onset was 330 days (range: 15 to 1350) with 70%<br />

of cases presenting after 6 months of minocycline exposure. At<br />

DILI onset, 50% were jaundiced, 19% reported rash, and 31%<br />

had fever but only 4% had eosinophilia. At presentation, 82%<br />

had acute hepatocellular liver injury, median ALT of 1212 IU/L<br />

(range: 70 to 2500), median bilirubin of 4.5 mg/dl (range:<br />

0.2 to 16.7), 92% had a + ANA and 29% had a + SmAb.<br />

During follow-up, 46% were treated with corticosteroids and<br />

the median time to ALT normalization was 93 days. None of<br />

the subjects died or required a liver transplant but 25% had evidence<br />

of chronic liver injury at 6 months of follow-up. GWAS<br />

revealed a strong association between HLA-B*35:02 and risk<br />

for minocycline-DILI, with a carrier frequency of 16% in DILI<br />

cases compared to 0.6% in population controls (Odds Ratio:<br />

29.6, 95% CI: 9.0-97.5, p=2.5 x 10 -8 ). Verification of HLA-<br />

B*35:02 imputation was confirmed by sequence-based HLA<br />

typing. HLA-B*35:02 carriers had a lower R-value at presentation<br />

(9 vs. 26, p=0.018) but otherwise had similar presenting<br />

features and outcomes compared to non-carriers. HLA-B*35:02<br />

has been shown to hasten progression to AIDS among HIV<br />

patients, but has not previously been associated with any drugor<br />

liver-related phenotypes. CONCLUSIONS: Long-term treatment<br />

with minocycline can result in infrequent severe acute<br />

hepatocellular liver injury with prominent autoimmune features<br />

in young female patients. HLA-B*35:02 is a rare HLA allele<br />

that is strongly associated with minocycline-DILI, and may be a<br />

useful diagnostic marker of this serious adverse event.<br />

Disclosures:<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

Paul B. Watkins - Consulting: Abbott, Actelion, Boerringer-Ingelheim, Cempra,<br />

Alecra, Roche, Merck, Reservlogix, Intercept, Janssen, Novartis, Otsuka, Pfizer,<br />

Sanolfi, Takeda, UCB, Bristol-Myers Squibb, GSK<br />

Robert J. Fontana - Consulting: GlaxoSmithKline, CLDF; Grant/Research Support:<br />

Gilead, vertex, BMS, Jansen, Gilead<br />

The following authors have nothing to disclose: Thomas J. Urban, Paola Nicoletti,<br />

Jose Serrano, Andrew Stolz, Ann K. Daly, Guruprasad P. Aithal, John F. Dillon,<br />

Huiman X. Barnhart<br />

1931<br />

Liver Disorders that can Masquerade as Drug-Induced<br />

Liver Injury<br />

Don C. Rockey 1 , Paul H. Hayashi 2 , Hans L. Tillmann 3 , Jiezhun<br />

Gu 3 , Marwan Ghabril 4 , Jawad Ahmad 5 , Andrew Stolz 7 , Robert J.<br />

Fontana 8 , Jay H. Hoofnagle 6 ; 1 Medical University of South Carolina,<br />

Charleston, SC; 2 University of North Carolina, Chapel Hill,<br />

NC; 3 Duke University, Durham, NC; 4 Indiana University, Indianapolis,<br />

IN; 5 Mount Sinai, New York, NY; 6 NIH, Bethesda, MD;<br />

7 University of Southern California, Los Angeles, CA; 8 University of<br />

Michigan, Ann Arbor, MI<br />

Background: Drug-induced liver injury (DILI) is a diagnosis of<br />

exclusion, and requires a high degree of clinical suspicion.<br />

The aim of this study was to better understand the features


1150A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and diagnoses of patients who were initially thought to have<br />

DILI, which was later judged to be unlikely. Methods: Between<br />

2004 and 2014, the Drug Induced Liver Injury Network (DILIN)<br />

Prospective Study recruited and assessed 1405 patients suspected<br />

to have DILI. All patients underwent evaluation for<br />

viral hepatitis (A, B, C, CMV, E, HSV), autoimmune hepatitis,<br />

pancreaticobiliary disease, and alcohol. Cases under went a<br />

structured causality assessment at 6 months of follow-up to be<br />

assigned into 5 categories of likelihood (definite, >95%; highly<br />

likely, 75-94%; probable, 50-74%, possible, 25-49%, unlikely,<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1151A<br />

1933<br />

Genome-wide investigation identifies common SNPs on<br />

Chr9p22.2 as risk factors for liver injury due to sulfamethoxazole/trimethoprim<br />

Thomas J. Urban 1,2 , Paola Nicoletti 3 , Robert J. Fontana 4 , Andrew<br />

Stolz 5 , Ann K. Daly 6 , Guruprasad P. Aithal 7 , M. I. Lucena 8 , Huiman<br />

X. Barnhart 9 , Paul Watkins 2 , Naga P. Chalasani 10 ; 1 Eshelman<br />

School of Pharmacy, University of North Carolina at Chapel<br />

Hill, Chapel Hill, NC; 2 Hamner Institute for Drug Safety Sciences,<br />

Durham, NC; 3 Columbia University, New York, NY; 4 University<br />

of Michigan, Ann Arbor, MI; 5 University of Southern California,<br />

Los Angeles, CA; 6 University of Newcastle, Newcastle upon Tyne,<br />

United Kingdom; 7 Nottingham University, Nottingham, United<br />

Kingdom; 8 University of Malaga, Malaga, Spain; 9 Duke University,<br />

Durham, NC; 10 Indiana University School of Medicine, Indianapolis,<br />

IN<br />

BACKGROUND &AIM: Treatment with the antibiotic combination<br />

sulfamethoxazole/trimethoprim (SMZ/TMP) can result in a<br />

rare but serious liver toxicity which typically presents with immunoallergic<br />

features. Currently there are no genetic or non-genetic<br />

predictors of toxicity. We aimed to uncover genetic risk<br />

factors associated with SMZ/TMP DILI. METHODS: Genomewide<br />

genotyping was performed on 27 individuals of European<br />

ancestry who experienced liver injury from SMZ/TMP, including<br />

22 DILIN patients and 5 from iDILIC/iSAEC. Genotype<br />

frequencies were compared to >5,000 population controls.<br />

Targeted genotyping by TaqMan assay was performed for validation<br />

in 33 cases (19 cases included in the GWAS and 14<br />

new cases not included in the GWAS). RESULTS: GWAS was<br />

conducted on 27 individuals with SMZ/TMP DILI with mean<br />

age 47 years, 48% females, and the pattern of liver injury was<br />

hepatocellular in 41% and cholestatic in 30%. Their peak ALT<br />

(mean ± sd) were 714 ± 542 U/L and peak bilirubin 11.5±<br />

10.1 mg/dL. The GWAS revealed a set of common SNPs in an<br />

intergenic region on Chr9p22.2 that showed strong evidence<br />

of association with SMZ/TMP-DILI. The minor allele frequency<br />

(MAF) of the risk allele was 26% in SMZ/TMP-DILI cases, vs.<br />

4.6% in controls (OR: 9.0, 95%CI: 4.3-19.1, p = 5 X 10 -9 ).<br />

Validation cohort consisted of 33 European Caucasians (19<br />

from the GWAS plus 14 independent replication samples) with<br />

mean age 46 years, 55% females and the pattern of liver injury<br />

was hepatocellular in 28% and cholestatic in 32%. Their peak<br />

ALT values were 745 ± 695 U/L and bilirubin 13.1± 10.1 mg/<br />

dL. Targeted assays confirmed the genotypes as measured by<br />

the GWAS arrays, and further identified 4 additional carriers<br />

among 14 new cases in the validation cohort (MAF=14.3%,<br />

p < 0.05). However, the phenotypic characteristics of DILI in<br />

patients who carried the risk alleles (age, sex, latency to onset,<br />

immunoallergic features, peak laboratory test values) were not<br />

significantly different from those who lacked the risk alleles at<br />

Chr9p22.2. CONCLUSIONS: Common SNPs on Chr9p22.2<br />

are strongly associated with risk for SMX/TMP-DILI. These<br />

SNPs are not located near any known protein-coding gene or<br />

miRNA, and have not previously been associated with mRNA<br />

expression of any gene. Further <strong>studies</strong> will be necessary to<br />

understand the mechanism underlying the association, and to<br />

extend these <strong>studies</strong> to other hypersensitivity reactions to SMZ/<br />

TMP as well as to DILI caused by other sulfonamides.<br />

Disclosures:<br />

Robert J. Fontana - Consulting: GlaxoSmithKline, CLDF; Grant/Research Support:<br />

Gilead, vertex, BMS, Jansen, Gilead<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

The following authors have nothing to disclose: Thomas J. Urban, Paola Nicoletti,<br />

Andrew Stolz, Ann K. Daly, Guruprasad P. Aithal, M. I. Lucena, Huiman X.<br />

Barnhart, Paul Watkins<br />

1934<br />

Co-activation of AKT and c-Met triggers rapid hepatocellular<br />

carcinoma development via mTORC1/FASN<br />

pathway in mice<br />

Junjie Hu 2,1 , Lei Li 3,1 , Li Che 1 , Diego Calvisi 4 , Xin Chen 1 ; 1 Bioengineering<br />

and Therapeutic Sciences, UCSF, San Francisoco,<br />

CA; 2 School of Pharmacy, Hubei University of Chinese Medicine,<br />

Wuhan, China; 3 School of Pharmacy, Huazhong university of science<br />

and technology, Wuhan, China; 4 Institude of Pathology, university<br />

of greifswald, Greifswald, Germany<br />

Background: Activation of the AKT/mTOR cascade and overexpression<br />

of c-Met have been implicated in the development<br />

of human hepatocellular carcinoma (HCC). Methods: To elucidate<br />

the functional crosstalk between the AKT and c-Met pathways,<br />

we generated a model characterized by the combined<br />

expression of activated AKT and c-Met in the mouse liver using<br />

hydrodynamic transfection. Results: We found that co-expression<br />

of AKT and c-Met results in faster liver tumor development<br />

when compared with mice overexpressing AKT alone, whereas<br />

c-Met alone does not lead to tumor formation. At the molecular<br />

level, liver tumors induced by AKT/c-Met display activation of<br />

AKT/mTOR and Ras/MAPK cascades as well as increased<br />

lipogenesis and glycolysis. Mechanistically, AKT/c-Met driven<br />

hepatocarcinogenesis was found to be mTORC1-dependent, as<br />

Raptor genetic deletion abrogates tumor development in AKT/<br />

c-Met mice. Furthermore, since increased de novo synthesis of<br />

fatty acid is a pivotal metabolic event downstream of mTORC1,<br />

we determined the requirement of lipogenesis in AKT/c-Met<br />

driven liver tumor development using conditional Fatty Acid<br />

Synthase (FASN) knockout mice. Of note, hepatocarcinogenesis<br />

induced by AKT/c-Met was fully inhibited by FASN ablation.<br />

In human HCC samples, coordinated expression of FASN,<br />

c-Met, activated AKT and ERK proteins was detected in a subgroup<br />

of biologically aggressive tumors. Conclutions: Our study<br />

demonstrates that co-activation of AKT and c-Met induces HCC<br />

development that depends on the mTORC1/FASN pathway.<br />

Suppression of mTORC1 and/or FASN might be highly detrimental<br />

for the growth of human HCC subsets characterized by<br />

concomitant induction of the AKT and c-Met cascades.<br />

Disclosures:<br />

The following authors have nothing to disclose: Junjie Hu, Lei Li, Li Che, Diego<br />

Calvisi, Xin Chen<br />

1935<br />

Classification of tumors based on the integrated profile<br />

of genetic and epigenetic alterations and the biological<br />

behavior of human hepatocellular carcinoma<br />

Naoshi Nishida 1 , Toshimi Kaido 2 , Masatoshi Kudo 1 ; 1 Department<br />

of Gastroenterology and Hepatology, Kinki University School of<br />

Medicine, Osaka-sayama, Japan; 2 Department of Surgery, Graduate<br />

School of Medicine Kyoto University, Kyoto, Japan<br />

Aims: We tried to classify HCCs based on the integrated<br />

genetic and epigenetic profile, and examined the association<br />

between the classification and biological characteristics<br />

such as high metastatic potential. Methods: (1) A total of 121<br />

HCCs were examined for p53, β-catenin and TERT promoter<br />

mutation, methylation of 8 tumor suppressor genes (TSGs)<br />

and hypomethylation on 3 repetitive DNA sequences (RSs),<br />

and fractional allelic loss (FAL) as a surrogate marker of chromosomal<br />

instability (CI) using direct sequencing, MethyLight,<br />

and microsatellite analyses with 200 markers, respectively. A<br />

degree of TSG methylation and hypomethylation of RSs was<br />

categorized as “extensive” vs. “limited” and “significant” vs.<br />

“slight” by hierarchical clustering, respectively. Each tumor was


1152A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

classified according to the integrated molecular profiles using<br />

corresponding and hierarchical clustering analyses, and their<br />

relationship to the clinicopathological backgrounds were examined.<br />

(2) Molecular profiles of 44 HCCs from the patients who<br />

underwent liver transplantation (LT) were determined based<br />

on the mutation, methylation and microsatellite analyses. We<br />

tried to examine the specific molecular features that related to<br />

the extrahepatic recurrence after LT of HCC. Results: (1) We<br />

successfully classified HCCs in to 4 subgroups based on the<br />

genetic and epigenetic alterations. The subgroup A1 (S-A1)<br />

was characterized by high frequencies of extensive TGS methylation,<br />

β-catenin and TERT promoter mutation; however, the<br />

molecular characteristics of the subgroup B2 (S-B2) were the<br />

presence of significant hypomethylation on RSs and p53 mutation<br />

accompanied by CI, which was mutually exclusive to that<br />

of S-A1. Similarly, the subgroup A2 (S-A2) was characterized<br />

by the presence of all the alterations examined. On the other<br />

hand, genetic and epigenetic alterations were rare in HCCs in<br />

the subgroup B1 (S-B1). Characteristics related to tumor progression,<br />

such as high serum AFP (> 200 ng/ml), larger tumor<br />

size (> 3 cm), presence of vascular invasion, and dedifferentiated<br />

tumor were more frequently observed in HCCs belonging<br />

to S-A2 and S-B2 than those in S-A1 and S-B1. (2) Among 44<br />

HCC patients after LT, 6 showed metastatic recurrence of HCC<br />

after LT and 38 had not shown any recurrence for more than 2<br />

years. Five of 6 HCC patients (83%) showing recurrence carried<br />

tumors with S-B2 molecular characteristics, whereas only<br />

11 of 38 HCCs in non-recurrence group (29%) showed S-B2<br />

characteristics. Conclusion: Integrated genetic and epigenetic<br />

profiles are associated with tumor characteristics and predict<br />

mild or aggressive behavior of HCCs.<br />

Disclosures:<br />

Masatoshi Kudo - Advisory Committees or Review Panels: Bayer HealthCare;<br />

Grant/Research Support: Bayer HealthCare<br />

The following authors have nothing to disclose: Naoshi Nishida, Toshimi Kaido<br />

1936<br />

Baicalein targets liver tumor initiating stem cell-like cells<br />

resistant to mTORC1 inhibition<br />

Raymond Wu 1 , Ramachandran Murali 2 , Yasuaki Kabe 3 , Keigo<br />

Machida 1 , Clay Wang 4 , Stan G. Louie 4 , Hidekazu Tsukamoto 1 ;<br />

1 Southern California Research Center for ALPD and Cirrhosis, Keck<br />

School of Medicine, University of Southern California, Los Angeles,<br />

CA; 2 Cedars-Sinai Medical Center, Los Angeles, CA; 3 Keio University<br />

School of Medicine, Tokyo, Japan; 4 Department of Pharmacology<br />

and Pharmaceutical Sciences, School of Pharmacy of the<br />

University of Southern California, Los Angeles, CA<br />

[Introduction] mTORC1 inhibitors are tested for the treatment<br />

of HCC based on hyperactive mTOR in this malignancy. However,<br />

evidence indicates mTORC1 inhibitors may promote<br />

CD133 upregulation and chemoresistance. CD133+ tumor initiating<br />

stem cell-like cells (TIC) isolated from mouse liver tumors<br />

(PNAS 106:1548; JCI 123:2832) are chemoresistant, and<br />

identification of an approach to abrogate this resistance is critical.<br />

[Objective] We searched for a natural compound which<br />

rescinds TIC’s resistance to mTORC1 inhibition and improves<br />

chemotherapy outcome. [Methods and Results] HPLC/MS and<br />

NMR analyses combined with TIC lethality assay, identified<br />

baicalein (BC) as a bioactive compound in the herbal medicine<br />

Yan Gang Wan shown to suppress DEN-induced liver<br />

tumorigenesis in mice. BC dose-dependently inhibits TIC self-renewal<br />

shown by spheroid formation and clonogenic assay<br />

and induces TIC apoptosis at IC50=9.5mM while having no<br />

adverse effects on mouse primary hepatocytes. BC represses<br />

stemness genes (Nanog, Sox2) while upregulating hepatocyte<br />

differentiation genes (Alb, Cyp7a1). Structure-function analysis<br />

reveals –OH groups on C5 and C7 are required for TIC killing.<br />

TIC which are resistant to rapamycin (2mM) and temsirolimus<br />

(~4.5mM), are killed ~90% when co-treated with BC (10mM).<br />

BC also enhances TIC cytotoxicity by sorafenib (3.3mM) and<br />

doxorubicin (1mM), causing ~95% elimination. BC suppresses<br />

autophagosome formation determined by LC3II IB and LC3-<br />

GFP analysis and mitochondrial respiration by Seahorse assay,<br />

and prevents temsirolimus-mediated CD133 induction. Pharmacokinetic<br />

<strong>studies</strong> reveal orally administered BC is converted to<br />

~50% less potent baicalin (BCi) with glucuronate substituting<br />

C7 -OH in plasma but largely reverted to BC in the mouse<br />

liver. A similar reversion is confirmed in cultured primary<br />

hepatocytes and TIC treated with BCi. BC treatment completely<br />

eliminates temsirolimus chemoresistance of TIC-derived subcutaneous<br />

tumor growth in nude mice. LC/MC analysis of TIC<br />

proteins pulled-down with BCi-conjugated nanobeads, identifies<br />

Rab1 and other Rab family members as potential targets of<br />

BC, and 3D structural analysis and docking <strong>studies</strong> predicted<br />

BC interaction at the GTP binding site of Rab1 via BC’s –OH<br />

of C5 and C7. Indeed, GTP but not ATP competitively inhibits<br />

BC-Rab1 interaction and Rab1 overexpression rescues BC-induced<br />

inhibition of autophagy, but not TIC killing, suggesting<br />

a role for additional Rab protein target(s) for the latter effect.<br />

[Conclusion] BC promotes killing of chemoresistant TICs, particularly<br />

when combined with mTORC1 inhibitor via mechanisms<br />

which appear to involve interactions with Rab family proteins.<br />

Disclosures:<br />

Hidekazu Tsukamoto - Consulting: Suntory Ltd.; Grant/Research Support: The<br />

Toray Co.<br />

The following authors have nothing to disclose: Raymond Wu, Ramachandran<br />

Murali, Yasuaki Kabe, Keigo Machida, Clay Wang, Stan G. Louie<br />

1937<br />

CD26 (DPP-4) as a molecular target for HCC treatment<br />

Sohji Nishina 1 , Akira Yamauchi 2 , Yuichi Hara 1 , Keisuke Hino 1 ;<br />

1 Hepatology and Pancreatology, Kawasaki medical univercity,<br />

Kurashiki city, Okayama, Japan; 2 Biochemistry, Kawasaki Medical<br />

School, Kurashiki city, Okayama, Japan<br />

Background and Aim: CD 26 is a multifunctional transmembrane<br />

glycoprotein and functions as dipeptidyl peptidase 4<br />

(DPP-4). Although CD26 is expressed in various cancers, the<br />

relationship between hepatocellular carcinoma (HCC) progression<br />

and CD26 expression remains unknown. The aim of this<br />

study was to investigate the potential role of CD26 as a molecular<br />

target for HCC treatment. Methods: CD26 expression was<br />

examined in 41 surgically resected liver specimens from patients<br />

with HCC. In vitro the effect of DPP-4 inhibitor, anagliptin, on<br />

cancer cell growth was investigated using Huh7/HepG2 cells<br />

that expresses CD26. In vivo nude mice (BALBc-nu/nu) were<br />

subcutaneously injected Huh7 cells and then fed control diet,<br />

low-dose anagliptin containing diet or high-dose anagliptin<br />

containing diet for 21 days. Results: CD26 expression was<br />

correlated with cell proliferation, angiogenesis and cell differentiation<br />

in HCC specimens obtained from patients. Anagliptin<br />

did not affect cell proliferation or cell cycle in vitro. However<br />

in nude mice, anagliptin significantly suppressed the growth of<br />

xenograft tumors in dose dependent manner. Anagliptin also<br />

induced NK cells infiltrations more vigorously and reduced<br />

angiogenesis in xenograft tumors, even though body weight<br />

and dietary intake through the feeding period, and glucose tolerance<br />

determined at the 14th day of feeding were not different<br />

among three groups. These results suggested that anagliptin<br />

potentially have antitumor activity against HCC. Furthermore,<br />

we examined whether anagliptin activated chemotaxis of NK


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1153A<br />

cells, since NK cells were more vigorously infiltrated in mice fed<br />

the anagliptin containing diet. NK cells obtained human volunteers<br />

were examined for their chemotaxis using EZ-TAXIScan in<br />

the presence of CXCL10, chemokine for NK cells, with or without<br />

anagliptin. Interestingly, anagliptin significantly activated<br />

the chemotaxis of NK cells probably through the reduction<br />

CXCL10 antagonism, because it has been shown that CXCL10<br />

is truncated at its N-terminus through DPP-4 activity and that a<br />

N-terminal truncated CXCL10 acts as an antagonist form. Conclusions:<br />

The present results indicated that CD26 expression<br />

was related to tumor progression in patients with HCC and<br />

that DPP-4 inhibitor, anagliptin potentially suppressed HCC<br />

progression through activation of chemotaxis of NK cells, and<br />

inhibition of angiogenesis.<br />

Disclosures:<br />

Akira Yamauchi - Grant/Research Support: Kobayashi Pharmaceutical Co., Ltd.<br />

The following authors have nothing to disclose: Sohji Nishina, Yuichi Hara,<br />

Keisuke Hino<br />

1938<br />

Liver Selective Acetyl-CoA Carboxylase Inhibitor<br />

ND-654 Improves Sorafenib Efficacy in the Treatment of<br />

Hepatocellular Carcinoma in Cirrhotic Rats<br />

Lan Wei 1 , Omeed Moaven 1 , Geraldine Harriman 2 , Jeremy R.<br />

Greenwood 3 , Sathesh P. Bhat 3 , William Westlin 2 , H. James Harwood<br />

2 , Rosana Kapeller 2 , Kenneth K. Tanabe 1 , Bryan C. Fuchs 1 ;<br />

1 Surgery, Massachusetts General Hospital, Boston, MA; 2 Nimbus<br />

Therapeutics, Cambridge, MA; 3 Schrodinger, New York, NY<br />

Background: Hepatocellular carcinoma (HCC) is increasing in<br />

incidence worldwide. Current treatment options for HCC are<br />

limited, and as such, prognosis is extremely poor with a 5-year<br />

survival less than 12%. Sorafenib is the only FDA-approved<br />

drug for the treatment of HCC but its effects are marginal as it<br />

only extends survival by a few months. Metabolic attenuation<br />

is a promising approach to cancer therapy, and rate-limiting<br />

steps in key biosynthetic pathways are particularly attractive<br />

targets. We recently identified ND-654, a hepatoselective<br />

(~3000:1 liver to muscle exposure), allosteric ACC inhibitor<br />

that binds to the ACC subunit dimerization site and inhibits the<br />

enzymatic activity of both ACC1 (IC50 = 3 nM) and ACC2<br />

(IC50 = 8 nM). We previously demonstrated that daily oral<br />

administration of 10 mg/kg ND-654 reduced tumor incidence<br />

by 55% and significantly improved median survival time in a<br />

rat model of cirrhosis and HCC. Here, we examine the effects<br />

of ND-654 alone and in combination with sorafenib on HCC<br />

development in cirrhotic rats. Methods: Male Wistar rats were<br />

treated once weekly with 50 mg/kg diethylnitrosamine (DEN)<br />

for 18 weeks to induce sequential development of fibrosis,<br />

cirrhosis and HCC. After establishment of cirrhosis and when<br />

HCCs are first developing (13 weeks), rats were treated daily<br />

by oral gavage with either 1) vehicle control, 2) 10 mg/kg<br />

ND-654, 3) 10 mg/kg sorafenib, or 4) 10 mg/kg ND-654<br />

and 10 mg/kg sorafenib. After 18 weeks, tumor nodules<br />

were counted and liver and tumor tissue was harvested for<br />

analysis. Results: Similar to our previous study, simultaneous<br />

inhibition of ACC1 and ACC2 significantly reduced HCC incidence<br />

by 41% (p < 0.05) which was comparable to results<br />

with sorafenib (57% reduction, p < 0.01). Remarkably, the<br />

combination of ND-654 and sorafenib significantly reduced<br />

HCC incidence by 81% (p < 0.001). Our previous results had<br />

shown that ND-654 decreased tumor proliferation and induced<br />

necrosis and we are currently examining the combination of<br />

ND-654 and sorafenib on these endpoints. Conclusions: These<br />

results provide further evidence that de novo lipogenesis is an<br />

important mediator of hepatic carcinogenesis and that selective<br />

inhibition of hepatic ACC is a viable cancer metabolism<br />

therapeutic strategy for treating HCC that could be combined<br />

with sorafenib.<br />

Disclosures:<br />

Jeremy R. Greenwood - Employment: Schrodinger Inc.; Patent Held/Filed: Nimbus<br />

Therapeutics; Stock Shareholder: Schrodinger Inc.<br />

Rosana Kapeller - Employment: Nimbus Therapeutics; Stock Shareholder: Aileron<br />

Therapeutics<br />

Kenneth K. Tanabe - Patent Held/Filed: EGF SNP and risk for HCC, EGFR inhibition<br />

and HCC, gene signature for prognosis in cirrhosis<br />

Bryan C. Fuchs - Consulting: Collagen Medical; Grant/Research Support: Nimbus<br />

Therapeutics<br />

The following authors have nothing to disclose: Lan Wei, Omeed Moaven, Geraldine<br />

Harriman, Sathesh P. Bhat, William Westlin, H. James Harwood<br />

1939<br />

The Long Noncoding RNA SNHG6 Functions As a Competing<br />

Endogenous RNA to Promote Hepatocellular Carcinoma<br />

Progression<br />

LI LIU 1 , Chuanhui Cao 2 , Ting Ting Zhang 1 , Dongyan Zhang 2 ,<br />

Dehua Wu 2 ; 1 Institution of hepatology, Guangzhou, China;<br />

2 Department of Radiation Oncology, Guangzhou, China<br />

Abstract: Expression of long noncoding RNAs (lncRNAs) has<br />

been reported to be dysregulated in hepatocellular carcinoma<br />

(HCC) but their contributions and functional mechanisms to<br />

HCC remain largely unknown. Here, we elucidate whether<br />

lncRNA SNHG6 is involved in HCC progression.Our results<br />

showed that SNHG6 was up-regulated in HCC in 160 paired<br />

cancerous and noncancerous tissues by quantitative real-time<br />

PCR and in situ hybridization. SNHG6 associated with portal<br />

vein tumor thrombus, tumor stage, metastasis, and shorter overall<br />

survival of HCC patients. Ectopic expression of the SNHG6<br />

in HCC cells promoted their proliferation and inhibited apoptosis<br />

both in vitro and in vivo. Furthermore, we revealed an<br />

endogenous interaction between SNHG6 and miR-26a/b by<br />

coimmunoprecipitation with anti-Ago2 in HCC cells. More<br />

importantly, SNHG6 controlled the expression of miR-26a/b<br />

target gene, TAK1. In our in vitro and in vivo system, and<br />

clinical HCC tissues, we observed that SNHG6 positively correlated<br />

with TAK1 and ectopic expression of SNHG6 was sufficient<br />

to increase TAK1. Moreover, knockdown TAK1 could<br />

arrest HCC cell proliferation inducing by overexpression of<br />

SNHG6, which was consistent with results of knockdown of<br />

SNHG6 expression in HCC cells. Notably, there also exists<br />

co-expression relationship between SNHG6 and TAK1 in liver,<br />

colon, rectum, and pancreas cancers.CONCLUSIONS: SNHG6<br />

may act as a biomarker of poor prognosis in HCC and confer<br />

malignant phenotype to tumor cells. The ceRNA network involving<br />

SNHG6 may contribute to a better understanding of liver<br />

carcinogenesis and tumor progression.


1154A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

follow-up, association of OPN mRNA expression in liver or in<br />

tumor tissue with pattern of HCC recurrence was determined.<br />

Male wild-type (WT) littermates, Opn global knockout (Opn -/- )<br />

and transgenic mice overexpressing Opn in hepatocytes in<br />

either WT (Opn Hep Tg) or in Opn -/- background (Opn -/-Hep Tg)<br />

were injected diethylnitrosamine (DEN) at 15 d of age and<br />

were followed for the development of liver tumors up to 1 yr.<br />

Upon sacrifice, histopathological scoring, immunohistochemistry<br />

for OPN, ELISA for serum and liver α-fetoprotein (AFP) and<br />

serum transaminases were analyzed. RESULTS: Elevated OPN<br />

mRNA in non-tumor liver tissue from 199 HBV-related HCC<br />

patients was associated with late tumor recurrence (>2 yrs<br />

after surgery, de novo carcinogenesis from remnant diseased<br />

liver) but not with early recurrence (dissemination/metastasis<br />

of resected primary tumor). The same trend was not observed<br />

in a second cohort of patients with HCV-related HCC thus suggesting<br />

HBV specificity. In contrast, OPN mRNA in tumor tissue<br />

correlated with early but not with late recurrence, indicating a<br />

dual oncogenic role of OPN according to the site of expression.<br />

To understand the mechanism whereby intrahepatic OPN<br />

and specifically hepatocyte OPN contributed to HCC, a wellknown<br />

mouse model mimicking HCC was used. After 1 yr<br />

of DEN injection, Opn -/- mice exhibited significantly reduced<br />

macroscopic HCC development, less number of HCC nodules<br />

per sample and lower serum and hepatic AFP levels compared<br />

to WT littermates. Yet, overexpression of Opn in hepatocytes<br />

rescued the phenotype from Opn -/- mice suggesting a key role<br />

for hepatocyte-derived OPN in HCC development, which was<br />

further confirmed by the leading incidence of tumors, the largest<br />

tumor size and the highest AFP in Opn Hep Tg mice. CON-<br />

CLUSIONS: 1) Liver and more importantly hepatocyte-derived<br />

OPN is a major player orchestrating the development of HCC;<br />

2) OPN specifically plays a role in HBV-related HCC and 3)<br />

Ablation or reduction of hepatocyte-derived OPN may be a<br />

useful strategy to prevent the onset and/or the progression to<br />

HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yu Chen, Xiaodong Ge, Ioana<br />

Abraham-Enachescu, Xiaochen Sun, Grace Guzman, Yujin Hoshida, Natalia<br />

Nieto<br />

Disclosures:<br />

The following authors have nothing to disclose: LI LIU, Chuanhui Cao, Ting Ting<br />

Zhang, Dongyan Zhang, Dehua Wu<br />

1940<br />

Hepatocyte-derived osteopontin promotes the development<br />

of hepatocellular carcinoma<br />

Yu Chen 1 , Xiaodong Ge 1,2 , Ioana Abraham-Enachescu 2 , Xiaochen<br />

Sun 2 , Grace Guzman 1 , Yujin Hoshida 2 , Natalia Nieto 1,2 ; 1 Pathology,<br />

University of Illinois at Chicago, Chicago, IL; 2 Department of<br />

Medicine, Icahn School of Medicine at Mount Sinai, New York,<br />

NY<br />

BACKGROUND & HYPOTHESIS: Previous work from our laboratory<br />

demonstrated that intrahepatic osteopontin (OPN) is highly<br />

profibrogenic. Since fibrosis progression is a major cause of<br />

hepatocellular carcinoma (HCC) and the role of OPN in the<br />

pathogenesis of HCC is unknown, we hypothesized that the<br />

presence of OPN in the liver microenvironment and specifically<br />

in hepatocytes could promote the development of HCC. METH-<br />

ODS: To prove this hypothesis, we analyzed clinical cohorts of<br />

HCC patients and a mouse model of HCC with genetic manipulation<br />

of Opn. In HCC patients’ cohorts with long-term clinical<br />

1941<br />

Efficacy and mechanism of dendritic cell-based immunotherapy<br />

in combination with T cell immunoglobulin and<br />

mucin protein 3 blockade in a murine hepatocellular<br />

carcinoma model<br />

Junichi Eguchi 1 , Eiichi Hayashi 2 , Hiroyoshi Doi 2 , Risa Omori 2 ,<br />

Atsushi Kajiwara 2 , Hitoshi Yoshida 2 , Takayoshi Ito 1 , Yoshio Deguchi<br />

1 , Norihiro Nomura 1 , Haruhiro Inoue 1 ; 1 Digestive Disease<br />

Center, Showa University Koto-Toyosu Hospital, Tokyo, Japan;<br />

2 Division of Gastroenterology, Department of Medicine, Showa<br />

University School of Medicine, Tokyo, Japan<br />

Progress in treatments for hepatocellular carcinoma (HCC) has<br />

improved the prognosis of patients with HCC. However, HCC<br />

is usually associated with cirrhosis and often recurs even after<br />

complete treatment of the tumors in the remaining part of the<br />

cirrhotic liver. Thus, there is a strong need for the development<br />

of a new intervention therapy that suppresses the occurrence or<br />

recurrence effectively with fewer side effects. Immunotherapy<br />

may be such a treatment and several clinical trials have been<br />

performed for HCC treatment. Dendritic cells (DCs) are known<br />

as professional antigen-presenting cells characterized by their<br />

potent ability to elicit immune responses to tumor-specific T<br />

cells against tumor-associated antigens. T cell immunoglobulin<br />

and mucin protein 3 (TIM-3), has been identified as a marker


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1155A<br />

of immunosuppressive statement. Interaction of TIM-3 with its<br />

ligand, galectin 9, triggers cell death in activated T cells. In this<br />

study, we evaluated the efficacy of the combination of DC-based<br />

vaccine and TIM-3 blockade, and investigated the mechanisms<br />

of the antitumor effects of the combined therapy. In protection<br />

model, mice were injected with DCs or/and TIM-3-antibody<br />

before the murine HCC tumor cell (BNL cell) challenge. 50% of<br />

mice treated with both DCs and TIM-3-antibody rejected tumor<br />

challenge, whereas other groups observed a palpable tumor in<br />

all mice tested. In therapeutic model, tumor-bearing mice were<br />

inoculated with DCs or/and TIM-3-antibody. Significant suppression<br />

of outgrowth of the established tumors was observed<br />

in the DCs and anti-TIM-3 combination treatment group (DCs +<br />

anti-TIM-3, 265.3 ± 53.09 mm 2 vs control, 535.1 ± 44.47 mm 2<br />

on day 30, P = 0.0046 vs controls). To investigate induction of<br />

tumor-specific immune responses, we stimulated splenocytes of<br />

DCs or/and TIM-3-antibody treated mice twice weekly by DCs<br />

in vitro. High tumor-specific immune response was detected<br />

when splenocytes of mice treated with both DCs and anti-TIM-3<br />

were used as effector cells in interferon-gamma enzyme-linked<br />

immunospot assay. In addition, incubation with TIM-3-antibody<br />

enhanced expression of IL-12p70, and cell surface expression<br />

of CD80 in bone marrow-derived DCs. Our findings suggest<br />

that blockade of the TIM-3 enhanced the maturation of DCs,<br />

and Th1-type antitumor immune responses induced by DC vaccine.<br />

The combination of DC vaccine and TIM-3 blockade may<br />

be a possible candidate for a cancer vaccine for clinical trials.<br />

Disclosures:<br />

The following authors have nothing to disclose: Junichi Eguchi, Eiichi Hayashi,<br />

Hiroyoshi Doi, Risa Omori, Atsushi Kajiwara, Hitoshi Yoshida, Takayoshi Ito,<br />

Yoshio Deguchi, Norihiro Nomura, Haruhiro Inoue<br />

1942<br />

PTEN Expression Affects the Outcome of Human Hepatocellular<br />

Carcinoma via Modulating the Oncogenic<br />

Behaviors and the angiogenic Processes<br />

Tsung-Hui Hu 1 , Po-Lin Tseng 1 , Pey-Ru Lin 1 , Ming-Hong Tai 2 , Chao-<br />

Cheng Huang 3 , Li-Na Yi 1 , Chih-Chi Wang 4 , Kuo-Chin Chang 1 ,<br />

Yi-Hao Yen 1 , Ming-Chao Tsai 1 , Ming-Tzung Lin 1 ; 1 Division of<br />

Hepato-Gastroenterology, Department of Internal Medicine,,<br />

Kaohsiung Chang Gung Memorial Hospital, Kaohsiung, Taiwan;<br />

2 Institute of Biological Sciences, National Sun Yat-Sen University,<br />

Kaohsiung, Taiwan; 3 Department of Pathology, Kaohsiung Chang<br />

Gung Memorial Hospital, Kaohsiung, Taiwan; 4 Department of<br />

Surgery, Kaohsiung Chang Gung Memorial Hospital, Kaohsiung,<br />

Taiwan<br />

Reduced phosphatase and tensin homolog (PTEN) expression<br />

is associated with vascular endothelial growth factor (VEGF)<br />

overexpression in various malignancies. Herein, we aimed to<br />

investigate the expression of PTEN and VEGF in pericancerous<br />

tissues and their association in hepatocellular carcinoma<br />

(HCC). PTEN expression was reduced in 43% and VEGF was<br />

overexpressed in 31% of 113 resected HCC clinical specimens.<br />

VEGF overexpression is positively correlated with young<br />

age, male gender, hepatitis B viremia, high α-fetoprotein<br />

levels, decline in PTEN expression, advanced tumor stage,<br />

and dedifferentiation of HCC. Survival analysis revealed that<br />

reduced PTEN expression, microvessel density, and advanced<br />

tumor stage were independent poor prognostic factors for disease-free<br />

survival. Adenovirus-mediated restoration of PTEN<br />

inhibited cell proliferation, clonogenic growth and invasive<br />

capability of PTEN-deficient human HCC cells in vitro. In xenograft<br />

model, PTEN gene delivery could efficiently suppress the<br />

incidence and growth rate of implanted tumor cells and reduce<br />

microvessel density. The expression of VEGF and HIF1a, and<br />

secreted IL-8 were decreased in PTEN restored HCC cells.<br />

Angiogenic networks of human umbilical vein endothelial cells<br />

were also suppressed by the increased PTEN in vitro. Conclusion:<br />

The tumor microenvironment may be affected by PTEN<br />

expression through modulating the oncogenic phenotypes of<br />

tumor cells and having a tendency to interfere in angiogenic<br />

processes, including expression of angiogenic factors and formation<br />

of angiogenic networks. The expression of PTEN and<br />

VEGF affecting the outcomes of HCC patients might serve as<br />

good prognostic biomarkers.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tsung-Hui Hu, Po-Lin Tseng,<br />

Pey-Ru Lin, Ming-Hong Tai, Chao-Cheng Huang, Li-Na Yi, Chih-Chi Wang, Kuo-<br />

Chin Chang, Yi-Hao Yen, Ming-Chao Tsai, Ming-Tzung Lin<br />

1943<br />

Inhibition of autophagy reverses the resistance to<br />

sorafenib induced by hypoxia in hepatocellular carcinoma<br />

Takayuki Kogure, Yasuteru Kondo, Jun Inoue, Yu Nakagome, Yuta<br />

Wakui, Tatsuki Morosawa, Yasuyuki Fujisaka, Teruyuki Umetsu,<br />

Tooru Shimosegawa; Tohoku University Hospital, Sendai, Japan<br />

Backgrounds: Sorafenib is the only effective systemic therapy<br />

for hepatocellular carcinoma (HCC). It shows cytotoxicity on<br />

HCC cells in vitro but its clinical response is disease stabilization.<br />

Sorafenib shows inhibition of angiogenesis, which could<br />

lead the tumor microenvironment to severe hypoxia. It has been<br />

reported that hypoxia is closely related to resistance to many<br />

types of anticancer drugs. We hypothesized that hypoxia<br />

induces resistance to sorafenib through activation of autophagy,<br />

which results in clinically modest efficacy of sorafenib.<br />

We investigated the involvement of autophagy in the resistance<br />

to sorafenib in HCC. Methods: HCC cell lines, PLC/PRF/5<br />

(PLC5), Li7, HepG2, and Huh7 were cultured in 95% air and<br />

5% CO 2<br />

(normoxia). For hypoxia, cells were transferred to a<br />

hypoxia chamber with 1% O 2<br />

, 5% CO 2<br />

, and 94% N 2<br />

. Proliferation<br />

of HCC cells was assessed with growth curve assay<br />

and MTS assay. Apoptosis was assessed morphologically and<br />

with activation of caspase-3/7 using a luminometric assay.<br />

Hypoxia inducible factor 1α (HIF1α) protein was detected by<br />

immunoblotting. Involvement of autophagy was assessed by<br />

detecting LC3 and p62 and by treating cells with chloroquine<br />

(CQ), an inhibitor of autophagy. Results: After 24 hr of cell<br />

culture in hypoxia, PLC5 and Li7 showed significant increase of<br />

HIF1α but no significant change was observed in growth curve<br />

through 96 hr of culture. Sorafenib inhibited cell growth in a<br />

concentration-dependent manner and this effect was reduced<br />

under hypoxic condition with an IC 50<br />

(normoxia/hypoxia, μM)<br />

of 5.9/9.4 in PLC5, 7.4/11.4 in Li7, 12.1/14.5 in Huh7, and<br />

6.9/11.7 in HepG2. Induction of apoptosis was suppressed<br />

under hypoxia. Incubation with sorafenib (5 and 10 μM) for<br />

24 hr induced apoptosis by 14.6% and 23.8% in PLC5 in normoxia<br />

whereas 10.5% and 13.1% in hypoxia. An increase of<br />

caspse-3/7 activity induced by sorafenib (5 and 10 μM) was<br />

reduced by 77.2% and 55.3% under hypoxia. To assess the<br />

involvement of autophagy, protein expression of LC3 and p62<br />

was detected. After 48 hr of culture in hypoxia, an increase of<br />

LC3-II and a decrease of p62 was observed in PLC5 and Li7,<br />

indicating the activation of autophagy. HCC cells were treated<br />

with sorafenib with or without CQ. Under hypoxia, viability<br />

of cells treated with sorafenib (10 μM, 72 hr) was 55.6% in<br />

PLC5 and 72.1% in Li7. Co-treatment of cells with CQ (50 μM)<br />

recovered the cytotoxicity of sorafenib with a viability of 86.4%<br />

in PLC5 and 89.9% in Li7. Conclusion: Under hypoxia, resistance<br />

to sorafenib is induced by the activation of autophagy.


1156A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Inhibition of autophagy could be an attractive approach for<br />

potentiating the antitumor effect of sorafenib in HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Takayuki Kogure, Yasuteru<br />

Kondo, Jun Inoue, Yu Nakagome, Yuta Wakui, Tatsuki Morosawa, Yasuyuki<br />

Fujisaka, Teruyuki Umetsu, Tooru Shimosegawa<br />

1944<br />

The relationship between hepatocellular carcinoma<br />

development and reduced expression levels of the<br />

major germline genotype of PROSC, which is frequently<br />

deleted in hepatocellular carcinoma<br />

Daiki Miki 1,2 , Hidenori Ochi 1,2 , C. Nelson Hayes 1,2 , Hiromi Abe 1,2 ,<br />

Sakura Akamatsu 1,2 , Atsushi Ono 1,2 , Takashi Nakahara 1,2 , Yizhou<br />

Zhang 1,2 , Keiichi Masaki 1,2 , Hatsue Fujino 1,2 , Eisuke Miyaki 1,2 ,<br />

Hiromi Kan 1,2 , Takuro Uchida 1,2 , Nobuhiko Hiraga 1,2 , Masataka<br />

Tsuge 1,2 , Tomokazu Kawaoka 1,2 , Michio Imamura 1,2 , Yoshiiku<br />

Kawakami 1,2 , Hiroshi Aikata 1,2 , Kazuaki Chayama 1,2 ; 1 Laboratory<br />

for Digestive Diseases, RIKEN Center for Integrative Medical<br />

Sciences, Hiroshima, Japan; 2 Department of Gastroenterology and<br />

Metabolism, Hiroshima University Hospital, Hiroshima, Japan<br />

Background & Aims: In this study, we focused on the recurrently<br />

amplified or deleted genes in hepatocellular carcinoma (HCC)<br />

tissues and examined the association between their baseline<br />

expression levels, which are regulated by single nucleotide<br />

polymorphisms (SNPs), and development of HCC. Methods:<br />

We searched for expression quantitative trait loci (eQTLs) for<br />

previously reported genes with copy number alterations from<br />

a public database of non-transformed peripheral blood samples<br />

that includes over 5,000 individuals. We then conducted<br />

a case-control analysis of these eQTL SNPs or SNPs linked<br />

to them (r 2 > 0.8 in both European and Asian populations)<br />

using our SNP typing data in HCC cases (631 HBV-related<br />

HCC, 554 HCV-related HCC) and 5,489 controls without<br />

any liver diseases or cancers. Results: Among 13 recurrently<br />

amplified genes, we found 10 cis-acting eQTL SNPs (i.e., the<br />

SNP is located near the expressed gene) and 2 trans-acting<br />

eQTL SNPs (i.e., the SNP is located far from the expressed<br />

gene or on another chromosome). Among 18 recurrently<br />

deleted genes, we found 15 eQTL SNPs and 3 trans-acting<br />

eQTL SNPs. Among these 30 eQTL SNPs, we have successfully<br />

genotyped 22 of either the target SNP, when available, or<br />

else a closely linked SNP. We genotyped 9 cis-acting and 1<br />

trans-acting eQTL SNPs for amplified genes, and 9 cis-acting<br />

and 3 trans-acting eQTL SNPs for deleted genes. In case-control<br />

analysis, no significant association was observed between<br />

SNP frequencies and HCC development both in HBV-related<br />

and HCV-related HCC <strong>studies</strong> (vs. controls). Subsequently, we<br />

compared each HCC case to chronic HBV and HCV patients<br />

without HCC (n = 1,997 and 1,354). After adjustment for<br />

multiple testing, we found 2 significantly associated eQTL SNPs<br />

with HCC for PROSC in the HBV study (Odds ratio [OR] =<br />

1.27) and for TNFAIP3 in the HCV study (OR = 1.37), suggesting<br />

their baseline expression levels may be important for<br />

virus-related hepatocarcinogenesis. The risk allele of the eQTL<br />

SNP for HBV-related HCC strongly down-regulates PROSC (P<br />

= 8.25 × 10 -149 ), whereas the risk allele of the eQTL SNP<br />

for HCV-related HCC weakly up-regulates TNFAIP3 (P = 2.60<br />

× 10 -4 ). Furthermore, both PROSC and TNFAIP3 are recurrently<br />

deleted genes in HCC. Taken together, the association<br />

of the eQTL SNP for PROSC with HBV-related HCC susceptibility<br />

seems to be more compatible than that for TNFAIP3 with<br />

HCV-related HCC susceptibility. Conclusions: As a result of<br />

integrative analysis using our large-scale genome-wide SNP<br />

genotype data and public eQTL data, we suggest that germline<br />

variants of recurrently deleted genes may affect viral hepatocarcinogenesis<br />

by regulating their expression levels.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Daiki Miki, Hidenori Ochi, C.<br />

Nelson Hayes, Hiromi Abe, Sakura Akamatsu, Atsushi Ono, Takashi Nakahara,<br />

Yizhou Zhang, Keiichi Masaki, Hatsue Fujino, Eisuke Miyaki, Hiromi Kan, Takuro<br />

Uchida, Nobuhiko Hiraga, Masataka Tsuge, Tomokazu Kawaoka, Michio<br />

Imamura, Yoshiiku Kawakami, Hiroshi Aikata<br />

1945<br />

Steatosis Induces Wnt/Β-Catenin Pathway To Stimulate<br />

Proliferation Of Hepatic Tumor Initiating Cells And Promote<br />

Liver Cancer Development<br />

Anketse D. Kassa 1 , Ni Zeng 1 , Vivian G. Medina 1 , Lina He 1 , Kaiting<br />

Yang 1 , Indra Mahajan 1 , Chien-Yu Chen 1 , Richa Aggarwal 1 ,<br />

Zhechu Peng 1 , Megan Rieger 2 , Chia-Lin Chen 2 , Keigo Machida 2 ,<br />

Zea Borok 2 , Michael Kahn 2 , Bangyan L. Stiles 1,2 ; 1 Pharmacology<br />

and Pharmaceutical Sciences, USC School of Pharmacy, Los Angeles,<br />

CA; 2 Keck School of Medicine, USC, Los Angeles, CA<br />

Obesity is a major compounding factor for liver cancer where<br />

male with a body mass index (BMI) >35 has 4.52 fold higher<br />

incidence of liver cancer vs. those with BMI between 18.5 and<br />

24.9. In mouse models, constitutive activation of the insulin/<br />

PI3K signaling pathway by deletion of the negative regulator<br />

Pten (phosphatase and tensin homologue deleted on chromosome<br />

10) leads to hepatosteatosis and later tumor phenotypes.<br />

In this model (Pten loxP/loxP ; Alb-Cre + ), we have shown that the<br />

presence of the underlying fatty liver condition is required for<br />

the development of liver cancers, which expands from a population<br />

of tumor initiating cells (TICs). In this study, we investigated<br />

how fatty liver disease may contribute to cancer formation.<br />

We show here that blocking fatty liver formation by genetic<br />

or non-genetic approaches significantly reduced the TIC population.<br />

Concomitantly, tumor development in the Pten loxP/loxP ;<br />

Alb-Cre + mice is inhibited despite of the tumor transformation<br />

signals of PTEN loss. We further demonstrated that hepatic<br />

steatosis resulting from Pten deletion or high fat diet feeding<br />

induces the activation of Wnt/b-catenin pathway. Genomic<br />

profiling of human liver cancer specimens has indicated a high<br />

prevalence of b-catenin activating mutation. However, consistent<br />

with mouse <strong>studies</strong> that genetically modify b-catenin,<br />

such mutation is not sufficient to drive malignant transformation.<br />

We show here that Wnt/b-catenin activation signal is the<br />

tumor promoting factor induced by lipid accumulation and that<br />

this signal is critical for the expansion of the transformed liver<br />

TICs. In the Pten loxP/loxP ; Alb-Cre + mice fed low caloric diet to<br />

inhibit hepatic steatosis, reduced expression of Wnt ligands<br />

is observed concurrent with inhibited expansion of TICs and<br />

complete blockage of tumor development. Blocking the signal<br />

of Wnt/b-catenin with shRNA or a pharmacological reagent<br />

ICG-001 inhibits the proliferation of TICs in vitro and lead to<br />

decreased accumulation of TICs and reduction of tumor grafts<br />

in vivo. Together, our study suggests that fatty liver activates the<br />

niche factor Wnt/b-catenin to promote tumorigenesis using the<br />

Pten null liver cancer model.<br />

Disclosures:<br />

The following authors have nothing to disclose: Anketse D. Kassa, Ni Zeng, Vivian<br />

G. Medina, Lina He, Kai-ting Yang, Indra Mahajan, Chien-Yu Chen, Richa<br />

Aggarwal, Zhechu Peng, Megan Rieger, Chia-Lin Chen, Keigo Machida, Zea<br />

Borok, Michael Kahn, Bangyan L. Stiles


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1157A<br />

1946<br />

Truncated Hepatitis B virus X protein up-regulates<br />

CD44+ cancer stem like cells through LTBP1<br />

Goki Suda, Seiji Tsunematsu, Masato Nakai, Jun Itoh, Mitsuteru<br />

Natsuizaka, Kenichi Morikawa, Koji Ogawa, Naoya Sakamoto;<br />

hokkaido university, Sapporo, Japan<br />

Background Aims: Hepatitis B virus (HBV) is a major pathogen<br />

of hepatocellular carcinoma (HCC). HBV X protein (HB) is<br />

reported to promote HCC development by activation of various<br />

signals. Recently genome-wide association study about HBV-related<br />

HCC revealed that COOH-truncated HBx was integrated<br />

in host genome at high frequency. Cancer stem like cells (CSCs)<br />

are defined as those cells that can self-renew, drive tumorigenesis.<br />

It is not clear whether COOH-truncated HBx integration<br />

promotes prevalence of CSCs or not. Therefore, we aimed<br />

to analyze the relationship between COOH-truncated HBx<br />

expression and CSCs. Methods: We firstly established stably<br />

HBx (HBx/HepG2) and COOH-truncated HBx (truncated HB/<br />

HepG2). Next we examined the expression levels of stemness<br />

makers in these established cells by Flow cytometry analysis.<br />

Subsequently we analyzed the anti-tumor agent resistance and<br />

tumorigenesis of these cells by MTS assay and soft gel agar<br />

colony formation assay. Finally we investigated the factors that<br />

promote prevalence of CSCs in truncated-HBx /HepG2 cells<br />

by PCR-array. Results: Both HBx/HepG2 and truncated HBx/<br />

HepG2 shows significantly resistant to anti-tumor agent, 5-FU,<br />

compared with control/HepG2 cells. Soft gel agar colony formation<br />

assay shows that truncated-HBx/HepG2 cells make significantly<br />

more colonies than control/HepG2 cells and HBx/<br />

HepG2 cells. Flow cytometry analysis reveals that truncated<br />

HBx/HepG2 cells express high levels of CD44, one of cancer<br />

stemness maker, compared with control/HepG2 and HBx/<br />

HepG2 cells. Next truncated-HBx cells were sorted in CD44<br />

high and low cells and were analyze potential of anti-tumor<br />

agent resistant and tumorigensis. CD44high cells shows resistant<br />

to anti-tumor agent 5-FU compared with CD44low cells,<br />

and soft gel agar colony formation assay shows that CD44high<br />

cells make significantly higher colonies than CD44low cells.<br />

Subsequently we investigated what factor up-regulate CD44 in<br />

truncated HBx/HepG2 cells by Human Stem Cell PCR Array<br />

(Qiagen).This PCR-array results indicated that LTBP1 is significantly<br />

up-regulated in only Truncated-HBx/HepG2 compared<br />

with HBx/HepG2 and con/HepG2 cells. The expression level<br />

of CD44 were significantly down-regulated by siRNA knockdown<br />

of LTBP1. In sorted CD44high cells, LTBP1 were significantly<br />

up-regulated compared with CD44low cells. Conclusion:<br />

COOH-truncated HBx protein up-regulates CD44+ cancer stem<br />

like cells via LTBP1. This finding might contribute development<br />

of efficient strategies for eliminating CSCs<br />

Disclosures:<br />

Naoya Sakamoto - Grant/Research Support: Gilead Sciences, TRSS; Speaking<br />

and Teaching: Bristol Myers Squibb, Janssen Pharm, Chugai co ltd<br />

The following authors have nothing to disclose: Goki Suda, Seiji Tsunematsu,<br />

Masato Nakai, Jun Itoh, Mitsuteru Natsuizaka, Kenichi Morikawa, Koji Ogawa<br />

1947<br />

P300 promotes TGF-β stimulated nuclear translocation<br />

of SMADs and activation of hepatic stellate cells into<br />

liver metastasis promoting myofibroblasts<br />

Luyang Guo 2 , Xiaoyu Xiang 2 , Vijay Shah 1 , Ningling Kang 2 ;<br />

1 Mayo Clinic, Rochester, MN; 2 the Hormel Institute, Austin, MN<br />

Introduction: Liver metastasis is dependent on bidirectional interactions<br />

between cancer cells and the microenvironment of the<br />

liver. TGF-β, released from cancer cells and other cells within<br />

the liver, induces activation of hepatic stellate cells (HSCs) into<br />

myofibroblasts (MFs). In turn, MF/activated HSCs promote liver<br />

metastatic growth. TGF-β induces intracellular signaling events<br />

in HSCs, including activation of receptors, nuclear translocation<br />

of SMADs and transcription of TGF-β target genes, which<br />

present targets to inhibit HSC activation. However, mechanisms<br />

governing these signaling events remain poorly understood.<br />

Hypothesis: We tested if p300, an acetyltransferase<br />

previously known to promote gene transcription, modulates<br />

TGF-β1 induced activation of HSCs into tumor promoting MFs.<br />

Methods: Lentiviruses encoding different p300 shRNAs were<br />

used to knockdown p300 and a small molecule C646 was<br />

used to inhibit p300 acetyltransferase activity. A tumor/HSC<br />

coculture and coimplantation mouse model were used to test<br />

the role of p300 knockdown HSCs for tumor growth. Immunofluorescence<br />

(IF) was used to study subcellular localization of<br />

p300 and SMAD nuclear translocation. Western Blot analysis<br />

(WB) and IF for α-SMA were used to assess HSC activation,<br />

co-immunoprecipitation (coIP) was used to study protein-protein<br />

interactions and microarray analysis was used to study<br />

gene expression profiles of HSCs. Results: p300 knockdown<br />

HSCs were less effective on promoting tumor cell proliferation<br />

than control HSCs in a tumor/HSC coculture and were less<br />

effective on promoting tumor growth in mice in a subcutaneous<br />

tumor/HSC coimplantation model. In vitro, p300 knockdown<br />

inhibited TGF-β1 stimulated activation of HSCs into MFs as<br />

assessed by both WB and IF for α-SMA (p200 cells<br />

per group). IF revealed that p300 resided in both cytoplasm<br />

and nucleus of HSCs and that p300 knockdown or inhibition<br />

by C646 inhibited TGF-β1 stimulated nuclear translocation of<br />

SMADs. Co-IP demonstrated that TGF-β1 stimulation promoted<br />

a p300/Importin7/8/SMADs complex in the cytoplasm that<br />

promotes nuclear transport of SMADs. Furthermore, microarray<br />

analysis identified that p300 knockdown suppressed transcription<br />

of TGF-β1 inducible genes, such as CTGF, TNC, PDGFC,<br />

FGF2 and POSTN encoding either prometastatic niche factors<br />

or growth factors of tumor cells. Conclusions: p300 is required<br />

for TGF-β1 stimulated nuclear translocation of SMADs and transcription<br />

of TGF-β1 target genes encoding numerous tumor<br />

promoting paracrine factors. p300 of HSCs thus presents a<br />

therapeutic target to inhibit HSC activation required for the<br />

development and progression of liver metastasis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Luyang Guo, Xiaoyu Xiang, Vijay<br />

Shah, Ningling Kang<br />

1948<br />

Critical role of Hippo cascade in regulating AKT/Rasdriven<br />

liver cancer development in mice<br />

Shanshan Zhang 2,1 , Junyan Tao 1 , Xiaolei Li 1 , Diego Calvisi 3 ,<br />

Xin Chen 1 ; 1 Department of Bioengineering and Therapeutic Sciences,<br />

University of California San Francisco, San Francisco, CA;<br />

2 Department of Pathology and Pathophysiology, Eastern Hepatobiliary<br />

Surgery Hospital, Second Military Medical University, Shanghai,<br />

China; 3 Institut für Pathologie, Ernst-Moritz-Arndt-Universität,<br />

Greifswald, Germany<br />

Background: Primary liver cancer consists of both hepatocellular<br />

carcinoma (HCC) and intrahepatic cholangiocarcinoma<br />

(ICC). Hippo pathway is a critical regulator in liver cancer<br />

development. However, the precise role of the Hippo cascade<br />

along hepatocarcinogenesis has not been fully explored. Purpose:<br />

In this study, we investigated the functional significance<br />

of Yap and TAZ, the transcriptional activators downstream of<br />

Hippo tumor suppressor kinases, in a murine model of liver<br />

cancer. Methods: We employed a mixed HCC and ICC mouse<br />

model via hydrodynamic transfection of activated forms of AKT


1158A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and Ras oncogenes (AKT/Ras), based on the evidence that activation<br />

of AKT/mTOR and Ras/MAPK pathways often occurs in<br />

human HCC and ICC samples. In AKT/Ras mice, either the<br />

Hippo cascade was activated by co-injecting Lats2 or Yap/<br />

TAZ activity was blocked via transfection of dominant negative<br />

TEAD2 (dnTEAD2). Furthermore, Yap or TAZ was silenced via<br />

mir30 based shRNA. qRT-PCR, Western blotting, and immunohistochemistry<br />

(IHC) were applied to detect gene and protein<br />

expression patterns in tissues and cell lines. Results: Both Yap<br />

and TAZ showed nuclear staining in AKT/Ras liver tumor samples.<br />

Co-injecting Lats2 or dnTEAD2 with AKT/Ras strongly<br />

delayed liver tumor development. Similar results were obtained<br />

when the co-expression of AKT/Ras was paralleled by Yap or<br />

TAZ silencing in the mouse liver. Histological analysis showed<br />

decrease in liver tumor number and tumor size as well as the<br />

prevalence of adenomas over carcinomas in the liver of AKT/<br />

Ras mice. Importantly, activation of Hippo or loss of Yap/TAZ<br />

led to the complete elimination of ICC-like lesions in the liver.<br />

The phenotype was highly similar to that observed when AKT/<br />

Ras mice were treated with anti-Notch2 antibodies. Indeed,<br />

further analysis demonstrated the down-regulation of Notch2<br />

in these mouse liver tumor tissues. Silencing of Yap or TAZ in<br />

human HCC cell lines also led to the decreased expression of<br />

CCA-like markers such as Sox9 and CK19. Conclusion: Our<br />

data demonstrate that Hippo pathway negatively regulates<br />

liver tumor formation in mice via regulating Notch2. Both Yap<br />

and TAZ are required for hepatocarcinogenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shanshan Zhang, Junyan Tao,<br />

Xiaolei Li, Diego Calvisi, Xin Chen<br />

1949<br />

Hepatocyte-specific overexpression of B cell leukemia-3<br />

in mice suppresses chemically-induced hepatocarcinogenesis<br />

Nadine Gehrke 1 , Marcus A. Woerns 1 , Yvonne Alt 1 , Nadine Hoevelmeyer<br />

2 , Ari Waisman 2 , Peter R. Galle 1 , Jörn M. Schattenberg 1 ;<br />

1 I. Department of Medicine, University Medical Center of the<br />

Johannes Gutenberg University Mainz, Mainz, Germany; 2 Institute<br />

for Molecular Medicine Mainz, University Medical Center, Mainz,<br />

Germany<br />

Background: B cell leukemia-3 (Bcl-3), which was originally<br />

described as a proto-oncogene in B cell lymphomas, is highly<br />

expressed in the liver. It binds tightly to the NFkB subunits<br />

p50 and p52 homodimers in order to activate or inhibit transcription<br />

from NFkB-dependent promoters regulating especially<br />

inflammation and cell death. Bcl-3, when dysregulated, has be<br />

shown to be widely expressed in several cancer types including<br />

hepatocellular carcinoma (HCC), but a complete understanding<br />

of the function of Bcl-3 in cancer development is still<br />

lacking. Methods: To evaluate the role of hepatic Bcl-3 in hepatocarcinogenesis<br />

we employed the two-step diethylnitrosamine<br />

(DEN)/phenobarbital (PB)-liver cancer mouse model. 7-10<br />

day-old, male mice exhibiting an increased hepatocyte-specific<br />

expression of Bcl-3 (alfp-Cre:bcl-3, Bcl-3 hepar mice) and wild<br />

type (wt) littermates received a single intraperitoneal injection<br />

of 25mg DEN followed, 4 weeks later, by continuous treatment<br />

with 0.5g/l PB dissolved in drinking water. Blood and liver<br />

tissue were harvested at 40 weeks of age for further analysis.<br />

Results: Remarkably, Bcl-3 hepar mice exhibited less tumours compared<br />

to wt mice in response to DEN/PB. Also their relative<br />

liver weight was significantly lower, which could be attributed<br />

to the smaller tumour number and size. Histopathological<br />

tumour grading revealed a comparable presence of dysplastic<br />

foci and nodules in all H&E stained liver sections irrespective<br />

of the genotype. However, more HCC as well as a greater<br />

relative HCC area were detectable in the wt compared to the<br />

Bcl-3 hepar group. In addition, K i<br />

-67 labeling and appropriate<br />

qRT-PCR analyses (K i<br />

-67, p53, cMyc, mTOR) showed significantly<br />

diminished levels of cell proliferation in the HCC region<br />

of Bcl-3 hepar livers compared to wt livers. Moreover, DEN/PB<br />

treatment resulted in a markedly reduced liver injury in Bcl-<br />

3 hepar mice relative to the wt as demonstrated by lower serum<br />

ALT and LDH levels. In agreement, JNK and ERK activation<br />

were more pronounced in tumour and residual tumour tissues of<br />

wt mice, whereas a higher grade of phosphorylated p38 and<br />

NFkB p65 was determined in Bcl-3 hepar livers protecting against<br />

DEN/PB-induced cell death. Compared to the wt, the absolute<br />

number and activation of intrahepatic B cells, that exert a suppressive<br />

function on tumour growth, was reduced in Bcl-3 hepar<br />

mice as measured by FACS analysis and decreased mRNA<br />

expression of CD81 and BLNK. Conclusion: Overexpression of<br />

hepatic Bcl-3 impaired DEN/PB-induced hepatocarcinogenesis<br />

through regulation of cell death and inflammatory processes.<br />

Thus, Bcl-3 could be an important target in the development of<br />

novel therapies.<br />

Disclosures:<br />

Marcus A. Woerns - Advisory Committees or Review Panels: Bayer, Bayer<br />

Peter R. Galle - Advisory Committees or Review Panels: Bayer, BMS, Lilly, Daiichi,<br />

Jennerex; Consulting: Medimmune; Grant/Research Support: Roche, Lilly; Speaking<br />

and Teaching: Bayer, BMS<br />

The following authors have nothing to disclose: Nadine Gehrke, Yvonne Alt,<br />

Nadine Hoevelmeyer, Ari Waisman, Jörn M. Schattenberg<br />

1950<br />

Platelet-Derived Growth Factor D (PDGF-D)-Induced<br />

Secretion of Vascular Endothelial Growth Factor-C by<br />

Cancer-Associated Fibroblasts (CAF) Stimulates Lymphangiogenesis<br />

in Cholangiocarcinoma<br />

Massimiliano Cadamuro 2,1 , Marta Vismara 2 , Simone Brivio 2 ,<br />

Mario Strazzabosco 2,3 , Luca Fabris 1,3 ; 1 Dep. of Molecular Medicine,<br />

University of Padova, Padova, Italy; 2 Dep. of Surgery &<br />

Translational Medicine, University of Milan-Bicocca, Monza, Italy;<br />

3 Section of Digestive Diseases, Yale University, New Haven, CT<br />

Cholangiocarcinoma (CCA) is characterized by a strong<br />

invasiveness and a dismal prognosis. Early metastasisation to<br />

regional lymph nodes often precludes surgery, the only potentially<br />

curative option to treat CCA. Metastatic dissemination<br />

is facilitated by the presence of an abundant reactive stroma,<br />

mainly composed by cancer-associated fibroblasts (CAF), and<br />

by a rich lymphatic vasculature. PDGF-D secreted by CCA cells<br />

plays a major role in CAF recruitment. Mechanisms of lymphangiogenesis<br />

in CCA are unknown. Therefore, we investigated<br />

the possible role of PDGF-D and CAF in this process. Methods/Results.<br />

Human CCA specimens were immunostained with<br />

D2-40 (lymphatic endothelial cells, LEC), CD34 (vascular blood<br />

endothelial cells, VEC), αSMA (CAF) alone or in combination<br />

with antibodies against VEGF-C or its receptor VEGFR-3 (n=6).<br />

This analysis showed that in CCA, lymphatic endothelial cells<br />

(LEC), positive for VEGFR-3, rather than VEC, were closely adjacent<br />

to VEGF-C expressing CAF. Secretion of VEGF-C, VEGF-D,<br />

Ang-1, Ang-2 was then evaluated in human primary cultured<br />

fibroblasts challenged with PDGF-D (n=3). Upon PDGF-D stimulation,<br />

fibroblasts secreted increased levels of the lymphangiogenic<br />

growth factor VEGF-C, whereas VEGF-D and Ang-2<br />

were not expressed. In fibroblasts, PDGF-D-stimulated VEGF-C<br />

expression was significantly inhibited by blocking PDGFRβ<br />

(the cognate receptor of PDGF-D) with imatinib mesylate, or<br />

by inhibiting JNK and ERK/NF-kB signaling with SP600125,<br />

U0126 and BAY11-7082, respectively. LEC recruitment (tran-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1159A<br />

swell migration) and vascular assembly by LEC were measured<br />

after exposure to conditioned medium (CM) of fibroblasts challenged<br />

with PDGF-D. CM stimulated LEC recruitment and their<br />

tubular assembly (lumen diameter, tube elongation and anastomization),<br />

to an extent comparable to that induced by VEGF-C<br />

alone. These effects of CM on LEC were inhibited by antagonizing<br />

PDGFRβ in fibroblasts and by inhibiting VEGFR-3 in LEC<br />

(n=6). In conclusion, this study provides a working model to<br />

understand the mechanisms by which the interaction between<br />

cancer cells and CAF promotes lymphangiogenesis in CCA.<br />

Our data show that PDGF-D secreted by CCA cells not only<br />

recruits CAF in the tumor reactive stroma, but also stimulates<br />

CAF to release VEGF-C acting through the JNK and ERK/NF-kB<br />

signaling. In turn, VEGF-C promotes tumor lymphangiogenesis<br />

by stimulating VEGFR-3 expressed on LEC. This interplay may<br />

be at the basis of the high rate of lymphatic metastasization<br />

in CCA. Furthermore, the ability to interfere with these mechanisms<br />

may prevent lymphatic metastatization of CCA.<br />

Disclosures:<br />

The following authors have nothing to disclose: Massimiliano Cadamuro, Marta<br />

Vismara, Simone Brivio, Mario Strazzabosco, Luca Fabris<br />

Notch1 up-regulation cells expressed both C-myc and OV6 (A,<br />

white arrow). Representative tumors in liver and lung (B, black<br />

arrow) and their H&E staining. VCAM-1 was activated in Notch1<br />

transfected cells by iTRAQ (C, red frame), immunofluoresecence<br />

(D, white arrow) and western blot analyses (E). VCAM-1 was<br />

overexpressed (F, black arrow) and TAMs were abnormally<br />

recruited in the pulmonary metastasis (G, black arrow).<br />

1951<br />

Notch-1 Up-regulated Rat Oval Cells Generated Poorly<br />

Differentiated Liver Cancer and Developed Lung Metastases<br />

in an Orthotopic Rat Model<br />

Wen-rui Wu, Xiang-de Shi, Rui Zhang, Lei-bo Xu, Mansheng Zhu,<br />

Xian-huan Yu, Hong Zeng, Chao Liu; Sun Yat-sen Memorial Hospital,<br />

Sun Yat-sen University, Guangzhou, China<br />

Background and aims: Oval cells may give rise to liver cancer.<br />

Notch signaling has been reported to play crucial role in hepatocellular<br />

carcinoma. The aim of this study was to investigate<br />

the biological role of Notch1 in rat oval cell line WB-F344.<br />

Methods: Notch1 was stably transfected into WB-F344 cells<br />

by lentivirus. Their proliferation, colony formation, cell cycle<br />

progression and invasion assays in vitro, as well as tumor<br />

formation assay in an orthotopic (liver) rat model were performed.<br />

Results: Gain-of-function analysis showed that Notch1<br />

up-regulation promoted proliferation, colony formation, G1/S<br />

cell cycle transition and invasion of WB-F344 cells in vitro.<br />

Notch1 transfected cells expressed oncogene C-myc and oval<br />

cell marker OV6 simultaneously. These cells showed a pile up<br />

appearance in culture and could generate poorly differentiated<br />

liver cancer orthotopically in isogenic rats (6/6, 100%).<br />

Interestingly, all tumor bearing rats developed lung metastases<br />

(6/6, 100%). In addition, iTRAQ analysis indicated that<br />

vascular cell adhesion molecule 1 (VCAM-1) was activated<br />

in Notch1 transfected cells in vitro, which was confirmed by<br />

immunofluorescence and western blot analysis. Moreover, we<br />

discovered that VCAM-1 was highly expressed in pulmonary<br />

metastasis tumors and tumor-associated macrophages (TAMs)<br />

were abnormally recruited in these tumors by immunohistochemistry.<br />

Conclusions: Notch1 is a crucial regulator of proliferation<br />

and malignant transformation of oval cells. Notch1<br />

may facilitate pulmonary metastasis of primary liver cancer via<br />

VCAM-1 activation and TAMs recruitment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Wen-rui Wu, Xiang-de Shi, Rui<br />

Zhang, Lei-bo Xu, Mansheng Zhu, Xian-huan Yu, Hong Zeng, Chao Liu<br />

1952<br />

Deregulated Interplay Between Methionine Adenosyltransferase<br />

α1, c-Myc and Maf Proteins During Chronic<br />

Cholestasis Facilitate Development of Cholangiocarcinoma<br />

Heping Yang 1,2 , Ting Liu 3 , Jiaohong Wang 1,2 , Tony Li 1,2 , Jose M.<br />

Mato 4 , Shelly C. Lu 1,2 ; 1 Medicine, Cedars-Sinai Medical Center,<br />

Los Angeles, CA; 2 USC Research Center for Liver Diseases, Los<br />

Angeles, CA; 3 Department of Gastroenterology, Xiangya Hospital<br />

Central South University, Changsha, China; 4 Centro de Investigación<br />

Biomédica en Red de Enfermedades Hepáticas y Digestivas<br />

(Ciberehd), CIC bioGUNE, Derio, Spain<br />

Purpose of Study: Cholangiocarcinoma (CCA) develops frequently<br />

in the setting of chronic inflammation and cholestasis.<br />

We reported the importance of c-Myc induction in the<br />

development of cholestatic liver injury and CCA in mice. We<br />

also showed induction of Maf proteins (MafG and c-Maf) contributed<br />

significantly to cholestatic liver injury. The aim of our<br />

current work was to determine whether there is any interplay<br />

between c-Myc and Maf proteins but in the process of our<br />

investigation we uncovered interesting reciprocal regulations<br />

that likely contribute to development of CCA, particularly in<br />

the setting of chronic cholestasis. Methods: Our study used<br />

bile duct ligation (BDL) and lithocholic acid (LCA) treatment<br />

as chronic cholestasis models, diethylnitrosamine followed by<br />

left and median BDL as a murine CCA model, human CCA<br />

cell lines KMCH and Huh-28, and human CCA specimens.<br />

We used immunoprecipitation followed by mass spectrometry<br />

and recombinant proteins to study protein-protein interactions,<br />

chromatin immunoprecipitation (ChIP) and sequential ChIP to<br />

examine co-occupancy of promoter region, overexpression and<br />

siRNA to vary protein expression. Results: We identified novel<br />

interacting proteins with c-Myc, specifically methionine adenosyltransferase<br />

α1 (MATα1, encoded by MAT1A), MafG and<br />

c-Maf. MAT1A expression fell at the mRNA and protein levels<br />

in whole liver and bile duct epithelial cells during BDL and LCA<br />

treatment, and in murine and human CCA samples. The opposite<br />

occurred with c-Myc, MafG and c-Maf expression. MATα1


1160A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

interacts mainly with Mnt in normal liver but this switches to<br />

c-Maf, MafG and c-Myc in cholestatic livers and CCA. Furthermore,<br />

MATα1 interacts directly with c-Myc, MafG and c-Maf.<br />

Interestingly promoter regions of MAT1A, c-Maf, MafG and<br />

c-Myc have E-box sequences that are bound by MATα1 and<br />

Mnt at baseline that switched to c-Myc, c-Maf and MafG after<br />

BDL and LCA treatment. MATα1 binds to the E-box as a complex<br />

with c-Myc and Max, but not by itself or with c-Myc. While<br />

E-box positively regulates c-Myc, MafG and c-Maf, it negatively<br />

regulates MAT1A. Furthermore, MATα1 binding to the E-box<br />

represses E-box and lowers the promoter activity of c-Myc,<br />

MafG and c-Maf, whereas c-Myc, MafG and c-Maf binding<br />

enhance E-box-driven promoter activity. This results in reciprocal<br />

regulation between MATα1 and c-Myc and Maf proteins.<br />

In CCA cell lines knockdown of MAT1A or overexpression of<br />

MafG or c-Maf enhanced tumorigenesis. Conclusions: we have<br />

uncovered a novel interplay between MATα1, c-Myc and Maf<br />

proteins and their deregulation during chronic cholestasis sets<br />

the stage for CCA oncogenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Heping Yang, Ting Liu, Jiaohong<br />

Wang, Tony Li, Jose M. Mato, Shelly C. Lu<br />

suppressed by regorafenib in the sorafenib-resistant clones,<br />

suggesting that sorafenib-resistance might be related to the<br />

acquisition of resistance against ERK signaling pathway inhibition<br />

mediated by sorafenib. Promisingly, regorafenib treatment<br />

suppressed the subcutaneous tumor growth of sorafenib-resistant<br />

Huh7 and Huh1 clones with significantly prolonged overall<br />

survival in vivo compared with sorafenib treatment. Conclusions:Regorafenib<br />

inhibited the growth of sorafenib-resistant<br />

HCC cells, potentially through suppression of the ERK signaling<br />

pathway. The efficacy of regorafenib on the sorafenib-resistant<br />

clones suggests its potential utility in HCC patients with resistance<br />

to sorafenib.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Mariko Yoshida - Grant/Research Support: Bayer<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Tomomi Hashiba, Taro Yamashita,<br />

Kouki Nio, Takehiro Hayashi, Yoshimoto Nomura, Tomoyuki Hayashi, Naoki<br />

Oishi, Masao Honda<br />

1953<br />

Regorafenib inhibits ERK signaling and suppresses the<br />

growth of sorafenib-resistant cells in human hepatocellular<br />

carcinoma<br />

Tomomi Hashiba, Taro Yamashita, Hikari Okada, Kouki Nio,<br />

Takehiro Hayashi, Yoshimoto Nomura, Mariko Yoshida, Tomoyuki<br />

Hayashi, Naoki Oishi, Masao Honda, Shuichi Kaneko; Gastroenterology,<br />

Kanazawa University Graduate School of Medical<br />

Science, Kanazawa, Japan<br />

Background:Regorafenib is a multi-kinase inhibitor targeting<br />

VEGFRs, PDGFRs, and RAF. A recent phase II study suggested<br />

acceptable tolerability and anti-tumor activity of regorafenib in<br />

patients with hepatocellular carcinoma (HCC) that progressed<br />

after initial sorafenib treatment. A confirmatory phase III study<br />

(NCT01774344) is ongoing. However, it remains unclear<br />

how regorafenib inhibits the growth signaling pathways of<br />

sorafenib-resistant HCC. In this study, we evaluated the effects<br />

of regorafenib and sorafenib on HCC growth and signaling<br />

pathways in HCC cell lines and primary HCC cells. Methods:Two<br />

HCC cell lines (Huh1 and Huh7) and three primary<br />

HCC cells obtained from surgically resected specimens were<br />

cultured routinely in DMEM supplemented with 10% fetal bovine<br />

serum and sorafenib tosylate (5-10 μM) for 3 months to generate<br />

sorafenib-resistant clones. Whole exome sequence analysis<br />

was performed to detect mutations in the sorafenib-resistant<br />

clones and their parental cells. Sensitivity against sorafenib<br />

or regorafenib was evaluated using a cell proliferation assay<br />

kit. Gene and protein expression was evaluated by quantitative<br />

RT-PCR and western blotting. A subcutaneous xenotransplantation<br />

model was used to evaluate the efficacy of orally<br />

administered sorafenib (30 mgkg -1 day -1 ) or regorafenib (20<br />

mgkg -1 day -1 ) on sorafenib-resistant clones in vivo. Results:All<br />

established sorafenib-resistant clones showed higher sensitivity<br />

to regorafenib compared with sorafenib in vitro, whereas<br />

parental cells showed similar sensitivity to both sorafenib and<br />

regorafenib. Sorafenib-resistant clones derived from cell lines<br />

showed abundant expression of the cancer stem cell marker<br />

CD44 compared with the parental cells. Although sorafenib-resistant<br />

clones showed frequent missense/nonsense mutations<br />

compared with the parental clones, no common mutations were<br />

detected among all sorafenib-resistant clones. Interestingly,<br />

ERK phosphorylation was not affected by sorafenib, but was<br />

1954<br />

The Dead Box Protein P68 Can Supress The Carcinogenesis<br />

And Α-Fetoprotein (AFP) Gene Expression via Binding<br />

The New Silencer Region Of AFP Gene<br />

Shunsuke Nojiri, Kei Fujiwara, Noboru Shinkai, Etsuko Iio, Takashi<br />

Joh; Department of Gastroenterology and Metabolism, Nagoya<br />

City University Graduate School of Medical Sciences, Nagoya,<br />

Japan<br />

The activity of the α-fetoprotein (AFP)<br />

gene decreases rapidly after birth but is reactivated in hepatocellular<br />

carcinoma (HCC) and also during proliferation of<br />

hepatocyte. In clinical therapy of hepatocellular carcinoma<br />

AFP reduction can be the good marker after anticancer therapies.<br />

We predicted a new silencer lesion that was located<br />

upstream from previous reported areas. The aim of this study<br />

is to identify the precise lesion of the silencer and demonstrate<br />

a new protein which suppresses AFP gene expression via its<br />

new suppresser lesion. A basic vector was constructed<br />

by inserting the 4.9-kb AFP5’-flanking sequence into<br />

PGL2 basic vector. The vectors deleted several sections were<br />

also constructed. The Luciferase activities were expressed in<br />

relation to the activity of the cells that were transfected with<br />

the basic vector and we found the 179 base pair new silencer<br />

lesion. After immunoprecipitation by biotinized PCR products<br />

including the silencer arrangement and the nuclear extracts, the<br />

protein was identified from using an LC-MS/MS assay. Chromatin<br />

immunoprecipitation (ChIP) was performed to confirm the<br />

protein binding to the new silencer. The candidate protein was<br />

suppressed or overexpressed using siRNA duplexes or overexpression<br />

vector in HepG2 cells and AFP expression measured<br />

by Luciferase activities and real time PCR and cell proliferation<br />

were measured by cell proliferation assay. The results<br />

of the LC-MS/MS assay demonstrated the presence of DEAD<br />

box protein p68 (p68). P68 protein could binding to the new<br />

silencer session was confirmed by ChIP. After using RNAi to<br />

p68 the Luciferase activities increased 3.5±1.1 times (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1161A<br />

same results. A new 179 base pair silencer position<br />

was identified which was located at 1153 bases upstream<br />

from which was reported previously. The p68 protein could<br />

suppress the AFP gene expression via binding to this lesion<br />

directly and suppress the cell proliferation. In hepatocellular<br />

carcinoma cells p68 may suppress not only AFP production<br />

but also cell proliferation. In the clinical therapies reduction<br />

of AFP can become the good prediction of carcinogenesis or<br />

reduction the mass of carcinoma. This report revealed the one<br />

aspect of the mechanism of these relationships.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shunsuke Nojiri, Kei Fujiwara,<br />

Noboru Shinkai, Etsuko Iio, Takashi Joh<br />

1955<br />

WITHDRAWN<br />

1956<br />

Macrophage colony-stimulating factor receptor antagonist<br />

inhibits progression of hepatocellular carcinoma in<br />

vivo.<br />

Hiroshi Kono, Yoshihiro Akazawa, Shinji Furuya, Hideki Fujii; First<br />

Department of Surgry, University of Yamanashi, Chuo, Japan<br />

Aim; We previously reported that macrophage colony-stimulating<br />

factor (M-CSF) is involved in hepatocarcinogenesis by<br />

inducing an angiogenic factor via the hepatic Mφs. Therefore,<br />

M-CSF could be targets of therapy in hepatocellular carcinoma<br />

(HCC). The spurpose of this study was to investigate effects of<br />

M-CSF receptor antagonist on HCC. Materials and Method;<br />

1; isolated mouse hepatic macrophages (Mφs) or monocytes<br />

(Mos) were cultured with media added the different dose of<br />

M-CSF. Production of vascular endothelial growth factor (VEGF)<br />

was assessed. Furthermore, isolated vascular endothelial cells<br />

(VEC) were co-cultured with or without hepatic Mφs in presense<br />

with M-CSF (100ng/ml) for 5 days and cell proliferations were<br />

assessed in vitro. 2; C57/BL6 mice were treated with diethyl<br />

nitrosamine (DEN) intraperitoneally. For treatment group,<br />

M-CSF receptor antagonist (GW2580) was treated every day.<br />

Incidence of tumors was assessed 38 weeks after treatment.<br />

Furthermore, angiogenesis and distribution of CD163-positive<br />

M2 Mφs were assessed. 3; Mouse HCC cells (MH134) were<br />

implanted to same strain C3H mice by subcutaneous injection<br />

(1 ×10 5 /animal). Tumor progression was assessed after<br />

3 weeks. 4; In the nude mouse, human HCC cells (HuH7)<br />

were implanted by intra-splenic injection (1 × 10 6 /animal).<br />

Tumor progression in the spleen was assessed after 3 weeks.<br />

5; MH134 or HuH7 (1 × 10 4 /well) was cultured with media<br />

containing M-CSF (100ng/ml) in the presence or absence of<br />

GW2580 (1ng/ml) for 7 days in vitro, and cell proliferation<br />

was assessed. Result; 1; Production of VEGF increased both in<br />

hepatic Mφs and MOs incubated with M-CSF in dose dependent<br />

and time dependent manner. Furthermore, the production<br />

was significantly greater in hepatic Mφs compared with that in<br />

Mos. Importantly, proliferation of VEC was greatest in cells cultured<br />

with hepatic Mφs in media with M-CSF. 2; Hepatic tumors<br />

diagnosed as hepatocellular adenoma or HCC were observed<br />

in animals treated with DEN. In contrast, tumor incidence was<br />

significantly reduced in DEN-treated animals with GW2580.<br />

Enhanced angiogenesis M2Mf population observed in the<br />

DEN-treated animals was significantly blunted by treatment of<br />

GW2580. 3 and 4; Growth of implanted both of HCC was<br />

significantly inhibited by GW2580 in vivo. 5; Proliferations<br />

both of HCC were not significant difference between in cells<br />

cultured with M-CSF and cells cultured without M-CSF. Furthermore,<br />

GW2580 did not inhibit proliferation of these cells in<br />

vitro. Conclusions; M-CSF receptor antagonist markedly inhibited<br />

tumor initiation and progression. Thus, M-CSF and/or its<br />

receptor could be a new therapeutic target for HCC. .<br />

Disclosures:<br />

The following authors have nothing to disclose: Hiroshi Kono, Yoshihiro Akazawa,<br />

Shinji Furuya, Hideki Fujii<br />

1957<br />

Promotion of Hepatocarcinogenesis by the Endocannabinoid<br />

Anandamide and its Receptor CB1<br />

Ki Tae Suk 1 , Ingmar Mederacke 1 , Geum Youn Gwak 2 , Robert F.<br />

Schwabe 1 ; 1 Columbia University, New York, NY; 2 Sungkyunkwan<br />

University, Seoul, Korea (the Republic of)<br />

BACKGROUND: The endocannabinoid system (ECS) exerts key<br />

roles in the development of liver fibrosis and fatty liver, two<br />

diseases that promote hepatocarcinogenesis. Although cannabinoids<br />

exert potent anti-tumor effects in vitro, the role of the<br />

ECS in carcinogenesis in vivo remains elusive. AIM: To understand<br />

the contribution of the ECS to hepatocarcinogenesis.<br />

METHODS: Expression endocannabinoids (EC), EC-degrading<br />

enzymes and EC receptors, was determined by LC-MS/MS,<br />

qPCR and immunohistochemistry in mouse and patient samples.<br />

The contribution of ECS to diethylnitrosamine (DEN)-induced<br />

hepatocellular carcinoma (HCC) was determined in<br />

mice deficient in fatty acid amide hydrolase (FAAH), the main<br />

anandamide (AEA)-degrading enzyme, and in mice deficient<br />

for cannabinoid receptor 1 (CB1), cannabinoid receptor 2<br />

(CB2), or transient receptor potential cation channel subfamily<br />

V member 1 (TRPV1). RESULTS: Murine and human HCCs displayed<br />

activation of the ECS with elevated expression of CB1<br />

and CB2 mRNA and protein, unaltered TRPV1 expression and<br />

only moderate alterations of EC levels. Surprisingly, FAAH-deficient<br />

mice, which have about 10-fold increased hepatic levels<br />

of CB1 agonist AEA, displayed a 129% (p


1162A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1958<br />

Chronic alcohol accelerates early hepatobiliary cancer<br />

by increasing inflammation, stemness, EMT and miR-<br />

122 mediated HIF-1α activation<br />

Aditya Ambade, Abhishek Satishchandran, Gyongyi Szabo; Medicine,<br />

UMass Medical School, Worcester, MA<br />

Background / Aims: Chronic alcohol use is a major risk factor<br />

in the development of hepatocellular carcinoma [HCC].<br />

Inflammation contributes to alcoholic liver disease, particularly<br />

in alcoholic hepatitis, and promotes HCC development. Other<br />

molecular events leading to HCC include proliferation of tumor<br />

stem cells, epithelial-mesenchymal transition (EMT) and altered<br />

microRNA expression in the liver. We hypothesized that the<br />

pro-inflammatory microenvironment in alcoholic hepatitis accelerates<br />

liver tumor development after pre-exposure to a chemical<br />

carcinogen. Methods: Four week old male C57BL/6 mice<br />

were injected intraperitoneally with N, N diethyl nitrosamine<br />

(DEN), 75 mg/kg/week for 3 consecutive weeks followed by<br />

100 mg/kg/week for 3 additional weeks. At age 8 weeks,<br />

mice were divided into alcohol (4% Lieber-DeCarli alcohol<br />

diet) or calorie matched control diet (pair-fed group). After 6<br />

weeks of alcohol feeding mice were sacrificed; blood, liver<br />

tissues were collected for analysis of cytokines, tumor markers<br />

and miR-122. Cyclin D1, p53, vimentin and sonic hedgehog<br />

proteins were evaluated by western blots. HIF-1a activity was<br />

assayed by EMSA. Results: Alcohol-fed DEN-injected mice<br />

had greater liver damage (ALT) and significantly increased<br />

liver-to-body weight ratio compared to pair-fed DEN-injected<br />

mice. Alcohol feeding resulted in steatohepatitis indicated by<br />

increased pro-inflammatory cytokines TNFa, IL-6, MCP-1, IL-17<br />

and fibrotic genes a-SMA, procollagen1a1 and TGFb, that<br />

were significantly higher in alcohol plus DEN mice compared<br />

to other groups. MRI and liver histology of alcohol plus DEN<br />

mice revealed hepatobiliary cysts, early hepatic neoplasia and<br />

increased AFP. No tumors were found in other groups. Proliferation<br />

makers (PCNA, cyclin D1, p53), EMT markers (vimentin,<br />

sonic hedgehog) and cancer stem cell markers (CD133<br />

and nanog) were significantly upregulated in livers of alcohol-fed<br />

DEN-injected mice compared to pair-fed DEN-injected<br />

and alcohol-fed saline-injected mice. Immunostaining for AFP,<br />

PCNA, cytokeratin7 and 19 showed increased number of<br />

stained cells in alcohol-fed DEN-injected mice compared to<br />

all other groups indicating proliferation of hepatobiliary progenitors.<br />

In livers with tumors from alcohol plus DEN mice, we<br />

found significant reduction of miR-122 with upregulation of<br />

miR-122 target gene, HIF-1α. Increased HIF-1a DNA binding<br />

activity and upregulation of its target, VEGFR1, was found in<br />

livers of alcohol plus DEN mice compared to all other groups.<br />

Conclusion: Our findings establish a new model of early HCC<br />

and demonstrate that alcoholic hepatitis accelerates hepatobiliary<br />

tumor development with molecular features of HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Aditya Ambade, Abhishek Satishchandran,<br />

Gyongyi Szabo<br />

1959<br />

Variations of the host genome and interaction of hepatitis<br />

B viral X protein associated with hepatocarcinogenesis<br />

Hiroko Nagata 1 , Yasuhiro Itsui 1,2 , Fukiko Kawai-Kitahata 1 , Shun<br />

Kaneko 1 , Miyako Murakawa 1,2 , Sayuri Nitta 1 , Mina Nakagawa 1 ,<br />

Seishin Azuma 1 , Sei Kakinuma 1 , Yasuhiro Asahina 1 ; 1 Department<br />

of Gastroenterology and Hepatology, Tokyo Medical and Dental<br />

University, Tokyo, Japan; 2 Department of Medical Education<br />

Research and Development, Tokyo Medical and Dental University,<br />

Tokyo, Japan<br />

Background and Aims: Association between variation of the<br />

host genome and viral proteins responsible for hepatocarcinogenesis<br />

is unclear. To clarify this, we comprehensively analyzed<br />

variations of the host genome in patients with HCC using<br />

a next-generation sequencer, and investigated the association<br />

between identified host genome variation and HBx protein.<br />

Methods: [1] We performed deep sequenceing in 104 pairs of<br />

HCC and non-tumor samples targeting 54 cancer-related genes<br />

containing 2820 mutational hotspots. [2] To assess the effect of<br />

TERT promoter variations (C228T, C250T), which were identified<br />

as most frequent mutations in our samples, on its transcriptional<br />

activity, reporter assays using TERT promoter-luciferase<br />

plasmids with or without these mutations were performed.<br />

[3] To assess the association between TERT promoter activity<br />

and HBx protein, reporter assays were performed 48 hours<br />

after co-transfection of plasmid expressing HBx protein and<br />

TERT promoter-luciferase plasmids in HepG2 cells and 293T<br />

cells. [4] Moreover, effects of transcriptional factor Sp1 and<br />

c-myc on TERT promoter activity were analyzed by additional<br />

co-transfection using plasmid expressing these transcriptional<br />

factors. Results: [1] We detected somatic mutations in 9 genes<br />

out of 93 samples: TERT, TP53, CTNNB1, PTEN, CDKN2A,<br />

HRAS, PIK3CA, STK11, GNAS and NFE2L2. Among them,<br />

TERT promoter mutations were most frequently found (65% of<br />

the patients) in overall. However, these mutations were significantly<br />

less frequent in HBV-related HCC, while integration of<br />

HBx gene into TERT region was frequently found, which was<br />

mutually exclusive to TERT promoter mutations. [2] The TERT<br />

mutations (C228T, C250T) significantly upregulated the promoter<br />

activity in comparison with wild-type prompter in HepG2<br />

and 293T cell (181±2.2%, 255±5.2%). [3] Expression of HBx<br />

protein resulted in further activation of TERT promoter activity<br />

in both of TERT-mutant and wild type (714±3.1%, 628±2.1%),<br />

and effect of HBx protein on TERT promoter activation was<br />

more remarkable in the TERT-wild promoter than the TERT-mutant<br />

one (560%, 435%). [4] The promoter activity was elevated<br />

by Sp1 or c-myc, which are TERT promoter-binding factors.<br />

Conclusions: TERT promoter mutations were most frequently<br />

detected in non-HBV related HCC, but rare in HBV-related<br />

HCC, in which integration of HBx into TERT region were frequently<br />

and exclusively found. Upregulation of TERT promoter<br />

activity by either TERT promoter mutations or interaction of HBx<br />

protein with TERT promoter via Sp1 and c-myc may associate<br />

with hepatocarcinogenesis or tumor progression.<br />

Disclosures:<br />

Sei Kakinuma - Grant/Research Support: The Japanese Society of Gastroenterology,<br />

MSD Co., Ltd.<br />

The following authors have nothing to disclose: Hiroko Nagata, Yasuhiro Itsui,<br />

Fukiko Kawai-Kitahata, Shun Kaneko, Miyako Murakawa, Sayuri Nitta, Mina<br />

Nakagawa, Seishin Azuma, Yasuhiro Asahina


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1163A<br />

1960<br />

Survival of hepatocellular carcinoma patients according<br />

to recommended treatment by the Barcelona Clinic Liver<br />

Cancer or Hong Kong Liver Cancer staging system in<br />

hepatitis B endemic area<br />

Ju-Il Yang, Dong Hyun Sinn, jung hee kim, Geum-Youn Gwak,<br />

Yong-Han Paik, Moon Seok Choi, Joon Hyeok Lee, Kwang Cheol<br />

Koh, Seung Woon Paik, Byung Chul Yoo; Gastroenterology, Samsung<br />

Medical Center, Seoul, Korea (the Republic of)<br />

Background: Cancer staging system aims to help to predict<br />

patient prognosis as well as select treatment modality. Several<br />

staging system has been proposed for hepatocellular carcinoma<br />

(HCC), and of those, only Barcelona Clinic Liver Cancer<br />

(BCLC) and Hong Kong Liver Cancer (HKLC) staging system<br />

has recommended treatment modality. We analyzed long-term<br />

outcome of patients according to the recommended treatment<br />

modality in both staging system. Methods: A total of 3,515 treatment-naïve,<br />

newly diagnosed HCC patients [age: 62.8 ± 10.5,<br />

male: 2,792 (79.4%), hepatitis B virus: 2,709 (77.1%)] in a<br />

single center were analyzed. Results: The median survival was<br />

5.4 years, with 5-years cumulative survival rate of 50%. BCLC<br />

and HKLC staging well-stratified patients prognosis (5-years<br />

survival rate: 79.1%, 62.9%, 40.3%, 21.3%, and 27.0% for<br />

BCLC 0, A, B, C and D, respectively; 72.3%, 54.9%, 50.6%,<br />

21.3%, 10.2%, 16.7%, 7.2%, 47.1% and 11.3% for stage<br />

I, IIa, IIb, IIIa, IIIb, IVa, IVb, Va and Vb, respectively). C-index<br />

of BCLC and HKLC staging system was 0.708 and 0.732.<br />

Overall, 49.5% patients received BCLC-recommended treatment,<br />

and patients survival was better when patient received<br />

recommended treatment in stage O and A, but survival was<br />

worse in stage B, C and D. For HKLC staging system, 55.6% of<br />

patients received HKLC-recommended treatment, and survival<br />

was better when patient received recommended treatment in<br />

stage I, IIa, IIb, IIIa, Va but not in stage IIIb, IVa, IVb, and Vb.<br />

Conclusion: Both BCLC and HKLC staging system could stratify<br />

patient prognosis, however, could not direct therapy for large<br />

proportion of patients, and in some stage, recommended therapy<br />

was associated with worse prognosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ju-Il Yang, Dong Hyun Sinn, jung<br />

hee kim, Geum-Youn Gwak, Yong-Han Paik, Moon Seok Choi, Joon Hyeok Lee,<br />

Kwang Cheol Koh, Seung Woon Paik, Byung Chul Yoo<br />

1961<br />

Inactivation of Aldose Reductase Prevents NAFLD,<br />

NASH, and Hepatocarcinogenesis in Transaldolase Deficiency<br />

Zachary Oaks 1 , Robert Hanczko 1 , Miguel Beckford 1 , Sookja K.<br />

Chung 3 , Steve Landas 1 , John Asara 2 , Andras Perl 1 ; 1 Medicine,<br />

SUNY Upstate, Syracuse, NY; 2 Medicine, Beth Israel Deaconess<br />

Medical Center, Boston, MA; 3 Anatomy, The University of Hong<br />

Kong, Hong Kong, China<br />

Background and Aims: Transaldolase (TAL) is a rate limiting<br />

enzyme of the pentose phosphate pathway. TAL deficiency<br />

elicits NAFLD, NASH, and hepatocellular carcinoma (HCC)<br />

in humans and mice. The pathogenesis is driven by oxidative<br />

stress and characterized by the accumulation of sugar phosphates,<br />

depletion of NADPH and glutathione (J. Clin. Invest.<br />

2009;119:1546-57). TAL-deficient mice exhibit the overexpression<br />

of the NADPH dependent enzyme aldose reductase<br />

(AR). Here, we investigated if the role of AR in the depletion of<br />

NADPH and the resultant pathologies. Methods: Metabolites<br />

were extracted from TAL, AR, and TAL/AR double-knockout<br />

livers that were sacrificed along with C57BL/6 controls, using<br />

5 age-matched mice of each strain. 258 metabolites were<br />

detected by LC-MS/MS with selective reaction monitoring.<br />

Relative metabolite concentrations and pathway fluxes were<br />

evaluated using Metaboanalyst. In addition, the development<br />

of NAFLD, NASH, and HCC were monitored up to 78 weeks<br />

of age. Results: 0/21 TAL/AR double-knockouts, 45/97 TAL<br />

knockouts, and 3/106 controls developed HCC by 78 weeks<br />

of age. Deletion of AR in TAL deficiency also rescued fatty liver;<br />

no hepatosteatosis was detected in mice lacking both enzymes.<br />

AR knockout livers had no significant pathology relative to<br />

control animals. The liver of double-knockouts were similar<br />

to wild-type from controls (p=1.0) and they had significantly<br />

reduced HCC compared to TAL knockouts (p


1164A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

This study suggests that TGFβ1 is involved in CCA tumor progression<br />

and may mediate this process through miR-34a downstream<br />

signaling pathways, raising the possibility that TGFβ1<br />

signaling may offer potential therapeutic targets to inhibit CCA<br />

development and growth.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chiung-Kuei Huang, Arihiro<br />

Aihara, Yoshifumi Iwagami, Tunan Yu, Rolf I. Carlson, Hironori Koga, Miran<br />

Kim, Jack R. Wands<br />

1963<br />

Hepatic Ruvbl1 is key regulator of glucose homeostasis<br />

and promotes hepatocellular carcinoma progression in<br />

vivo.<br />

Tommaso Mello 1,2 , Francesca Zanieri 1 , Elisabetta Ceni 1 , Mirko<br />

Tarocchi 1 , Giada Marroncini 1 , Simone Polvani 1 , Sara Tempesti 1 ,<br />

Oxana Bereshchenko 3 , Stefano Milani 1,2 , Andrea Galli 1,2 ; 1 University<br />

of Florence, Florence, Italy; 2 Careggi University Hospital,<br />

Florence, Italy; 3 University of Perugia, Perugia, Italy<br />

The AAA+ ATPase Ruvbl1 is overexpressed in several human<br />

cancers, including hepatocellular carcinoma (HCC), in which<br />

high Ruvbl1 expression correlates with a poor prognosis. A<br />

growing body of data from in vitro models supports the concept<br />

that deregulation of Ruvbl1 expression occurs in cancer to promote<br />

its growth and progression, making this protein an attractive<br />

target for anti-cancer therapies. However, whether Ruvbl1<br />

actively drives HCC onset and progression remains speculative.<br />

To challenge this question we realized an hepatocyte-conditional<br />

Ruvbl1 +/- mouse and evaluated the number and size<br />

of tumor foci induced by DEN. Ruvbl1 +/- mice were obtained<br />

by crossing Ruvbl1-floxed with Albumin-Cre deleter mice. The<br />

male offspring were subjected to a single i.p. injection of DEN<br />

(5mg/kg) to induce liver cancer, and were monitored up to one<br />

year after cancer induction. Contrary to our expectations, we<br />

found that despite an initial delay in the appearance of liver<br />

cancer 6 month after DEN injection, conditional Ruvbl1 +/- mice<br />

eventually developed larger tumors that control floxed mice<br />

9 and 12 months after tumor induction. Ruvbl1 +/- mice (not<br />

DEN-injected) showed impaired basal hepatocyte turnover and<br />

strikingly lower regeneration after 70% hepatectomy, confirming<br />

that Ruvbl1 support cell growth in vivo. We observed that<br />

in the hemizygous mice Ruvbl1 expression was reduced by<br />

50% in normal hepatocytes but not within the tumors, which<br />

indeed overexpressed Ruvbl1 regardless of the mouse genetic<br />

background, suggesting that Ruvbl1 overexpression is a central<br />

event in HCC onset. Hepatic-Ruvbl1 +/- mice developed mild<br />

hyperglycemia, impaired glucose tolerance, insulin resistance,<br />

had reduced liver glycogen content and increased somatic<br />

growth. Moreover, mRNA expression of several mTOR targets<br />

(SREBP-1c, HIF1α, PGC-1α, PPARγ) were reduced in Ruvbl1 +/-<br />

mice. Ruvbl1 silencing in non-transformed hepatocytes (AML-<br />

12) resulted in reduced mTOR expression and impaired<br />

PI3K-Akt-mTOR signaling in response to insulin, as well reduced<br />

cell growth. By 2D-DIGE proteomics we identified key enzymes<br />

of glycolysis as down-regulated proteins in Ruvbl1-silenced<br />

hepatoma cells. Taken together these results highlight a novel<br />

role of Ruvbl1 in the maintenance of hepatic glucose homeostasis<br />

and insulin-sensitivity through the Akt-mTOR pathway, and<br />

strongly suggest that deregulation of this pathway by Ruvbl1<br />

overexpression is a key event driving HCC progression.<br />

Disclosures:<br />

Stefano Milani - Grant/Research Support: Gilead Sciences, Merck Sharp and<br />

Dohme, Abbvie<br />

The following authors have nothing to disclose: Tommaso Mello, Francesca Zanieri,<br />

Elisabetta Ceni, Mirko Tarocchi, Giada Marroncini, Simone Polvani, Sara<br />

Tempesti, Oxana Bereshchenko, Andrea Galli<br />

1964<br />

MicroRNA-207 controls Prp19 expression in hepatocellular<br />

carcinoma<br />

Ji-Min Zhu, Xi-Zhong Shen; Department of Gastroenterology,<br />

Zhongshan Hospital of Fudan University, Shanghai, China<br />

INTRODUCTION: microRNAs (miRNAs) are small non-coding<br />

RNA molecules of 21-24 nt that regulate the expression of<br />

numerous target genes by transcriptional inhibition or translational<br />

repression. Multiple lines of evidence suggest that miR-<br />

NAs play important roles in tumor (such as, hepatocellular<br />

carcinoma (HCC)) development, progression, invasion, metastasis<br />

and prognosis. Our previous unpublished work suggested<br />

that microRNA-207 (miR-207) upregulates in HCC tissues than<br />

in normal liver tissues. Unfortunately, our understanding of the<br />

molecular mechanisms that governs its role in development,<br />

progression, invasion, metastasis and prognosis remains fragmentary.<br />

AIMS & METHODS: In this study, a genome-wide<br />

miRNA microarray was used to identify differentially expressed<br />

miRNAs in HCCs patients. Next, miR-207 expression in normal<br />

liver tissues, HCC tissues and adjacent non-tumor tissues was<br />

validated, and the predictive values of miR-207 for the prognosis<br />

of HCC patients were explored using χ2 test, Student’s<br />

t test, paired t test, and Mann-Whitney test. The biological<br />

roles of miR-207 and its underlying roles in HCC were also<br />

investigated in the Huh7, MHCC-97H, SMMC-7721, L02 and<br />

Hep3B cell lines. RESULTS: Forty-two miRNAs were differentially<br />

expressed in HCCs. Several of these miRNAs were previously<br />

found deregulated in other cancers. Additionally, the<br />

miR-207 was found upregulated in ∼80% of HCCs and in all<br />

HCC-derived cell lines. Overexpressed intratumoral miR-207<br />

was associated with poor survival rate (P < 0.001), and was<br />

an independent prognostic factor for overall survival rate (P <<br />

0.001) in the patients with HCC. Overexpression of miR-207 in<br />

L02 cell line remarkably enhanced cell proliferation, migration,<br />

and invasion; while inhibition of miR-207 in Hep3B cell line<br />

caused the opposite effects. Further study found that miR-207<br />

suppressed the expression of pre-Mrna processing factor 19<br />

(Prp19) by directly binding to its 3’-untranslated region, and<br />

sensitized Hep3B cell to CDDP-induced apoptosis. Moreover,<br />

miR-207 expression levels correlated positively with Prp19 in<br />

human HCC tissues. Western blot showed that overexpression<br />

of miR-207 in L02 cell upregulates the expression of Prp19,<br />

while inhibition of miR-207 in Hep3B cell downregulates the<br />

expression of Prp19, suggesting that Prp19 might play as a<br />

oncogene in HCC. CONCLUSION: Taken together, our findings<br />

suggest that miR-207 could play a role in the development of<br />

HCC, at least in part, by modulating Prp19. These findings<br />

establish a basis toward the development of therapeutic strategies<br />

aimed at blocking miR-207 in HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ji-Min Zhu, Xi-Zhong Shen


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1165A<br />

1965<br />

Somatic Mutational Analysis by Next Generation RNA<br />

Sequencing Reveals Frequent Mutation of Chromatin<br />

and Chromosome Remodeling Genes in Gallbladder<br />

Cancer<br />

Loretta K. Allotey 1 , Roongruedee Chaiteerakij 1,2 , Gavin R. Oliver<br />

3 , Vivekananda Sarangi 3 , Renumathy Dhanasekaran 1 , Catherine<br />

D. Moser 1 , Nasra H. Giama 1 , Daniel R. O’Brien 3 , Raymond<br />

M. Moore 3 , Mia D. Champion 4 , Eric W. Klee 3 , Mitesh J. Borad 5 ,<br />

Lewis R. Roberts 1 ; 1 Division of Gastroenterology and Hepatology,<br />

Mayo Clinic College of Medicine, and Mayo Clinic Cancer Center,<br />

Rochester, MN; 2 Department of Medicine, Faculty of Medicine,<br />

Chulalongkorn University and King Chulalongkorn Memorial Hospital,<br />

Thai Red Cross Society, Bangkok, Thailand; 3 Department of<br />

Biomedical Statistics and Informatics, Mayo Clinic, Rochester, MN;<br />

4 Department of Biomedical Statistics and Informatics, Mayo Clinic,<br />

Scottsdale, AZ; 5 Division of Hematology and Medical Oncology,<br />

Mayo Clinic, Scottsdale, AZ<br />

Background/Aim: Gallbladder cancer (GBC) is uncommon,<br />

with an incidence of 2 per 100,000/year in the US. Due<br />

to its asymptomatic nature, most patients present at a late<br />

stage. Current standard chemotherapy and radiation therapy<br />

only achieve a 5-year survival rate of 10% in patients with<br />

advanced cancer. Thus improved understanding of the molecular<br />

pathogenesis of GBC is urgently needed for rational design<br />

of effective therapies. We therefore aimed to investigate the<br />

mutational landscape of human gallbladder cancer. Methods:<br />

12 GBCs were RNA sequenced on the Illumina Hi-Seq 2000<br />

platform. Single Nucleotide Variants (SNVs) were called from<br />

the genome-aligned reads using Unified Genotyper. Stringent<br />

filtering criteria were used to identify variants of high<br />

confidence (Inclusion criteria: Depth of Coverage>10, Quality<br />

Score>0.05, Minor Allele Frequency1<br />

patient). To identify variants in clinically relevant genes, variants<br />

meeting the above criteria were selected and compared to<br />

a generated list of 1,434 genes reported in COSMIC, Cancer<br />

Panels, or in the literature to have relevance to cancer and epigenetics<br />

processes. Survival in months was determined for each<br />

patient and select genes were investigated for their association<br />

with survival by Kaplan Meier analysis. The significance of the<br />

differences in survival between groups was analyzed using the<br />

Log-Rank test. Results: 69 SNVs were identified in 28 clinically<br />

relevant genes. Fifteen of the 28 genes (53%) have known<br />

effects on epigenetic processes, including TP53 (mutated in<br />

50% of tumors), SETD1A (33%), MLL2 (33%), SRCAP (25%),<br />

MSH6 (16%), MSL1 (16%), ATXN7 (16%), SUV420H1 (16%),<br />

LSG1 (16%), RSF1 (16%), MCM7 (16%), DNAJC2 (16%),<br />

AHCTF1 (16%), BRD4 (16%), and NUP214 (16%). DAVID<br />

Functional annotation and WebGestalt Gene Ontology Analysis<br />

of the 28 genes revealed that 11 genes (ATXN7, SRCAP,<br />

DNAJC2, MSL1, SUV420H1, BRD4, MLL2, SETD1A, TP53,<br />

and RSF1) are involved in chromatin and chromosome remodeling.<br />

Of the 11 genes, mutations in the SRCAP gene were<br />

found to have a marginally significant effect on survival. The<br />

three patients with mutation in the SRCAP gene had a survival<br />

of 4 months while patients without mutations in SRCAP had<br />

a survival of 9 months (P = 0.06). Conclusions: Chromatin<br />

and chromosome remodeling genes are frequently mutated in<br />

GBC and provide a starting point for further functional <strong>studies</strong>.<br />

Patients with SRCAP mutations show a trend towards poorer<br />

survival than patients with wild-type SRCAP. Validation of these<br />

findings in a larger cohort is underway.<br />

Disclosures:<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Loretta K. Allotey, Roongruedee<br />

Chaiteerakij, Gavin R. Oliver, Vivekananda Sarangi, Renumathy Dhanasekaran,<br />

Catherine D. Moser, Nasra H. Giama, Daniel R. O’Brien, Raymond M. Moore,<br />

Mia D. Champion, Eric W. Klee, Mitesh J. Borad<br />

1966<br />

Differential Antitumor Effects of FGFR Inhibitors on a<br />

Novel Cholangiocarcinoma Patient-Derived Xenograft<br />

Mouse Model Expressing an FGFR2-CCDC6 Fusion Protein<br />

Yu Wang 1,2 , Hassan M. Shaleh 2 , Xiwei Ding 2,3 , Shaoqing<br />

Wang 2,4 , Roongruedee Chaiteerakij 2,5 , Kais Zakharia 2 , Loretta<br />

K. Allotey 2 , Essa A. Mohamed 2 , Catherine D. Moser 2 , Steven<br />

Alberts 6 , Joseph M. Gozgit 7 , Mitesh J. Borad 8 , Lewis R. Roberts 2 ;<br />

1 Department of Hepatobiliary Surgery, Nanfang Hospital, Southern<br />

Medical University, Guangzhou, China; 2 Division of Gastroenterology<br />

and Hepatology, Mayo Clinic College of Medicine, and<br />

Mayo Clinic Cancer Center, Rochester, MN; 3 Department of Gastroenterology,<br />

Drum Tower Hospital, Affiliated to Medical School<br />

of Nanjing University, Nanjing, China; 4 Department of Pathology,<br />

Qiqihar Medical University, Qiqihar, China; 5 Department of Medicine,<br />

Faculty of Medicine, Chulalongkorn University and King<br />

Chulalongkorn Memorial Hospital, Thai Red Cross Society, Bangkok,<br />

Thailand; 6 Department of Oncology, Mayo Clinic, Rochester,<br />

MN; 7 ARIAD Pharmaceuticals, Inc., Cambridge, MA; 8 Division of<br />

Hematology and Medical Oncology, Mayo Clinic, Scottsdale, AZ<br />

Introduction: Cholangiocarcinoma (CCA) is a highly lethal<br />

cancer with limited therapeutic options and a poor prognosis.<br />

Recent genomic analysis has revealed the presence of<br />

FGFR2 fusion proteins in up to 13% of intrahepatic CCAs.<br />

The FGFR fusion proteins appear to play a key role in cancer<br />

progression. Tumors harboring FGFR fusions have demonstrated<br />

enhanced sensitivity to FGFR inhibitors, suggesting that<br />

CCA patients with FGFR2 fusions may benefit from targeted<br />

FGFR2 kinase inhibition. The effects of FGFR kinase inhibitors<br />

ponatinib, dovitinib and BGJ398 have not been evaluated in<br />

CCA bearing FGFR fusions. We aimed to assess the relative<br />

sensitivity of a novel in vivo CCA PDX mouse model bearing<br />

an FGFR2-CCDC6 fusion protein to different FGFR inhibitors.<br />

Methods: LIV31 is a PDX model derived in NSG SCID mice<br />

from a metastatic lung nodule resected from a patient with<br />

stage IV intrahepatic CCA. Subsequent passage of the PDX<br />

tumor into nude mice was successful. Detailed assay confirmed<br />

that LIV31 harbors a transcript encoding an FGFR2-CCDC6<br />

fusion protein. LIV31 cells were implanted subcutaneously into<br />

the flanks of nude mice. When tumor volumes reached 150-<br />

250 mm3 the mice were randomized into four groups: vehicle,<br />

ponatinib 25 mg/kg, dovitinib 30 mg/kg, or BGJ398 15<br />

mg/kg. The drugs were administered daily by oral gavage at<br />

doses previously shown to be tolerated by mice. Tumor volumes<br />

and mouse weights were measured once a week. Tumor<br />

growth curves were compared. After 63 days treatment, the<br />

animals were euthanized and xenografts were examined by<br />

histology, Ki-67 IHC, Tunel staining and Western blotting for<br />

pFGFR2 and the FGFR pathway downstream mediators pFRS2,<br />

pAKT, and pERK. Results: All three FGFR inhibitors significantly<br />

inhibited the growth of the FGFR2-CCDC6 fusion mouse PDX<br />

tumors when compared with vehicle (P


1166A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The Ki-67 and Tunel staining results show that these inhibitors<br />

can also inhibit proliferation and induce apoptosis of the CCA<br />

xenograft cells. Conclusions: These three TKIs all significantly<br />

inhibited the growth of patient-derived CCA xenografts bearing<br />

an oncogenic FGFR2-CCDC6 fusion gene. Treatment efficacy<br />

from most to least effective was BGJ398 > dovitinib > ponatinib.<br />

Additional <strong>studies</strong> exploring the potential mechanisms<br />

underlying differences in response to particular FGFR inhibitors<br />

and the effects of combinations of other therapeutic agents with<br />

FGFR inhibitors are underway.<br />

Disclosures:<br />

Joseph M. Gozgit - Employment: ARIAD Pharmaceuticals<br />

Lewis R. Roberts - Grant/Research Support: Bristol Myers Squibb, ARIAD Pharmaceuticals,<br />

BTG, Wako Diagnostics, Inova Diagnostics, Gilead Sciences, Five<br />

Prime Therapeutics<br />

The following authors have nothing to disclose: Yu Wang, Hassan M. Shaleh,<br />

Xiwei Ding, Shaoqing Wang, Roongruedee Chaiteerakij, Kais Zakharia, Loretta<br />

K. Allotey, Essa A. Mohamed, Catherine D. Moser, Steven Alberts, Mitesh J.<br />

Borad<br />

1967<br />

Glypican 3 contributes to HCC invasiveness and tumorigenesis<br />

by up-regulating Angioprotein 2 and initiating<br />

epithelial mesenchymal transition<br />

Thu Le Trinh, William M. Puszyk, Olorunseun Ogunwobi, Quanfeng<br />

Wu, Chen Liu; Department of Pathology, University of Florida,<br />

Gainesville, FL<br />

Background and Rationale: Glypican 3 (GPC3) has been<br />

shown to be up-regulated in HCC and is known to be associated<br />

with HCC advanced stages as well as cancer invasiveness<br />

and recurrence. The aim of this study is to define the<br />

mechanisms of GPC3 in HCC carcinogenesis and metastasis.<br />

Methods: Representations of murine and human HCCs were<br />

created using GPC3 knock-in in 1MEA (1MEA hGPC3 ), a murine<br />

HCC cell line which doesn’t have GPC3 expression; and GPC3<br />

knock-down in Huh7 (Huh7 GPC3 shRNA ), a human HCC cell line<br />

which highly expresses GPC3. Cell proliferation and tumor<br />

development were access in vitro and in vivo using the 1MEA/<br />

Balb/c and Huh7/NSG mice models. The impact of GPC3<br />

knock down on different cancer pathways was approached by<br />

qPCR array analysis on the Huh7 control shRNA and GPC3-knockdown<br />

Huh7 GPC3 shRNA cell lines using Cancer Pathway Finder<br />

array (Qiagen), which consists of 84 different genes that represent<br />

9 different biological pathways involved in carcinogenesis.<br />

The effect of GPC3 repression on cell invasion ability<br />

was also examined by wound-healing assay. Results: We<br />

found that GPC3 didn’t induce a different proliferation rate<br />

in 1MEA hGPC3 cell line, but it promoted tumor development<br />

significantly when 1MEA hGPC3 was implanted into Balb/c mice<br />

compared to 1MEA control. Knock-down of GPC3 resulted in<br />

cell growth inhibition and cell cycle arrest at S phase. Similarly,<br />

tumor proliferation was greatly retarded in Huh7 GPC3<br />

shRNA<br />

xenograft mouse model in comparison to Huh7control shRNA<br />

tumor. In the 84 cancer pathway focused genes, 29 genes<br />

showed at least a 2-fold difference in gene expression between<br />

GPC3-knockdown cells compared to non-silenced cells. These<br />

included genes that maintain telomeres (TINKS, TINKS2, TINF2,<br />

TERF2IP), anti-apoptotic genes (XIAP, NOL3, FASLG), genes<br />

involved in cell cycle regulation (CCND2, CCND3, E2F4),<br />

markers of epithelial to mesenchymal transition (EMT) (CDH2,<br />

DSP, OCLN, SNAI3), and genes that promote angiogenesis<br />

(ANGPT2). Among these, ANGPT2 gene expression was<br />

down-regulated up to 150 fold in Huh7 shRNA GPC3 compared to<br />

Huh7 control shRNA . GPC3 roles in EMT were confirmed by migration<br />

assay and showed significant inhibition of Huh7GPC3 shRNA<br />

migration. Likewise, EMT markers and inducers were upregulated<br />

in 1MEA hGPC3 cell line in comparison to 1MEA control.<br />

Conclusions: These results indicated that GPC3 played a role<br />

in HCC tumorigenesis via promoting EMT, angiogenesis and<br />

cell migration, and that GPC3 could serve as a key connection<br />

between angiogenesis and metastasis in HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Thu Le Trinh, William M. Puszyk,<br />

Olorunseun Ogunwobi, Quanfeng Wu, Chen Liu<br />

1968<br />

Promoted malignant tumor phenotype of hepatitis C<br />

virus transgenic mice fed an atherogenic high-fat diet is<br />

associated with the abnormal expression of the glycolysis-related<br />

gene pyruvate kinase M2.<br />

Riuta Takabatake 1 , Masao Honda 1 , Hikari Okada 1 , Kai Takegoshi<br />

1 , Yoshio Sakai 1 , Taro Yamashita 1 , Naoto Nagata 3 , Toshinari<br />

Takamura 3 , Takuji Tanaka 2 , Shuichi Kaneko 1 ; 1 Gastroenterology,<br />

Kanazawa University Graduate School of Medical Science,<br />

Kanazawa, Japan; 2 Cancer Research and Prevention, The Tohkai<br />

Cytopathology Institute, Gifu, Japan; 3 Department of Disease Control<br />

and Homeostasis, Kanazawa University Graduate School of<br />

Medical Sciences, Kanazawa, Japan<br />

Objective The relationship between hepatitis C virus (HCV)-related<br />

metabolic abnormalities and the progression of hepatic<br />

tumorigenesis remains unclear. The expression of the HCV<br />

open reading frame in the liver of transgenic (Tg) mice induces<br />

steatohepatitis, which is followed by the development of hepatocellular<br />

carcinoma. The atherogenic high-fat diet (Ath HFD)<br />

mouse model develops steatohepatitis similar to human nonalcoholic<br />

steatohepatitis. We investigated the expression of<br />

pyruvate kinase M2 (PKM2), an isoenzyme of the glycolytic<br />

enzyme pyruvate kinase and recently recognized as a transcription<br />

factor promoting cancer cell growth and the progression<br />

of tumorigenesis, in HCV Tg mice fed an Ath HFD<br />

diet. Materials and Methods Four groups of mice (n = 15–20/<br />

group) were fed with a control diet or an Ath HFD diet starting<br />

at 8 weeks after birth for 30 and 60 weeks. The degree of<br />

liver steatosis, liver fibrosis, liver weight, and tumor incidence<br />

were measured. The expression of collagen I and IV, TGF-β,<br />

p-SMAD3, PKM2, Foxo1, p-Akt, Hes1, and Hey1 was evaluated<br />

by immunohistochemical staining, TaqMan PCR, and<br />

western blotting. In vitro analysis was performed by using<br />

PKM2-stable knockdown Huh-7 cells. Results HCV Tg mice fed<br />

an Ath HFD diet showed markedly promoted liver fibrosis at<br />

30 and 60 weeks compared with wild-type (WT) mice fed an<br />

Ath HFD diet. In addition, HCV Tg mice fed an Ath HFD diet<br />

showed markedly increased liver weight (1.5-fold) and tumor<br />

size (1.5-fold), but with no change in tumor frequency compared<br />

to WT mice. HCV Tg mice fed an Ath HFD diet showed<br />

a significant increase in the expression of PKM2 and fibrosis-related<br />

genes, such as collagen I/IV, p-smad3, and TGF-β<br />

together with Notch signal-related genes. HCV core protein<br />

was found to activate the PKM2 promoter through the induction<br />

of CREB. In vitro, PKM2-stable knockdown Huh-7 cells showed<br />

significant repression of EpCAM, AFP, Sox2, Oct4, vimentin,<br />

and KRT19 expression, which are markers of cancer stem cells.<br />

Interestingly, the addition of recombinant Notch/jagged1 to<br />

Huh-7 cells further activated the nuclear translocation of PKM2;<br />

conversely, Notch-stimulated Hes1 and Hey1 expression was<br />

repressed in PKM2-knockdown cells. Conclusion Promoted liver<br />

fibrosis and a malignant tumor phenotype were observed in<br />

HCV Tg mice fed an Ath HFD diet, which were associated with<br />

the abnormal expression of PKM2. PKM2 might be involved in<br />

the expression of pro-fibrotic genes and epithelial-mesenchymal<br />

transition signaling by activation of the Notch signaling


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1167A<br />

pathway. These findings could reveal the relationships between<br />

glucose metabolism, liver fibrosis, and tumorigenesis in chronic<br />

hepatitis C.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan<br />

The following authors have nothing to disclose: Riuta Takabatake, Masao<br />

Honda, Kai Takegoshi, Yoshio Sakai, Taro Yamashita, Naoto Nagata, Toshinari<br />

Takamura, Takuji Tanaka<br />

1969<br />

Colorectal liver metastasis is enhanced in the alcoholic<br />

liver: role of ethanol-mediated alterations in the liver<br />

microenvironment and carcinoembryonic antigen processing<br />

Ashley M. Mohr 4 , Robert S. Bridge 1 , Peter Thomas 3 , John Gould 2 ,<br />

Jacy L. Kubik 1 , Carol A. Casey 2,1 , Dean J. Tuma 1 , Benita L.<br />

McVicker 2,1 ; 1 Internal Medicine, University of Nebraska Medical<br />

Center, Omaha, NE; 2 VA Nebraska-Western Iowa Health Care<br />

System, Omaha, NE; 3 Surgery & Biomedical Sciences, Creighton<br />

University, Omaha, NE; 4 Biochemistry & Molecular Biology, University<br />

of Nebraska Medical Center, Omaha, NE<br />

Colorectal cancer (CRC) is a leading cause of cancer mortality.<br />

The majority of CRC deaths are associated with the development<br />

of colorectal liver metastasis (CRLM). The characterization<br />

of CRLM including the role of the liver environment is ongoing<br />

and vital for the development of treatment options. Although it<br />

is known that alcohol consumption is an independent risk factor<br />

for CRLM, the mechanisms by which alcohol affects the environment<br />

influencing metastasis are undefined. It has been shown<br />

that the liver processing of a CRC-associated protein, carcinoembryonic<br />

antigen (CEA), correlates with CRLM. In particular,<br />

CEA is processed by Kupffer cells (KCs) and hepatocytes<br />

(HCs) resulting in KC-produced cytokines or the physiological<br />

degradation of CEA in HCs; events that respectively can potentiate<br />

or regulate CRLM. Interestingly, alcoholic liver disease<br />

(ALD) is associated with increased serum CEA levels. However,<br />

alcohol’s effects on CEA processing and CRLM is not known.<br />

In the current study, we examined the role of alcohol in a novel<br />

model of ALD and CRLM mediated by human CEA-expressing<br />

LS174T CRC cells. Methods: Immunodeficient Rag1 tm1Mom mice<br />

were fed control (C) or ethanol (E) diets for four weeks followed<br />

by intrasplenic injection of LS174T cells. The C/E diets were<br />

continued for 3-5 more weeks at which time ALD and CRLM<br />

were assessed. Results: ALD was confirmed by several parameters<br />

including decreased HC endocytosis via the asialoglycoprotein<br />

receptor (ASGPR), a well-defined early impairment in<br />

ALD and receptor-mediated degradation of CEA. Importantly,<br />

the rate and burden of CRC tumors was enhanced in the alcohol-fed<br />

mice. Specifically, 3 weeks following LS174T injections,<br />

50% of E-fed mice had CRLM compared to no evidence of<br />

metastasis in controls. By 5 weeks, all C/E mice had CRLM yet<br />

the tumor size and number were markedly increased in alcohol-affected<br />

livers. Analysis of liver tissue adjacent to tumors<br />

indicated enhanced levels of CEA and KC-associated cytokines<br />

(TNF-α, IL1-b, IL-10). Moreover, adhesion molecules (ICAM-1<br />

and E-selectin) known to be involved in CRLM in response to<br />

KC-produced cytokines, were elevated (2-fold) in E-fed mice.<br />

Overall, ALD exacerbates CRLM by potentiating CEA-mediated<br />

KC cytokine production and downstream adhesion molecules<br />

facilitating CRC adherence in the liver. Additionally,<br />

alcohol-mediated defects to ASGPRs led to impaired CEA degradation<br />

that can contribute to pro-metastatic signaling. Further<br />

investigation could significantly impact the understanding of<br />

involved mechanims as well as identify potential targets for the<br />

treatment of colorectal metastases in the alcohol-injured liver.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ashley M. Mohr, Robert S.<br />

Bridge, Peter Thomas, John Gould, Jacy L. Kubik, Carol A. Casey, Dean J. Tuma,<br />

Benita L. McVicker<br />

1970<br />

Fibrotic liver microenvironment promotes stemness of<br />

liver cancer through the activation of TGF-beta and<br />

histone demethylase KDM2B in a p53-status dependent<br />

manner.<br />

Taro Yamashita, Mitsumasa Kondo, Hikari Okada, Naoki Oishi,<br />

Masao Honda, Shuichi Kaneko; Department of Gastroenterology,<br />

Kanazawa University Graduate School of Medical Science,<br />

Kanazawa, Japan<br />

[Background] Liver cirrhosis is one of the most important risk<br />

factors for the development of hepatocellular carcinoma (HCC).<br />

Fibroblasts are known to regulate malignant nature of tumors,<br />

but its mechanisms remains to be elucidated in detail. In this<br />

study, we evaluated the role of fibroblasts on stemness of the<br />

tumor, which is closely associated with the aggressiveness of<br />

liver cancer. [Methods]Primary HCC cells obtained from surgically<br />

resected specimens as well as Huh7, Hep3B, and HepG2<br />

cells were co-cultured with various fibroblasts in vitro using cell<br />

culture inserts. Gene and protein expression was evaluated<br />

by qRT-PCR and Western blotting. Fluorescence-activated cell<br />

sorting (FACS) was used to evaluate the frequency of cancer<br />

stem cells expressing EpCAM/CD133. Cancer stem cell characteristics<br />

were evaluated by spheroid formation, invasion,<br />

and tumorigenicity in immune deficient mice. Time-lapse image<br />

analysis was performed to monitor cell motility affected by stromal<br />

cells. KDM2B and H3K36 methylation status was evaluated<br />

by Western blotting and immunofluorescence. [Results] Co-culture<br />

of HCC cells with fibroblasts resulted in the enhanced cell<br />

motility, spheroid formation, and invasion capacities of HCC<br />

cells. FACS analyses demonstrated the enrichment of EpCAM/<br />

CD133-positive Huh7 cells when co-cultured with fibroblasts.<br />

Gene expression analysis of selected cytokines and growth factors<br />

indicated the abundant expression of TGFB1 in fibroblasts.<br />

Supplementation of recombinant transforming growth factor<br />

beta (TGF-beta) enhanced the spheroid formation capacity of<br />

HCC cells, and this effect was abolished when cultured with<br />

neutralizing antibodies against TGF-beta. Co-injection of HCC<br />

cells with fibroblasts resulted in the high frequency of lung<br />

metastasis. Interestingly, acquisition of stemness induced by<br />

TGF-beta was associated with the activation of histone demethylase<br />

KDM2B in p53-mutant Huh7 cells but not in p53-wildtype<br />

HepG2. Introduction of mutant p53 (R175H) enhanced the<br />

activation of KDM2B induced by TGF-beta in p53-null Hep3B<br />

cells. [Conclusions] Fibrotic liver microenvironment accelerates<br />

the malignant natures of HCC with enrichment of EpCAM/<br />

CD133-positive cancer stem cells through the activation of TGFbeta<br />

signaling. TGF-beta may activate histone demethylase<br />

KDM2B to modify epigenetic status to induce the stemness of<br />

HCC in a p53-status dependent manner.<br />

Disclosures:<br />

Hikari Okada - Employment: Kanazawa University<br />

Shuichi Kaneko - Grant/Research Support: MDS, Co., Inc, Chugai Pharma., Co.,<br />

Inc, Toray Co., Inc, Daiichi Sankyo., Co., Inc, Dainippon Sumitomo, Co., Inc,<br />

Ajinomoto Co., Inc, Bristol Myers Squibb., Inc, Pfizer., Co., Inc, Astellas., Inc,<br />

Takeda., Co., Inc, Otsuka„ÄÄPharmaceutical, Co., Inc, Eizai Co., Inc, Bayer<br />

Japan, Eli lilly Japan


1168A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Taro Yamashita, Mitsumasa<br />

Kondo, Naoki Oishi, Masao Honda<br />

1971<br />

Development of HCC is dependent upon hepatocellular<br />

apoptosis in a murine model of NASH associated tumoriogenesis<br />

Casey Johnson 1 , Alexander Wree 1,2 , Joan Font-Burgada 3 , Akiko<br />

Eguchi 1 , Davide Povero 1 , Michael Karin 3 , Ariel E. Feldstein 1 ;<br />

1 Pediatrics, UCSD, La Jolla, CA; 2 Department of Internal Medicine<br />

III, University Hospital RWTH-Aachen, Aachen, Germany; 3 Laboratory<br />

of Gene Regulation and Signal Transduction, Departments of<br />

Pharmacology and Pathology, UCSD, La Jolla, CA<br />

Background: Nonalcoholic steatohepatitis (NASH) is increasingly<br />

associated with the development of hepatocellular carcinoma<br />

(HCC), even in patients without established liver cirrhosis.<br />

While the oncogenic mechanisms of blocking persistent hepatocyte<br />

cell death, a common feature among various chronic liver<br />

diseases, remain unclear it nevertheless presents as a logical<br />

treatment modality. Therefore, we aimed at investigating the<br />

influence of hepatocyte specific Bid depletion – a BH3-only<br />

Bcl-2 family member that amplifies apoptotic death signals –<br />

on tumor development in a murine model of NASH induced<br />

HCC. Methods: Hepatocyte-specific conditional Bid-knockout<br />

mice (Bid Δhep ) and functional Bid-expressing control mice (Bidflo/flo<br />

) were injected with 100mg/Kgr streptozotocin (STZ) at<br />

2-days of age. Mice were weaned onto a 60% Kcal from fat<br />

diet and liver tumorigenesis was investigated 5 months later.<br />

Results: Western Blot analysis confirmed hepatocyte specific<br />

Bid depletion in Bid Δhep mice. Survival rate was impaired in<br />

Bid flo/flo when compared to Bid Δhep mice (Bid flo/flo 3/12 vs.<br />

Bid Δhep 0/12; p < 0.05). Bid flo/flo mice exhibited higher incidences<br />

of liver tumors larger than 5mm when compared to<br />

Bid Δhep mice, which translated into an increased tumor load<br />

(Bid flo/flo 10/10 vs. Bid Δhep 4/10; p < 0.05). Liver transaminase<br />

levels remained normal within Bid Δhep mice, while Bid flo/<br />

flo<br />

mice were elevated (Bid Δhep 249 IU/l ±12 IU/l (mean ±<br />

SEM) vs. Bid flo/flo 84±12; p < 0.05). Tissue section quantification<br />

revealed an increased rate of hepatocellular apoptosis<br />

in Bid flo/flo mice when compared to Bid Δhep mice. Quantitative<br />

PCR analysis of mRNA levels of AFP (alphafetoprotein), HGF<br />

(hepatocyte growth factor), Cyclin D1 and PCNA were significantly<br />

increased in Bid flo/flo mice as compared to Bid Δhep<br />

animals indicating cellular proliferation. Liver sections of Bid flo/<br />

flo<br />

mice also exhibited increased infiltration of inflammatory<br />

cells and concomitant upregulation of mRNA levels of F4/80,<br />

Ly6c, TNFa (tumor necrosis factor), and IL-1beta (interleukin<br />

1 beta). Liver fibrosis developed in both mouse groups, while<br />

mRNA levels indicating hepatic stellate cell activation (aSMA<br />

– smooth muscle actin; and TIMP1 – tissue inhibitor of metalloproteinase)<br />

were significantly increased in Bid flo/flo mice when<br />

compared to Bid Δhep mice. Conclusion: Our study demonstrates<br />

that blocking hepatic apoptosis ameliorated HCC development<br />

in a NASH-related tumorigenesis model. These results suggest<br />

that reducing hepatocyte cell death, liver inflammation and<br />

compensatory proliferation presents a strong beneficial effect<br />

that supersedes the potential side effect of enhancing tumor<br />

cell survival.<br />

Disclosures:<br />

Akiko Eguchi - Grant/Research Support: Gilead, Conatus<br />

The following authors have nothing to disclose: Casey Johnson, Alexander Wree,<br />

Joan Font-Burgada, Davide Povero, Michael Karin, Ariel E. Feldstein<br />

1972<br />

Cytotoxic synergy between sorafenib and cyclooxygenase<br />

2 inhibitor celecoxib in vitro: induction of cell death<br />

of hepatocellular carcinoma through ER stress mediated<br />

apoptosis<br />

Sera Yang 1 , SoHee Kang 1 , Soohyun Park 2 , Jonghwa Kim 1 , Yong<br />

Han Paik 1,2 ; 1 Department of Medicine, Samsung Medical Center,<br />

Sungkyunkwan University School of Medicine, Seoul, Korea<br />

(the Republic of); 2 Department of Health Science and Technology,<br />

Samsung Advanced Institute for Health Science and Technology,<br />

Sungkyunkwan University, Seoul, Korea (the Republic of)<br />

Background/Aim: Sorafenib, a multi-kinase inhibitor, is currently<br />

recommended for the treatment of advanced hepatocellular<br />

carcinoma (HCC). However, the response rate of sorafenib<br />

in advanced HCC treatment are not satisfactory and there is<br />

a rationale for investigating its use in combination with other<br />

agents. Celecoxib is known as a selective cyclooxygenase-2<br />

(COX-2) inhibitor and has been known to exhibit anti-tumor<br />

effects in HCC cells but the mechanism is still largely unknown.<br />

The aim of this study was to investigate the synergistic effect of<br />

celecoxib on the anti-tumor activity of sorafenib in HCC and<br />

to determine the underlying molecular mechanism. Methods:<br />

HCC cell line PLC/PRF/5 were treated with sorafenib and/<br />

or celecoxib, and then proliferation was analyzed by MTS<br />

assay. Cell death was evaluated by annexin staining and<br />

FACS analysis. Western blot was also used to study the mechanism<br />

of the induced cell death. Results: MTS assay revealed<br />

that each of sorafenib and celecoxib alone marginally inhibited<br />

HCC cell proliferation. However, the combination of<br />

sorafenib and celecoxib showed a synergistic effect to inhibit<br />

cell proliferation, while the effect was less strongly observed<br />

with 2,5-dimethyl-celecoxib (DMC, a derivative of celecoxib<br />

that lacks COX2-inhibitory function). Using flow cytometry, we<br />

found that the combined treatment synergistically increased the<br />

population of Annexin + /PI + at 48h, indicating the increased<br />

apoptosis. The increased cleavage of caspase-8 and PARP was<br />

also demonstrated in Western blot. In addition, we also found<br />

that ER stress marker proteins, such as GRP78/Bip, CHOP,<br />

and spliced XBP-1, were dramatically increased in the combined<br />

treatment of sorafenib and celecoxib. The activation of<br />

caspases and cell death induced by sorafenib plus celecoxib<br />

was significantly inhibited by pretreatment with 4μ8C, a ER<br />

stress inhibitor. These data indicated that sorafenib and celecoxib<br />

synergistically induce cell death of HCC cell line through<br />

ER-stress-mediated apoptosis. Conclusion: Combined treatment<br />

of sorafenib and celecoxib synergistically inhibit the growth<br />

of HCC cells via inducing ER-stress mediated apoptosis. This<br />

finding raises the possibility of the combined treatment using<br />

sorafenib and celecoxib for the augmentation of anticancer<br />

effect of sorafenib in HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sera Yang, SoHee Kang,<br />

Soohyun Park, Jonghwa Kim, Yong Han Paik


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1169A<br />

1973<br />

PPPDE1 is a deubiquitinase that promotes HCC development<br />

by stabilizing a 25kd MDM2 N terminal fragment<br />

and subsequently suppressing the p53 pathway<br />

Xing-Wang Xie 1,2 , Yang-Jing Zhao 1,2 , Xue-Yan Wang 1,2 , Ran<br />

Fei 1,2 , Heng-Hui Zhang 1,2 , Wei-Jia Liao 3 , Lai Wei 1,2 , Yu Wang 4 ,<br />

Hong-Song Chen 1,2 ; 1 Peking University People’s Hospital, Peking<br />

University Hepatology Institute, Beijing, China; 2 Beijing Key Laboratory<br />

of Hepatitis C and Immunotherapy for Liver Diseases, Beijing,<br />

China; 3 Guilin Medical College, Guilin, China; 4 Chinese<br />

Center for Disease Control and Prevention, Beijing, China<br />

Introduction: In our previous study, PPPDE1 (DESI2) was identified<br />

and validated as a HCC driver gene, but the underlying<br />

mechanism was unclear. The PPPDE1 protein belongs to a protein<br />

family named PPPDE (Permuted Papain fold Peptidases of<br />

Ds-RNA viruses and Eukaryotes) that was predicted to be a<br />

deubiquitinating enzyme (DUB) family by protein homology<br />

search. Hence, this study aims to validate the DUB activity of<br />

PPPDE1 and to elucidate how it is involved in the HCC development.<br />

Method: Lentvirus expressing wild type PPPDE1 and<br />

PPPDE1 C108S mutant (the cysteine 108 was replaced by serine)<br />

were used for over-expression and protein purification. Lentivirus<br />

expressing PPPDE1 specific shRNAs were used to silencing<br />

PPPDE1. The DUB activity of PPPDE1 and PPPDE1 C108S<br />

were determind by in vitro and in vivo de-ubiquitination assay.<br />

Co-immunoprecipitation (Co-IP) was used in combination with<br />

Mass Spectrometry and Western blot to identify and validation<br />

protein binding partners of PPPDE1. Colony formation<br />

assay and subcutaneous tumor model were used to evaluate<br />

the clonogenic and tumorigenic ability of HCC cells. Results:<br />

Purified PPPDE1 protein cleaved K48-linked and K63-linked<br />

Tetra-Ubiquitin into monoubiquitin in vitro. Transfected PPPDE1<br />

also markedly reduced the K48-linked and K63-linked protein<br />

ubiqutination level in cultured cells. These results showed that<br />

PPPDE1 is a novel DUB and the PPPDE protein family is a new<br />

DUBs family with cysteine proteases activity. The C108S substitution<br />

of PPPDE1 C108S destroyed the deubiquitination activity<br />

of PPPDE1. Moreover, PPPDE1 C108S lost the ability to promote<br />

the clonogenic growth of HCC cell lines compared to wild<br />

type PPPDE1, which suggested that PPPDE1 promote the HCC<br />

development mainly through its DUB activity. We also found<br />

that PPPDE1 stabilized a 25kd N termal MDM2 protein fragment<br />

(MDM2 p25), a degradation fragment from MDM2-FL<br />

(full length) and MDM2 p60 (60kd) protein, by binding and<br />

deubiquitinated this protein fragment. When over-expressed<br />

in HCC cell lines, PPPDE1 significantly promoted the degradation<br />

of p53 through a MDM2 p25 dependent manner and<br />

down regulated the protein level of BAX, a downstream effector<br />

of p53-induced apoptosis. Reversely, knockdown of PPPDE1<br />

in HCC cell lines considerably increased the p53 and BAX<br />

protein level while greatly down-regulated the MDM2 p25<br />

level. Most importantly, PPPDE1 silencing with lentiviral shRNA<br />

tremendously inhibited the in vivo growth of subcutaneous<br />

xenografts tumor of HCC cell lines in nude mice. Conclusion:<br />

PPPDE1 is a novel DUB that can promote the development of<br />

HCC by stabilizing a 25kd N terminal MDM2 fragment and<br />

subsequently suppressing the p53 pathway.<br />

Disclosures:<br />

Lai Wei - Advisory Committees or Review Panels: Gilead, AbbVie; Grant/<br />

Research Support: BMS<br />

The following authors have nothing to disclose: Xing-Wang Xie, Yang-Jing Zhao,<br />

Xue-Yan Wang, Ran Fei, Heng-Hui Zhang, Wei-Jia Liao, Yu Wang, Hong-Song<br />

Chen<br />

1974<br />

Mevalonate pathway targets FoxM1 transcription factor<br />

via protein geranylgeranylation in human hepatocellular<br />

carcinoma<br />

Satoshi Ogura, Yuichi Yoshida, Mayumi Egawa, Tomohide Kurahashi,<br />

Kunimaro Furuta, Shinichi Kiso, Yoshihiro Kamada, Tetsuo<br />

Takehara; Department of Gastroenterology and Hepatology,<br />

Osaka University, Suita, Osaka, Japan<br />

Background & Aims: Mevalonate (MV) pathway, which is<br />

required for cholesterol biosynthesis, has been implicated in<br />

the pathogenesis of hepatocellular carcinoma (HCC). Inhibition<br />

of HMG-CoA reductase (HMGCR), a rate limiting enzyme for<br />

MV pathway, has been also reported to reduce incidence of<br />

HCC in animal and clinical <strong>studies</strong>. The Forkhead box M1<br />

(FoxM1) is a proliferation-specific transcription factor and is<br />

over-expressed in a variety of cancers, including HCC. In addition,<br />

loss of FoxM1 is shown to reduce hepatocarcinogenesis<br />

in mouse models. However, to date, it remains unclear whether<br />

FoxM1 is involved in MV pathway of HCC. In this study, we<br />

aimed to elucidate this issue using in vitro culture systems.<br />

Method: Human hepatoma cell lines HepG2, Huh7, and HLF,<br />

were treated with chemical inhibitors of enzymes in MV pathway<br />

for 24 hours and the effect of these inhibitors on FoxM1<br />

protein expression was examined by Western blot analysis.<br />

Results: Administration of statin (pitavastatin), HMGCR inhibitor,<br />

induced an about 50 % reduction of FoxM1 expression<br />

in hepatoma cells (p


1170A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

1975<br />

Retrospective cohort study for liver carcinogenesis prediction<br />

among HBs antigen negative and anti-HCV antibody<br />

negative hepatitis patients<br />

Tomoko Aoki 2,1 , Hiroko Iijima 1 , Takashi Nishimura 1 , Chikage<br />

Nakano 1 , Tomoyuki Takashima 1 , Nobuhiro Aizawa 1 , Naoto<br />

Ikeda 1 , Hironori Tanaka 1 , Yoshinori Iwata 1 , Hirayuki Enomoto 1 ,<br />

Masaki Saito 1 , Shuhei Nishiguchi 1 ; 1 Hyogo College of Medicine,<br />

Nishinomiya city, Hyogo prifecture, Japan; 2 Internal medicine,<br />

YOKA HOSPITAL, Yoka 1878-1, Japan<br />

BACKGROUND & AIMS:Recently, we can use new ultrasonographic<br />

methods for non-invasive diagnosis of liver fibrosis.<br />

The aim for this study is to reveal the risk of liver carcinogenesis<br />

among HBs antigen negative and anti-HCV negative<br />

(nonBnonC) cases using Virtual Touch Quantification (VTQ).<br />

METHODS:Our research model was a retrospective cohort<br />

study. We continued to follow up 910 patients for 45.2<br />

months: 23 patients developed into hepatocellular carcinoma<br />

(HCC), we named them “carcinogenesis group”. The other<br />

887 patients did never develop HCC, we named them “carcinogenesis-free<br />

group”. We examined factors that contribute<br />

to carcinogenesis by Cox regression analysis. RESULT:The univariate<br />

analysis between two groups: there were significant difference<br />

in age, sex, platelet counts, Hyaluronic Acid (HA), FIB4<br />

index, VTQ, PT-INR, serum albumin, fasting plasma glucose<br />

(FPG), and total cholesterol(p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1171A<br />

expression of keratin 19(K19), epithelial cell adhesion molecule(EpCAM),<br />

matrix metalloproteinase 2/9(MMP2/9) and<br />

CXC chemokine receptor 4(CXCR4) were also down-regulated<br />

in Bel 7402 cells; Migration and invasion, expression of K19,<br />

EpCAM, MMP2/9 and CXCR4 were significant enhanced<br />

when HLE cells(non-AFP-producing) were transfected with AFP<br />

expressed vector. Conclusions: AFP plays a critical role in promoting<br />

metastasis of HCC; AFP is significantly promote HCC<br />

cell invasion and metastasis via up-regulated expression of<br />

metastasis-related proteins. Thus, AFP may be used as a novel<br />

therapeutic target for treating HCC patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Mingyue Zhu, Junli Guo, Wei<br />

Li, Mengsen Li<br />

1978<br />

Osteopontin-c isoform promotes a mesenchymal phenotype<br />

in human cholangiocarcinoma cells<br />

Marco A. Briones-Orta 1 , Jason D. Coombes 1 , Massimiliano Mellone<br />

7 , Rossella Rispoli 5 , Paul P. Manka 1,3 , Rasha Younis 1 , Naoto<br />

Kitamura 1 , Shannon S. Glaser 6,8 , Gianfranco Alpini 6,8 , Alberto<br />

Quaglia 9 , Roger Williams 1 , Salvatore Papa 4 , Ali Canbay 3 , Wing-<br />

Kin Syn 1,2 ; 1 Regeneration and Repair, The Institute of Hepatology,<br />

London, United Kingdom; 2 Liver Unit, Barts Health NHS Trust, London,<br />

United Kingdom; 3 Gastroenterology and Hepatology, University<br />

Hospital Essen, Essen, Germany; 4 Cell Signaling, The Institute<br />

of Hepatology, London, United Kingdom; 5 Weatherall Institute of<br />

Molecular Medicine, John Radcliffe Hospital, Oxford, United Kingdom;<br />

6 Research, Central Texas Veterans Health Care Systems,<br />

Temple, TX; 7 Experimental Pathology Group, Cancer Sciences<br />

Unit, Southampton General Hospital, Southampton, United Kingdom;<br />

8 Scott & White Digestive Disease Research Center, Texas<br />

A&M Health Science Center College of Medicine, Temple, TX;<br />

9 Liver Pathology, Kings College London, London, United Kingdom<br />

Introduction: Osteopontin (OPN) is upregulated in tissue fibrosis<br />

and cancer. The OPN gene generates 3 isoforms (a, b and<br />

c) by alternative splicing, which can induce differing cellular<br />

responses in tumorigenesis. Although these isoforms have been<br />

investigated in breast, prostate, and ovarian cancers, <strong>studies</strong> in<br />

cholangiocarcinoma (CCA) have been limited to OPN-a (total<br />

OPN). We hypothesized that OPN isoforms are overexpressed<br />

in CCA, and the pattern of isoform overexpression is associated<br />

with either a more or less aggressive phenotype. Crosstalk<br />

between OPN and TGF-b signaling was evaluated as both<br />

are key modulators of cancer progression. Methods: HuCCT1,<br />

SG231 and CCLP1, human intrahepatic cholangiocarcinoma<br />

cell lines were used. Plasmids for each OPN isoform were used<br />

for overexpression. Expression of OPN-a, b, c, and epithelial-mesenchymal<br />

transition (EMT) markers (vimentin, aSMA,<br />

E-cadherin) were evaluated by qRT-PCR. Global gene expression<br />

changes were examined by microarray using SG231.<br />

OPN isoform overexpression and their effects on the components<br />

of TGF-b pathway were evaluated by immunoblots. To<br />

evaluate the in vivo significance of the individual isoforms, a<br />

xenograft model was used. FFPE human cholangiocarcinoma<br />

and healthy liver sections were stained for total OPN. Results:<br />

Total OPN was upregulated in cholangiocarcinoma. CCLP1<br />

cells expressed the highest levels of total OPN and exhibited<br />

the most mesenchymal phenotype (high vimentin and low<br />

E-cadherin). In contrast, HuCCT1 expressed the lowest amount<br />

of OPN and exhibited an epithelial phenotype (high E-cadherin<br />

and low vimentin). In all 3 cell lines, OPN-a, and b mRNA<br />

were more abundant than OPN-c (~10 fold). Overexpression<br />

of OPN-c led to the greatest increase in mesenchymal genes<br />

and proteins (vimentin and aSMA), and was associated with<br />

the highest level of TGF-β signalling as shown by Smad2/3<br />

phosphorylation. Microarray confirmed that overexpression<br />

of OPN-c induced the largest alterations in gene expression<br />

profile, and an enrichment analysis revealed that OPN modulated<br />

genes which regulate phosphoprotein and acetylation. In<br />

the xenograft model, overexpression of all OPN isoforms was<br />

associated with greater rates of tumour growth than control,<br />

but OPN-c had the greatest rate. Conclusions: OPN expression<br />

correlates with the degree of EMT (marker of aggressiveness)<br />

in human cholangiocarcinoma. OPN-c is associated with the<br />

most mesenchymal phenotype and TGF-β pathway activation.<br />

OPN-c also induces greater changes in the transcriptome and<br />

supports greater rates of tumor growth than OPN-a, and -b.<br />

High OPN-c expression may be a predictor of worse clinical<br />

outcomes.<br />

Disclosures:<br />

The following authors have nothing to disclose: Marco A. Briones-Orta, Jason D.<br />

Coombes, Massimiliano Mellone, Rossella Rispoli, Paul P. Manka, Rasha Younis,<br />

Naoto Kitamura, Shannon S. Glaser, Gianfranco Alpini, Alberto Quaglia, Roger<br />

Williams, Salvatore Papa, Ali Canbay, Wing-Kin Syn<br />

1979<br />

Glypican 3 (GPC3)-CD81 axis in regulation Hippo pathway<br />

in hepatocytes and hepatocellular carcinoma (HCC)<br />

Yuhua Xue, Kelly Koral, William C. Bowen, Anne Orr, Meagan<br />

Haynes, Wendy M. Mars, George K. Michalopoulos; Pathology,<br />

University of Pittsburgh, Pittsburgh, PA<br />

GPC3 is over-expressed in HCC. Mutations of GPC3 in humans<br />

are the cause of Simpson-Golabi-Behmel syndrome, associated<br />

to organomegaly. Previous <strong>studies</strong> have shown that GPC3 is<br />

associated with inhibition of hepatocyte growth and GPC3<br />

transgenic mice with targeted expression in hepatocytes have<br />

decreased liver regeneration. We have also shown that GPC3<br />

binds to the membrane tetraspanin CD81, which is one of the<br />

portals of entry of hepatitis C virus (HCV). Studies elsewhere<br />

have shown that activation of CD81 is associated with phosphorylation<br />

of Ezrin. The latter has been shown to regulate<br />

the Hippo pathway, responsible for regulating nuclear levels<br />

of the protein known as Yes-associated protein (Yap). Expression<br />

of Yap in the nucleus relates to activation of hepatocyte<br />

growth. Mice over-expressing GPC3 have decreased levels of<br />

nuclear Yap. Since we have demonstrated that GPC3 binds<br />

to CD81, the purpose of this study is to explore the role of<br />

GPC3 in regulation of Yap via CD81 and Hippo pathway and<br />

to investigate the possible mechanisms. We found that treatment<br />

of primary rat hepatocytes with CD81 agonist antibody<br />

led to higher levels of phosphorylated Ezrin and lower Hippo<br />

activity. Partial hepatectomy (PHx) exhibited that loss of GPC3<br />

antagonistic modulation during early liver regeneration, CD81<br />

activated Ezrin and down regulated Hippo activity in vivo. Rat<br />

hepatoma cell lines JM1 and JM2 had dramatically down regulated<br />

CD81 expression and Hippo activity, and up regulated<br />

Ezrin activity in vitro. Re-gain of CD81 expression in JM2 rat<br />

hepatoma cells up regulated Hippo activity while down regulated<br />

Ezrin activity in vitro. Human HCC tissue array revealed<br />

absence of CD81 in most HCCs but present in all the normal<br />

tissue controls, while GPC3 was present in most HCCs and all<br />

the normal tissue controls. Treatment with HCV envelope protein<br />

E2 led to more Hippo activity and decrease in nuclear Yap<br />

in HepaRG cells and primary human hepatocytes. Conclusions:<br />

The CD81-phosphoEzrin signaling enhances nuclear Yap levels.<br />

This pathway in normal hepatocytes is inhibited by GPC3.<br />

In human hepatocytes, GPC3 and the E2 protein of HCV also<br />

inhibit the CD81-phosphoEzrin effect on Hippo pathway and<br />

cause a decrease in nuclear Yap. Since most HCC lack expression<br />

of CD81, they are likely to escape from antagonistic mod-


1172A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ulation of over-expressed GPC3. HCC may also have a growth<br />

advantage over normal hepatocytes in HCV infection by missing<br />

CD81, thus being able to undergo clonal expansion, facilitating<br />

HCC early development and growth. Persistent HCV<br />

infection strengthens this expansion process and contributes<br />

carcinogenesis by E2 protein interacting with CD81.<br />

Disclosures:<br />

George K. Michalopoulos - Consulting: Vital Therapies<br />

The following authors have nothing to disclose: Yuhua Xue, Kelly Koral, William<br />

C. Bowen, Anne Orr, Meagan Haynes, Wendy M. Mars<br />

gene that acts as a driving force for the amplification at 3q26<br />

in HCC and that the oncoprotein EVI1 suppresses TGF-β-mediated<br />

growth inhibition of HCC cells.<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

The following authors have nothing to disclose: Kohichiroh Yasui, Atsushi<br />

Umemura, Taichiro Nishikawa, Kanji Yamaguchi, Michihisa Moriguchi, Yoshio<br />

Sumida, Hironori Mitsuyoshi, Shinji Tanaka, Shigeki Arii<br />

1980<br />

EVI1 is a target for gene amplification at 3q26 and<br />

suppresses growth inhibition by transforming growth<br />

factor-β in hepatocellular carcinoma<br />

Kohichiroh Yasui 1 , Atsushi Umemura 1 , Taichiro Nishikawa 1 , Kanji<br />

Yamaguchi 1 , Michihisa Moriguchi 1 , Yoshio Sumida 1 , Hironori Mitsuyoshi<br />

1 , Shinji Tanaka 2 , Shigeki Arii 3,4 , Yoshito Itoh 1 ; 1 Department<br />

of Molecular Gastroenterology and Hepatology, Kyoto<br />

Prefectural University of Medicine, Kyoto, Japan; 2 Department<br />

of Molecular Oncology, Tokyo Medical and Dental University,<br />

Tokyo, Japan; 3 Hamamatsu Rosai Hospital, Hamamatsu, Japan;<br />

4 Department of Hepato-Biliary-Pancreatic Surgery, Tokyo Medical<br />

and Dental University, Tokyo, Japan<br />

Background: Amplification of DNA in certain regions of chromosomes<br />

plays a crucial role in the development and progression<br />

of human malignancies, specifically when proto-oncogenic<br />

target genes within those amplicons are overexpressed. A common<br />

criterion for designation of a gene as a putative target<br />

of amplification is that gene amplification leads to its overexpression.<br />

The EVI1 (ecotropic viral integration site 1) gene<br />

codes for a zinc finger transcriptional factor that is involved in<br />

leukemic transformation of hematopoietic cells. EVI1 is one of<br />

the most aggressive oncogenes associated with myeloid leukemia.<br />

EVI1 has been reported to suppress transforming growth<br />

factor (TGF)-β signaling by inhibiting Smad3 in leukemic cells.<br />

However, little is known about its relevance for hepatocellular<br />

carcinoma (HCC). Aim: To identify a novel gene that acts as a<br />

driving force for gene amplification in HCC and is involved in<br />

the development and progression of HCC. Methods: We investigated<br />

DNA copy number aberrations in 20 HCC cell lines<br />

using a high-density oligonucleotide microarray (GeneChip<br />

Mapping 100K and 250K arrays, Affymetrix). DNA copy<br />

number and expression of EVI1 were determined by using fluorescence<br />

in situ hybridization (FISH) and quantitative PCR in<br />

HCC cell lines and primary HCCs. Knockdown experiments of<br />

EVI1 were performed. Results: We found that a novel amplification<br />

at the chromosomal region 3q26 occurs in the HCC cell<br />

line, JHH-1, and that MDS1 (myelodysplastic syndrome 1) and<br />

EVI1 complex locus (MECOM ), which lies within the 3q26<br />

region, was amplified. Quantitative PCR analysis of the three<br />

transcripts transcribed from MECOM indicated that only EVI1,<br />

but not the fusion transcript MDS1-EVI1 or MDS1, was overexpressed<br />

in JHH-1 cells and was significantly up-regulated in 22<br />

(61%) of 36 primary HCC tumors when compared with their<br />

non-tumorous counterparts. A copy number gain of EVI1 was<br />

observed in 24 (36%) of 66 primary HCC tumors. High EVI1<br />

expression was significantly associated with larger tumor size<br />

and higher level of des-γ-carboxy prothrombin (DCP), a tumor<br />

marker for HCC. Knockdown of EVI1 resulted in increased<br />

induction of the cyclin-dependent kinase inhibitor p15 INK4B by<br />

TGF-β and decreased expression of c-Myc, cyclin D1, and<br />

phosphorylated Rb in TGF-β-treated cells. Consequently, knockdown<br />

of EVI1 led to reduced DNA synthesis and cell viability.<br />

Conclusions: Our results suggest that EVI1 is a probable target<br />

1981<br />

Hyperinsulinemia, not hyperglycemia accelerates the<br />

progression of hepatocellular carcinoma in neonatal<br />

streptozotocin induced mouse model<br />

Koichi Tsuneyama, Ryosuke Bessho, Masahiro Miki, Hayato Baba,<br />

Tetsuyuki Takahashi, Hirohisa Ogawa, Hisanori Uehara; Department<br />

of Pathology and Laboratory Medicine, Institute of Biomedical<br />

Sciences, Tokushima University Graduate School, Tokushima,<br />

Japan<br />

Purpose: Diabetes mellitus (DM) is a well-known risk factor<br />

for hepatocellular carcinoma (HCC); however, the underlying<br />

mechanisms are not well understood. We have previously<br />

reported that neonatal streptozotocin (STZ) treatment in DIAR<br />

mice (DIAR-nSTZ) causes type 1 diabetes and subsequent HCC<br />

without special diet. In the following study, to examine the<br />

relation between hyperglycemia and HCC, we normalized the<br />

blood glucose level by applying excess insulin in DIAR-nSTZ<br />

mice. Although diabetic features were completely improved,<br />

frequency of HCC did not improve in 12 weeks of age. We<br />

hypothesized that hyperinsulinemia, not hyperglycemia have<br />

a close association to carcinogenesis in this model. To clarify<br />

this hypothesis, we examined neonatal STZ treatment in IV CS<br />

mice, which showed quick normalization of blood glucose level<br />

after STZ treatment by rapid regeneration of islet cells. Methods:<br />

Newborn male IV-CS mice were prepared at institute of<br />

animal reproduction (Ibaraki, Japan). STZ (60 mg/g) was subcutaneously<br />

injected into the IV CS-STZ treated group (IV-CSnSTZ<br />

mice, N13), whereas physiologic solution was injected<br />

into the control group (IV-CS-control mice, N=13) at 1.5 days<br />

after birth. All mice were fed a normal diet, and physiologically<br />

and histopathologically assessed at 8 (N=2), 12 (N=3) and<br />

16 (N-8) weeks of age. Blood was taken once in month and<br />

measured glucose and insulin level in all mice. Results: Blood<br />

glucose level of IV CS-nSTZ mice showed slightly increase till<br />

5 week of age, but it was normalized after 6 weeks of age. In<br />

contrast, blood insulin level showed higher titer after 6 weeks<br />

of age. Even at 8 weeks of age, one out of two mice showed<br />

hepatic nodule in a diameter of 2mm. Histopathologically, it<br />

showed mild structural and nuclear atypia showing dysplastic<br />

nodule. At 12 weeks of age, one out of three mice contained<br />

2 hepatic nodules of dysplastic nodule. At 16 weeks of age,<br />

six out of eight mice showed one or more hepatic nodules.<br />

Histologically, well differentiated hepatocellular carcinoma<br />

as well as dysplastic nodules were observed. In contrast, no<br />

hepatic nodules were recognized in IV CS-control mice. Conclusions:<br />

IV CS mice showed hepatocarcinogenesis after neonatal<br />

STZ treatment. Differently from other strains including ddY<br />

and DIAR, blood glucose level did not show marked increase<br />

after STZ treatment, and it was normalized after 6 weeks of<br />

age. However, frequency of HCC showed similar tendency<br />

of DIAR-nSTZ mice. In addition to our previous data of insulin<br />

treatment of DIAR-nSTZ mice, we strongly hypothesized that


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1173A<br />

hyperinsulinaemia rather than hyperglycemia can accelerate<br />

the progression of HCC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Koichi Tsuneyama, Ryosuke<br />

Bessho, Masahiro Miki, Hayato Baba, Tetsuyuki Takahashi, Hirohisa Ogawa,<br />

Hisanori Uehara<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yoshiki Onishi, Tomohide Tatsumi,<br />

Satoshi Aono, Seiichi Tawara, Akira Nishio, Atsuo Takigawa, Takahiro<br />

Suda, Tadashi Kegasawa, Shigeki Suemura, Minoru Shigekawa, Ryotaro Sakamori,<br />

Naoki Hiramatsu<br />

1982<br />

Branched chain amino acids promote sorafenib-mediated<br />

apoptosis in hepatocellular carcinoma via<br />

down-regulation of Mcl-1<br />

Yoshiki Onishi, Tomohide Tatsumi, Satoshi Aono, Seiichi Tawara,<br />

Akira Nishio, Atsuo Takigawa, Takahiro Suda, Tadashi Kegasawa,<br />

Shigeki Suemura, Minoru Shigekawa, Hayato Hikita,<br />

Ryotaro Sakamori, Naoki Hiramatsu, Tetsuo Takehara; Department<br />

of Gastroenterology and Hepatology, Osaka University Gradute<br />

School of Medicine, Suita, Japan<br />

Background&aim: Sorafenib is an only effective chemotherapeutic<br />

agent in the treatment of hepatocellular carcinoma<br />

(HCC). However, the survival benefit is limited. Branched<br />

chain amino acids (BCAAs) promote albumin synthesis through<br />

the activation of mTOR pathway. BCAAs treatment has been<br />

reported to reduce the incidence of HCC in patients with liver<br />

cirrhosis and to reduce the recurrence of HCC after curative<br />

treatment. We previously reported a retrospective study that<br />

the combined use of sorafenib and BCAAs prolonged the<br />

overall survival in HCC patients. In this study, we examined<br />

the mechanism of anticancer effect of the combination therapy.<br />

Method: Human HCC cell lines, HepG2, PLC/PRF/5 and<br />

Huh7, were treated with sorafenib and BCAAs. Cell viabiliy<br />

was evaluated by WST assay. To evaluate apoptosis of HCC<br />

cells, we measured caspase3/7 activities in the culture supernatants<br />

and Annexin-V positive cells by flow cytometry. Growth<br />

signals and regulatory proteins of apoptosis were analyzed<br />

by western blotting. mRNA levels were evaluated by using<br />

quantitative RT-PCR. Result: The combined use of sorafenib and<br />

BCAAs significantly decreases in cell viability of all treated<br />

HCC cells. Western blotting analysis revealed that addition<br />

of BCAAs enhanced inhibition of ERK pathway in sorafenibtreated<br />

HCC cells. Sorafenib treatment induced Akt phosphorylation<br />

in HCC cells as previously reported. However, addition<br />

of BCAAs antagonized the Akt phosphorylation. The combined<br />

use of sorafenib and BCAAs resulted in significant increase<br />

in caspase 3/7 activities and Annexin-V positive cell rates in<br />

HCC cells compared with sorafenib alone. We next examined<br />

pro-and anti-apoptotic proteins in HCC cells treated with<br />

sorafenib and BCAAs. Addition of BCAAs enhanced reduction<br />

of anti-apoptotic protein Mcl-1 expression in sorafeib-treated<br />

HCC cells. In contrast, both pro-apoptotic proteins Bak/Bax<br />

and anti-apoptotic protein Bcl-xL did not changed. mRNA<br />

expression levels of Mcl-1 were significantly lower in HCC cells<br />

treated by sorafenib with BCAAs than those without BCAAs.<br />

Furthermore, we evaluated Mcl-1 degradation in the HCC cells<br />

by using cycloheximide, a protein synthesis inhibitor. After inhibition<br />

of protein synthesis, Mcl-1 levels were decreased more in<br />

sorafenib-treated cells with BCAAs than those without BCAAs.<br />

These results indicate that BCAAs reduce Mcl-1 expression at<br />

both transcriptional and post-transcriptional levels in sorafenibtreated<br />

HCC cells. Conclusion: BCAAs enhanced sorafenib-mediated<br />

apoptosis in HCC cells via down regulation of Mcl-1.<br />

These results suggest that the combined use of sorafenib and<br />

BCAAs is effective in the treatment of advanced HCC patients.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

1983<br />

DNA methylome and cancer-specific expression change<br />

of clustered miRNAs in non-B non-C hepatocellular carcinoma<br />

Takeshi Matsui 2,1 , Masanori Nojima 8 , Etsuko Iio 2 , Akihiro Tamori 3 ,<br />

Shoji Kubo 4 , Ken Shirabe 5 , Koichi Kimura 9 , Mitsuo Shimada 9 ,<br />

Tohru Utsunomiya 6 , Yasuteru Kondo 7 , Yasuhito Tanaka 2 ; 1 Center<br />

for Gastroenterology, Teine-Keijinkai Hospital, Sapporo, Japan;<br />

2 Department of Virology unit, Nagoya City University Graduate<br />

School of Medical Sciences, Nagoya, Japan; 3 Department of<br />

Hepatology, Osaka City University Graduate School of Medicine,<br />

Osaka, Japan; 4 Department of Hepato-Biliary-Pancreatic Surgery,<br />

Graduate School of Medicine, Osaka City University, Osaka,<br />

Japan; 5 Department of Surgery and Science, Graduate School<br />

of Medical Sciences, Kyushu University, Fukuoka, Japan; 6 Oita<br />

prefectual hospital, Oita, Japan; 7 Division of Gastroenterology,<br />

Tohoku University Hospital, Sendai, Japan; 8 Institute of Medical<br />

Science Hospital, the University of Tokyo, Tokyo, Japan; 9 Department<br />

of Surgery, the University of Tokushima, Tokushima, Japan<br />

[Background] While HBV and HCV infection have been suppressed<br />

with development of public health measures and<br />

medical treatment, non-B non-C hepatocellular carcinoma<br />

(NBNC-HCC) is considered increasing based on increase of<br />

nonalcoholic fatty liver disease (NAFLD) in Japan. Molecular<br />

biological mechanism should be pursued against this disease<br />

structure change. [Aim] To explore critical changes of DNA<br />

methylation in micro RNA (miRNA) coding regions and its<br />

expression in tumorigenesis of NBNC-HCC. [Methods] Infinium<br />

HumanMethylation450 BeadChip Kit (Illumina) and micro<br />

3D-Gene miRNA Oligo Chip (Toray) were applied in 26 pairs<br />

of tumor and non-tumor background tissue samples. Pooled<br />

analysis using the data from Gene Expression Omnibus (GEO,<br />

NCBI) was also conducted to strengthen generalizability, and<br />

to seek common DNA methylation changes across infection<br />

status (HBV, HCV or NBNC). Gene ontology analysis were<br />

applied for gene clusters defined by clustering and other statistical<br />

characteristics. [Results and plan] Hierarchical clustering<br />

of the DNA methylome status demonstrates clear discrimination<br />

between tumor and non-tumor samples. While approx. 20% of<br />

CpG islands in promoters were dominantly hypermethylated<br />

in the tumor, other regions were globally hypomethylated in<br />

the tumor. The mean difference in methylation levels of total<br />

probes was -10.5% in non-CpG islands, and 0.68% in CpG<br />

islands (tumor vs. non-tumor). It suggests that the change of<br />

methylation status was totally different between CpG and non-<br />

CpG islands. Next, to investigate the molecular biological significance<br />

of DNA hypomethylation in the tumor, we determine<br />

the characteristics of hypomethylated sites, and then found several<br />

miRNA clusters were markedly hypomethylated (-30% to<br />

-40%) in tumor tissues (e.g. miR-493 in 14q32.2, miR-154 in<br />

14q32.31, miR-518F in 19q13.41, miR-888 in Xq27.3). Average<br />

methylation level in miRNA coding region were lower than<br />

other regions. In detail, 52.8% of total miRNA probes were significantly<br />

hypomethylated, and mean difference in methylation<br />

levels were -14.4% in non-CpG islands, and 0.14% in CpG<br />

islands. Consistent with the hypomethylation tendency, average<br />

miRNA expression level was 28.0% up-regulated. In particular,<br />

miRNA expression significantly around 80% to 100%


1174A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

up-regurated in the above hypomethylated miRNA clusters.<br />

[Conclusions] Several major miRNA clusters were markedly<br />

hypomethylated in tumor tissues, and expression of miRNAs<br />

coded in clustered-regions were broadly up-regulated. We will<br />

present the results of further advanced analyses including the<br />

influence of hypermethylation for target genes, and integration<br />

of clinical information.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Takeshi Matsui, Masanori<br />

Nojima, Etsuko Iio, Akihiro Tamori, Shoji Kubo, Ken Shirabe, Koichi Kimura,<br />

Mitsuo Shimada, Tohru Utsunomiya, Yasuteru Kondo<br />

1984<br />

The analyses of oxidative DNA repair genes and hepatocarcinogenesis<br />

in patients with chronic hepatitis C<br />

Koji Miyanishi, Masayoshi Kobune, Shingo Tanaka, Toshifumi<br />

Hoki, Tsutomu Sato, Yasushi Sato, Rishu Takimoto, Junji Kato;<br />

Medical Oncology and Hematology, Sapporo Medical University,<br />

Sapporo, Japan<br />

Background and aim: Recent <strong>studies</strong> have shown that excess<br />

hepatic iron accumulation contributes to liver injury in patients<br />

with chronic hepatitis C (CHC). Free iron in the liver is believed<br />

to facilitate the formation of reactive oxygen species (ROS),<br />

including hydroxyl (OH) radicals, which cause oxidative damage<br />

of numerous cellular components such as nucleic acids.<br />

The OH radical is known to generate promutagenic bases such<br />

as 8-hydroxy-2-deoxyguanosine (8-OHdG), which has been<br />

implicated in DNA mutagenesis and carcinogenesis. 8-OHdG<br />

lesions may be promptly repaired by human 8-oxo-guanine<br />

DNA glycosylase (hOGG1) and human MUTY homolog<br />

(MUTYH). Therefore, it is plausible that the development of<br />

hepatocellular carcinoma (HCC) from CHC is determined by<br />

the balance between DNA damage and repair. In this study,<br />

we conducted analyses of single nucleotide polymorphisms<br />

(SNPs) of hOGG1 and MUTYH genes in CHC patients with<br />

or without development of HCC. Further, we examined the<br />

efficacy of N-acetyl cysteine (NAC) as a treatment to prevent<br />

the development of HCC using MUYTH null mice. Methods:<br />

Genomic DNA and total RNA were extracted from peripheral-blood<br />

mononuclear cells obtained from 63 CHC patients with<br />

(HCC group, n=39) or without HCC (non-HCC group, n=24).<br />

We analyzed the SNPs of hOGG1 and MUTYH by target<br />

resequence using next generation sequencer (ION PGM) and<br />

mRNA of hOGG1 and MUTYH by Taqman RT-PCR. The HCC<br />

incidence rate was calculated based on the period between<br />

the diagnosis of CHC and the appearance of HCC, using the<br />

Kaplan-Meier technique. In vivo study, male C57BL/6 MUTYH<br />

null mice were fed an excess-iron diet or control diet with or<br />

without NAC. Results: The frequency of SNPs in hOGG1 did<br />

not differ between HCC group and non-HCC group. In one<br />

of MUTYH SNPs (rs3219487), the frequency of minor allele<br />

carrier in the HCC group significantly higher than that in HCC<br />

group. In addition, the expression of MUTYH mRNA in minor<br />

homo allele were lower than that in major homo allele. Multivariate<br />

Cox regression analysis indicated that age, liver fibrosis<br />

factor and the minor allele carrier independently associated<br />

with hepatocarcinogenesis. Hepatocarcinogenesis rate in<br />

MUTYH null mice which were fed an excess-iron diet were significantly<br />

higher than that in wild type C57BL/6 or in null mice<br />

which were fed a control diet. NAC administration decreased<br />

the development of HCC in null mice. Conclusions: These results<br />

indicated that the minor allele of this SNP in MUTYH could be<br />

a significant risk factor of liver carcinogenesis and could be<br />

a predictive marker of HCC development in CHC patients.<br />

Further, NAC may be useful for prevention of HCC in such<br />

patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Koji Miyanishi, Masayoshi<br />

Kobune, Shingo Tanaka, Toshifumi Hoki, Tsutomu Sato, Yasushi Sato, Rishu<br />

Takimoto, Junji Kato<br />

1985<br />

Continuous hepatocyte apoptosis accelerates diethylnitrosamine-induced<br />

tumor development in the liver<br />

Yasutoshi Nozaki, Hayato Hikita, Satoshi Tanaka, Sadatsugu<br />

Sakane, Yugo Kai, Yuki Makino, Tasuku Nakabori, Yoshinobu<br />

Saito, Ryotaro Sakamori, Naoki Hiramatsu, Tomohide Tatsumi,<br />

Tetsuo Takehara; Osaka University Graduate School of Medicine,<br />

Suita, Osaka, Japan<br />

Background and Aim: Apoptosis serves as an important mechanism<br />

for removing DNA damaged-cells and is considered to<br />

inhibit carcinogenesis. On the other hand, hepatocyte apoptosis<br />

is a key feature of chronic liver disease including viral hepatitis<br />

and steatohepatitis, which are well-known risk factors for<br />

HCC. Thus, it remains unclear whether the apoptosis inhibition<br />

accelerates or decelerates liver tumor development. The present<br />

study examined the impact of continuous hepatocyte apoptosis<br />

and inhibition of apoptosis on liver tumor development.<br />

Methods: We used male hepatocyte-specific knockout (KO)<br />

mice of Mcl-1, one of anti-apoptotic proteins, as a model of<br />

apoptosis-prone liver, and hepatocyte-specific single or double<br />

KO mice of Bak and Bax, proapoptotic proteins, as a model<br />

of apoptosis-resistant liver . To induce liver tumors those KO<br />

mice were intraperitoneally administered 20 mg/kg diethylnitrosamine<br />

(DEN) at the age of 2 weeks. Results: Mcl-1 KO<br />

mice spontaneous hepatocyte apoptosis as evidenced by HE<br />

staining of the liver sections, increasing serum ALT levels and<br />

serum caspase-3/7 activities. The number of TUNEL positive<br />

hepatocytes also increased in Mcl-1 KO mice. No liver tumors<br />

were observed at 6 months in both Mcl-1 KO and wild-type<br />

(WT) mice without DEN. On the other hand, while only 11.8%<br />

(2/17) of WT mice treated with DEN developed macroscopic<br />

liver tumors in 6 months, 100% (7/7) of Mcl-1 KO littermates<br />

developed (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1175A<br />

The following authors have nothing to disclose: Yasutoshi Nozaki, Satoshi<br />

Tanaka, Sadatsugu Sakane, Yugo Kai, Yuki Makino, Tasuku Nakabori, Yoshinobu<br />

Saito, Ryotaro Sakamori, Naoki Hiramatsu, Tomohide Tatsumi<br />

1986<br />

FGF15 and FGF19 Induce Disparate FGFR4-Mediated<br />

Hepatocarcinogenicity In-Vitro and In Two Murine Models:<br />

Implications for Drug-Associated Carcinogenicity<br />

Risk Assessments<br />

Lei Ling, Mei Zhou, Darrin Lindhout, Jian Luo, Michael Chen,<br />

Hong Yang, Mark Humphrey, Van Phung, Kalyani Mondal, Hugo<br />

Matern, Marc Learned, Stephen Rossi, Alex M. DePaoli, Hui Tian;<br />

NGM Biopharmaceuticals, S. San Francisco, CA<br />

Background and Aims: FGF19 is a regulator of bile acid (BA)<br />

synthesis in humans. However, FGF19 produces hepatocellular<br />

carcinoma (HCC) in transgenic mice and is linked to increased<br />

risk of HCC post-resection recurrence in humans. FGF15 is the<br />

rodent ortholog of human FGF19 with 50% amino acid (AA)<br />

homology and used in rodents to assess the pharmacologic and<br />

caricongenic activity of FGF19. Both FGF15 and FGF19 are<br />

potent inhibitors of Cyp7a1-mediated BA synthesis in rodents<br />

but the carcinogenicity risk of FGF15 is not well characterized.<br />

The hepatocarcinogenicity of FGF15 and FGF19 was<br />

evaluated in leptin receptor-deficient (db/db) mice after 24<br />

wks or diet-induced obese (DIO) mice after 52 weeks of treatment.<br />

In vitro receptor interaction and activation <strong>studies</strong> were<br />

performed. Methods: FGF15, FGF19 or control was dosed<br />

by long-term transgene expression using 1 dose of adeno-associated<br />

virus (AAV)-mediated gene delivery. Liver tissue was<br />

examined for tumors and liver weight at 24wks post-dose (db/<br />

db) or 52wks post-dose (DIO). Gene expression of HCC-related<br />

and cell proliferation markers were measured in liver<br />

tissue. FGFR4-Klothoβ (FGFR4-KLB) receptor complex binding<br />

was assessed with SPR assays and transfected rat L6 cells were<br />

used to assess receptor activation. Results: FGF19 induced<br />

liver tumors in both models post-treatment at concentrations<br />

as low as 1 ng/ml (Figures 1a, 1b). In contrast, FGF15 failed<br />

to induce liver tumors at supraphysiologic concentrations and<br />

maintained normal liver weight and liver-body weight ratios.<br />

Liver tissue expression of Ki-67 (Figure 1c), α-fetoprotein, glypican-3,<br />

cyclin-a2, Ccnb1 and Ccnb2 were induced with FGF19<br />

but not FGF15. Reduced activation of FGFR4-KLB receptor complex<br />

was only seen with FGF15, as measured by decreased<br />

expression of STAT-3 target genes. These differences may be<br />

mediated by AA sequences unique to FGF15. Conclusions:<br />

Significant differences in hepatocellular proliferation and tumorigenesis<br />

were observed with FGF15 vs FGF19. These data<br />

demonstrate the potential challenges of assessing the risks of<br />

increased FGF19 levels in humans using FGF15 in rodent carcinogenicity<br />

models.<br />

Disclosures:<br />

Lei Ling - Employment: NGM Biopharmaceuticals, Inc.<br />

Mei Zhou - Stock Shareholder: NGM Biopharmaceuticals<br />

Darrin Lindhout - Patent Held/Filed: NGM Biopharmaceuticals; Stock Shareholder:<br />

NGM Biopharmaceuticals<br />

Jian Luo - Employment: NGM Biopharmaceuticals<br />

Hong Yang - Employment: NGM Biopharmaceuticals<br />

Mark Humphrey - Employment: NGM Biopharmaceuticals; Stock Shareholder:<br />

NGM Biopharmaceuticals<br />

Van Phung - Employment: NGM Biopharmaceuticals; Stock Shareholder: NGM<br />

Biopharmaceuticals<br />

Kalyani Mondal - Stock Shareholder: NGM BioPharmaceuticals<br />

Hugo Matern - Employment: NGM; Stock Shareholder: NGM<br />

Marc Learned - Employment: NGM Biopharmaceuticals, Inc.; Stock Shareholder:<br />

NGM Biopharmaceuticals, Inc.<br />

Stephen Rossi - Employment: NGM Biopharmaceuticals, Inc; Stock Shareholder:<br />

NGM Biopharmaceuticals, Gilead Sciences<br />

Alex M. DePaoli - Employment: NGM Biopharmaceuticals<br />

Hui Tian - Management Position: NGM Biopharma<br />

The following authors have nothing to disclose: Michael Chen<br />

1987<br />

Lymphotoxin–β regulated by NF–κB is a useful biomarker<br />

related to poor prognosis in hepatocellular carcinoma<br />

Takehisa Watanabe, Satomi Fujie, Youko Yoshimaru, Katsuya<br />

Nagaoka, Hiroko Setoyama, Kotaro Fukubayashi, Masakuni Tateyama,<br />

Hideaki Naoe, Jiro Fujimoto, Motohiko Tanaka, Yutaka<br />

Sasaki; Hepatology, Kumamoto Univ., Kunamoto, Japan<br />

The Lymphotoxin-β (LTβ) is a major pro-inflammatory cytokine<br />

that belongs to the TNF superfamily and mediates local<br />

inflammatory responses as a membrane-bound cytokine. The<br />

LTβ pathway is considered to be an important pathway in<br />

the hepato-carcinogenesis. We previously reported that the<br />

CCCTC-binding factor (CTCF) mediated higher order chromatin<br />

conformation of TNF/LT locus, including LTβ gene, is deeply<br />

involved in the regulation of LTβ expression in human HCC,<br />

by the global analysis with ChIP-chip and ChIP-seq. In this<br />

study, we attempted to reveal the molecular mechanism of LTβ<br />

regulation and the clinical characteristics of the patients with<br />

HCC expressing LTβ. Because chronic inflammation is one of<br />

the most crucial factors for hepato-carcinogenesis, we firstly<br />

focused on NF-κB, a major pro-inflammatory transcription factor,<br />

and investigated the influence of NF-κB on LTβ expression<br />

and higher order chromatin conformation, by using cell lines<br />

differed on the activity of NF-κB. As a result, LTβ was expressed<br />

continuously in the cells with constitutive active NF-κB in<br />

response to the stimulation with TNF. On the other hand, LTβ<br />

degraded rapidly in the cells without constitutive active NF-κB.<br />

These cells differ on the higher order chromatin conformation<br />

to control the interaction between the LTβ promoter and appropriate<br />

NF-κB responsive enhancer. These results indicated that<br />

NF-κB was involved in the higher order chromatin conformation<br />

and regulation of LTβ. Additionally, we found that LTβ<br />

could be quantified in the exosomes extracted from serum of<br />

the patient, whereas liver biopsy had been considered to be<br />

indispensable to describe how LTβ is expressed in hepatic lobules.<br />

To examine whether exosomal LTβ reflects LTβ expression<br />

in HCC tissue, we performed immunohistochemistry with LTβ<br />

antibodies in 24 samples from HCC patients and compare with<br />

the amount of exosomal LTβ. Consequently, high level exosomal<br />

LTβ reflected the high expression LTβ in the tumor tissues,<br />

not in the adjacent non-tumor tissues. Moreover, the patients<br />

with the high level exosomal LTβ exhibited significantly shorter<br />

survival time. Therefore, these results suggest that exosomal LTβ<br />

might be a useful biomarker related to poor prognosis in HCC.<br />

Disclosures:<br />

Yutaka Sasaki - Grant/Research Support: Chugai Pharmaceutical Co. JAPAN,<br />

MSD Co. JAPAN


1176A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Takehisa Watanabe, Satomi<br />

Fujie, Youko Yoshimaru, Katsuya Nagaoka, Hiroko Setoyama, Kotaro Fukubayashi,<br />

Masakuni Tateyama, Hideaki Naoe, Jiro Fujimoto, Motohiko Tanaka<br />

1988<br />

DDB2 loss protects mice from H-ras12V-driven HCCs<br />

Dragana Kopanja, Akshay Pandey, Shuo Huang, Damiano Fantini,<br />

Pradip Raychaudhuri; Biochemistry and Molecular Genetics,<br />

University of Illinois at Chicago, Chicago, IL<br />

DDB2 (Damaged DNA-binding protein (DDB)-2,) is a protein<br />

encoded by the nucleotide excision repair (NER) gene xeroderma<br />

pigmentosum group E (XPE). DDB2 is involved in early<br />

steps of NER and in DNA damage-induced apoptosis. Interestingly,<br />

DDB2 also acts as transcriptional repressor and regulates<br />

expression of antioxidant genes MnSOD and Catalase. Furthermore,<br />

DDB2 represses transcription of EMT-inducing factors<br />

Snail, VEGF and ZEB1, and its loss in high grade colon cancers<br />

promotes metastasis by promoting EMT. Moreover, DDB2<br />

-/- mice develop spontaneous tumors at higher frequency than<br />

wt mice, suggesting that it can act as a tumor suppressor. In<br />

this study, we investigated role of DDB2 in liver cancer. Since<br />

Ras-signaling pathway is found to be ubiquitously activated<br />

in human hepatocellular carcinoma (HCC) through epigenetic<br />

silencing of the Ras-regulators, we used mice that express<br />

H-ras12V under control of albumin promoter. Male mice of<br />

this strain develop liver tumors at age of 8-9 months with penetrance<br />

of almost 100%. We crossed H-ras12V mice with DDB2<br />

knockout mice (DDB2 -/-). Surprisingly, in that model of HCCs,<br />

DDB2 deletion protects mice from liver tumor development. We<br />

observed lower number of liver tumor nodules in DDB2 knockout<br />

H-ras12V mice compared to the DDB2 +/+ H-ras12V mice.<br />

Moreover, unlike in colon cancer, DDB2 loss does not induce<br />

metastasis of H-ras12V driven liver cancer. To understand why<br />

DDB2 -/- H-ras12V mice develop lower number of tumor nodules<br />

compared to DDB2 +/+ H-ras12V, we harvested livers<br />

from 5 months old male mice of both genotypes. We found<br />

that DDB2 null livers have increased expression of MnSOD<br />

and decreased accumulation of ROS. Therefore, an impairment<br />

of oxidative damage could explain observed reduction in the<br />

number of H-ras12V-driven HCC nodules in DDB2 -/- mice.<br />

Since DDB2 is considered to act as a tumor suppressor, our<br />

finding that DDB2 loss protects mice from H-ras12V-driven liver<br />

tumor development is interesting and surprising, and it demonstrates<br />

that depending of the context of tumor development<br />

DDB2 can either inhibit or promote tumorigenesis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Dragana Kopanja, Akshay Pandey,<br />

Shuo Huang, Damiano Fantini, Pradip Raychaudhuri<br />

1989<br />

Inflammation contributes to hepatocarcinogenesis<br />

through an acceleration of transcription-coupled mutagenesis<br />

Tomonori Matsumoto, Norihiro Nishijima, Tsutomu Chiba, Hiroyuki<br />

Marusawa; Department of Gastroenterology and Hepatology,<br />

Graduate School of Medicine, Kyoto University, Kyoto, Japan<br />

Chronic inflammation plays important roles in cancer development<br />

in association with various intrinsic mediators including<br />

cytokines and growth factors, however, accumulation of<br />

genetic alterations in tumor-related genes is a critical step<br />

required for malignant transformation. Activation-induced cytidine<br />

deaminase (AID), an intrinsic DNA mutator enzyme, has<br />

been demonstrated to be aberrantly elicited in epithelial cells<br />

by inflammatory stimulation, and to contribute to tumorigenesis<br />

by inducing genetic aberrations. To gain further insight into the<br />

inflammation-mediated genotoxic events required for carcinogenesis,<br />

we examined the role of chronic inflammation in the<br />

emergence of genetic aberrations in the liver with constitutive<br />

AID expression. Treatment with thioacetamide (TAA) at lowdose<br />

concentrations caused minimal hepatic inflammation in<br />

both wild-type and AID transgenic (Tg) mice. Surprisingly, all<br />

the TAA-treated AID Tg mice developed multiple liver cancers in<br />

6 months, while wild-type mice with TAA or AID Tg mice without<br />

TAA rarely develop tumors during the same periods. Whole<br />

exome sequencing demonstrated the enhanced accumulation<br />

of somatic mutations in various genes in the liver of TAA-treated<br />

AID Tg mice. Microarray analyses showed the transcriptional<br />

upregulation of several putative tumor suppressor genes under<br />

hepatic inflammatory conditions induced by TAA treatment.<br />

Additional quantitative reverse transcription-polymerase chain<br />

reaction analyses and deep-sequencing analyses revealed<br />

the transcriptional upregulation and enhanced mutagenesis<br />

of putative tumor suppressor genes, including dual specificity<br />

phosphatase 6 (Dusp6), early growth response 1 (Egr1) and<br />

inhibitor of DNA binding 2 (Id2), in AID-expressing liver with<br />

TAA-mediated hepatic inflammation. Together, these findings<br />

suggest that inflammation-mediated transcriptional upregulation<br />

of target genes, including putative tumor suppressor genes,<br />

enhances the opportunity for inflamed cells to acquire somatic<br />

mutations and contributes to the acceleration of tumorigenesis<br />

in the inflamed liver tissues.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tomonori Matsumoto, Norihiro<br />

Nishijima, Tsutomu Chiba, Hiroyuki Marusawa<br />

1990<br />

Regulator of G-protein signaling 5 enhances portal vein<br />

invasion in hepatocellular carcinoma<br />

Yumi Umeno 1 , Sachiko Ogasawara 1 , Jun Akiba 1 , Hironori<br />

Kusano 1 , Osamu Nakashima 2 , Hironori Koga 3 , Takuji Torimura 3 ,<br />

Hirohisa Yano 1 ; 1 Department of Pathology, Kurume University<br />

School of Medicine, Kurume, Japan; 2 Clinical Laboratory Medicine,<br />

Kurume University Hospital, Kurume, Japan; 3 Department of<br />

Medicine, Division of Gastroenterology, Kurume University School<br />

of Medicine, Kurume, Japan<br />

Aim: Hepatocellular carcinoma (HCC) is one of the most<br />

common cancers in the world and a leading cause of cancer<br />

death. Portal vein invasion (PVI) is one of the major prognostic<br />

factors in patients with HCC. The aim of this study is to identify<br />

molecules to regulate PVI. Materials and Methods: We<br />

selectively collected tissues from each part of cancerous area,<br />

paired noncancerous area and PVI area by using laser microdissection<br />

(LMD) and frozen sections obtained from 3 HCC<br />

patients. Subsequently, cDNA microarray was conducted. Five<br />

upregulated or downregulated molecules in plural cases were<br />

identified and subjected to the following examinations. In order<br />

to evaluate the expression and the relationship to clinicopathologic<br />

factors, quantitative RT-PCR (qRT-PCR) and immunohistochemical<br />

stain were performed in 32 HCCs and in 60 HCCs,<br />

respectively. Results: We focused on 3 upregulated molecules,<br />

i.e., integrin beta 3 (ITGB3), osteopontin (OPN) and regulator<br />

of G-protein signaling 5 (RGS5) and 2 downregulated molecules,<br />

i.e., metallothionein 1G (MT1G) and metallothionein 1H<br />

(MT1H) in PVI tissue compared with cancerous tissue by cDNA<br />

microarray analysis. In qRT-PCR, RGS5 was significantly overexpressed<br />

in cancerous tissue compared with noncancerous<br />

tissue (p = 0.02). However, there was no significant difference<br />

in ITGB3 and OPN expression (respectively: p = 0.13 and p<br />

= 0.12). The expression of MT1G and MT1H was significantly


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1177A<br />

lower in cancerous tissue compared with noncancerous tissue<br />

(respectively: p < 0.01 and p < 0.01). There were no significant<br />

differences between the expression of these five molecules<br />

and any clinicopathologic factors, including PVI. Immunohistochemical<br />

stain for RGS5 demonstrated that RGS5 expression<br />

was higher in cancerous tissue than in paired noncancerous<br />

tissue in 63.3% (38/60) of HCC. Furthermore, high expression<br />

of RGS5 in cancerous tissue was significantly associated with<br />

PVI (p < 0.01) and tended to be associated with intrahepatic<br />

metastasis (p = 0.0626). The cases with high expression of<br />

RGS5 in cancerous tissue were significantly frequent in simple<br />

nodular type with extranodular growth or confluent multinodular<br />

type than in simple nodular type (p < 0.01). Conclusion: By<br />

using LMD and cDNA microarray, we detected molecules associated<br />

with PVI. RGS5 expression in cancerous tissue was significantly<br />

upregulated at mRNA and protein levels compared<br />

with noncancerous tissue and was significantly associated with<br />

PVI of HCC. RGS5 may be a useful prognostic biomarker as<br />

well as a target of molecular therapy of HCC. In vitro functional<br />

analysis of RGS5 in HCC is underway.<br />

Disclosures:<br />

The following authors have nothing to disclose: Yumi Umeno, Sachiko Ogasawara,<br />

Jun Akiba, Hironori Kusano, Osamu Nakashima, Hironori Koga, Takuji<br />

Torimura, Hirohisa Yano<br />

1991<br />

Correlation between Incidence Rate of Hepatocellular<br />

Carcinoma and Chemical Air Pollution in the State of<br />

Texas.<br />

Ali Shirafkan 1 , Cristiana Rastellini 1 , Katherine Ensor 2 , Loren Raun 2 ,<br />

Daria Zorzi 1 , Laura Campos 2 , Mauro Montalbano 1 , Tiziana Corsello<br />

1 , Luca Cicalese 1 ; 1 Department of Surgery, University of Texas<br />

Medical Branch, Galveston, TX; 2 Department of Statistics, Rice<br />

University, Houston, TX<br />

Introduction: Primary liver cancer is the third cause of cancer<br />

death in the world and the seventh in the United States. The<br />

known etiologies of Hepatocellular Carcinoma (HCC) are comprised<br />

of nonspecific cirrhosis, alcohol induced liver disease,<br />

hepatitis C (HCV) and B (HBV) infection. In addition, obesity<br />

and diabetes mellitus type two (T2DM) have been shown to<br />

increase the risk. Texas ranks third in the US with HCC incidence<br />

rate of 9.9 per 100,000. This is being compared to<br />

national average of 7.1 per 100,000, while distribution of the<br />

known risk factors in the state of Texas is similar to the national<br />

data. Texas has the largest oil and gas industry and is the<br />

second largest agricultural producer in the nation. Chemicals,<br />

fertilizers, herbicides and pesticides could potentially contribute<br />

to HCC development. Methods: The mean concentration<br />

value of Vinyl Chloride, Arsenic, Benzene and 1, 3 Butadiene<br />

reported by EPA (Environmental Protection Agency). Incidence<br />

rate of HCC in Texas by counties for the time period between<br />

2002 and 2010 was obtained from the Texas Cancer Registry.<br />

A scatter plot matrix, a logistic regression model and a generalized<br />

additive (GAM) logistic regression model were used<br />

to observe any correlation between these variables. Results:<br />

Increases in vinyl chloride and total Benzene and 1, 3 Butadiene<br />

were associated with an increased risk of HCC within a<br />

county. Arsenic levels appear protective at low levels, but at<br />

high levels the probability of HCC within a county increases.<br />

Conclusion: It is possible that vinyl chloride, arsenic, BZ and 1,<br />

3 BD produced in different areas of Texas probably contribute<br />

to the high incidence rates of HCC observed in this state.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ali Shirafkan, Cristiana Rastellini,<br />

Katherine Ensor, Loren Raun, Daria Zorzi, Laura Campos, Mauro Montalbano,<br />

Tiziana Corsello, Luca Cicalese<br />

1992<br />

MicroRNA profiling of human cholangiocarcinoma<br />

Hong Joo Kim, Yong Kyun Cho, Woo Kyu Jeon, Byung Ik Kim;<br />

Internal Medicine, Sungkyunkwan University Kangbuk Samsung<br />

Hospital, Seoul, Korea (the Republic of)<br />

Background and aims: Cholangiocarcinoma (CC) is a highly<br />

malignant epithelial cancer of the biliary tract with high morbidity<br />

and mortality. In the current study, we performed microR-<br />

NAs (miRNAs) expression microarrays to identify the up- and<br />

downregulated miRNAs in CC. Thereafter, we performed oligomer<br />

and antagomir transfection experiments for the most prominently<br />

up- and downregulated miRNA in CC to identify the<br />

underlying molecular pathogenic mechanisms for proliferation,<br />

invasion and metastasis of CC. Methods: Human hepatic duct<br />

bifurcation site cancer cells SNU-1196 (KCLB No. 01196),<br />

and human common bile duct cancer cells SNU-245 (KCLB No.<br />

00245) were purchased from Korean Cell Line Bank and maintained<br />

in RPMI1640 medium. Affymetrix miRNA expression<br />

microarrays were used to determine the miRNA expression profile<br />

of each CC cell line and normal control cells obtained from<br />

primary culture of normal bile duct epithelium. Transfections<br />

were performed using Lipofectamine RNAiMAX (Invitrogen).<br />

Cell proliferation was determined by MTT assay after mimics<br />

or antisense oligonucleotide transfection. Western blotting of<br />

cell extracts was performed using rabbit monoclonal antibodies<br />

against human SRGAP2, Rac1 (both from BD Bioscience) and<br />

β-actin. Results: To profile the miRNA expression patterns in<br />

human CC cells, we compared the miRNA expression profile<br />

of 2 CC cell lines with that of normal control cells obtained<br />

from primary culture of normal bile duct epithelium. MicroR-<br />

NAs downregulated more than 10-folds log ratio in CC cell<br />

lines than normal control cells were as follows: miR-145-5p,<br />

miR-125b-5p, miR-125a-5p, and miR-100-5p. Upregulated<br />

miRNAs more than 8-folds log ratio in CC cell lines than normal<br />

control cells were as follows: miR-182-5p, miR-196a-5p,<br />

miR-194-5p, miR-192-5p, and miR-182-5p. Thereafter, we<br />

performed transfection experiments using the mimics and<br />

antagomirs targeting miR-145-5p. There were no significant<br />

differences in cell proliferation measured by MTT assay among<br />

the CC and normal control cell lines transfected with mimics<br />

and antagomirs. Downregulation of miR-145-5p by antagomir<br />

transfection was associated with increase in protein expression<br />

levels of SRGAP2 and Rac1. Conclusions: Our data identify<br />

that miR-145-5p is the most profoundly downregulated miRNA<br />

in human CC cells. SRGAP2 and Rac1 can be novel targets of<br />

miR-145-5p, establishing a role for miR-145-5p in modulating<br />

motility and metastasis of CC cells. Our data provide a possibility<br />

for further investigations for future miR-145-5p targeted<br />

approaches of anticancer treatment in.CC.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hong Joo Kim, Yong Kyun Cho,<br />

Woo Kyu Jeon, Byung Ik Kim<br />

1993<br />

Cholangiocarcinoma cells interact with cancer-associated<br />

myofibroblasts in 3-D culture to provoke a strong<br />

desmoplastic-like reaction mediated by TGF-β.<br />

Alphonse E. Sirica, Akihiro Usui, Deanna J. Campbell; Pathology,<br />

Virginia Commonwealth University School of Medicine, Richmond,<br />

VA<br />

Intrahepatic cholangiocarcinomas (ICC) typically exhibit<br />

a prominent desmoplastic reaction marked by a dramatic<br />

increase in α-smooth muscle actin-positive cancer-associated<br />

fibroblasts (α-SMA+CAFs) together with the production of a


1178A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

dense collagen-enriched fibrous stroma. In order to further our<br />

understanding of the biological relationships mediating the desmoplastic<br />

reaction in ICC, we employed a novel rat 3-D culture<br />

model to investigate fibrogenic promoting interactions between<br />

cholangiocarcinoma cells (BDEsp-TDE cc<br />

cells) and α-SMA+CAFs<br />

(BDEsp-TDF sm<br />

cells) when co-cultured in a dilute collagen-type I<br />

gel matrix. Both BDEsp-TDE cc<br />

and BDEsp-TDF sm<br />

cells were established<br />

from a desmoplastic rat cholangiocarcinoma from liver.<br />

Compared with monoculture controls, co-culturing BDEsp-TDE cc<br />

cholangiocarcinoma cells constitutively expressing TGF-β1 with<br />

BDEsp-TDF sm<br />

CAFs provoked an intense desmoplastic-like reaction<br />

within the gel culture matrix, characterized by a prominent<br />

deposition of dense collagen fibers, together with a significant<br />

increase in the numbers of BDEsp-TDF sm<br />

CAFs accumulated<br />

within the gel matrix. This reaction was greatly suppressed by<br />

the TGF-β receptor 1 kinase inhibitor LY2157299. Similarly,<br />

treatment of BDEsp-TDF sm<br />

cells cultured alone within 3-D gel cultures<br />

with TGF-β1 significantly increased the numbers of these<br />

SMA+ CAFs within the gel matrix (Figure), which correlated<br />

with an overproduction of dense collagen fibers, and which<br />

was also inhibited by >70% by LY2157299. These results<br />

strongly suggest that paracrine TGF-β signaling plays a major<br />

role in generating the desmoplastic reaction in ICC and that<br />

targeting this pathway may have important implications for<br />

stromal depletion therapy for desmoplastic ICC. Supported by<br />

NIH R01 CA 083650.<br />

TGF-β1-induced enhancement of BDEsp-TDF sm<br />

CAF cell accumulation<br />

within 3-D gel monoculture compared with vehicle control<br />

cultures (VC).<br />

associated with oxidation of lipids and formation of oxysterols.<br />

Oxysterols are allosteric regulators of Smoothened (Smo), the<br />

receptor of the Hedgehog (Hh) signaling pathway, which is<br />

activated in cholangiocarcinoma. The step wise process of<br />

malignant transformation of biliary epithelial cells is poorly<br />

described. Purpose: We aimed to interrogate malignant transformation<br />

of the biliary epithelium upon stimulation of the Hh<br />

pathway by oxysterols. Methods: The human non-malignant<br />

(H69 and NHC) and primary PSC patient-derived cholangiocyte<br />

cell lines were employed for this study. Upon treatment<br />

with oxysterols [7-keto-27-hydrocholesterol and 20(S)-hydroxycholesterol],<br />

cholangiocytes were analyzed for Hh pathway<br />

stimulation with Smo translocation to the plasma membrane<br />

and cilia by immunofluorescence, migration in modified Boyden<br />

chambers, and colony-formation assay in a soft-agar. A<br />

miRNA array was conducted to identify miRNAs that were<br />

induced in cholangiocytes in an oxysterol- and Smo-dependent<br />

manner. miRNA targets were analyzed in silico. Results: Cholangiocytes<br />

expressed Hh pathway components at a baseline<br />

(Gli2, Smo). Smo translocation to the plasma membrane and<br />

cilia were observed in the cells treated with oxysterols (10mM<br />

for 16 hours) but not in the vehicle treated cells. Stimulation<br />

with oxysterols (10 mM for 7 hours) lead to a 1.3 – 2 (p <<br />

0.01) fold increase in cholangiocyte migration. Oxysterols promoted<br />

anchorage-independent growth evidenced by formation<br />

of colony-like structures by non-malignant cholangiocytes in a<br />

soft-agar over the 6 week period. miR-1246 was upregulated<br />

(logFC = 2.7) in H69 cells upon stimulation with 7-keto-27-hydrocholesterol<br />

and was inhibited (logFC = -2.2) with Vismodegib,<br />

direct-inhibitor of Smo, pre-treatment. The Hh pathway<br />

inhibitory receptor Patched-1, Axin2, and GSK-3β, all of each<br />

are involved in Hh signaling and carcinogenesis, were found<br />

to be a target for miR-1246 in silico. Conclusions: Products of<br />

lipid oxidation, oxysterols, promote malignant transformation<br />

of cholangiocytes in the Hh pathway-dependent manner. The<br />

oxysterol-Smo axis upregulation of miR-1246 might provide a<br />

positive feedback mechanism in this transformation.<br />

Disclosures:<br />

Gregory J. Gores - Advisory Committees or Review Panels: Conatus<br />

The following authors have nothing to disclose: Nataliya Razumilava, Anuradha<br />

Krishnan, Steven P. O’Hara<br />

Disclosures:<br />

The following authors have nothing to disclose: Alphonse E. Sirica, Akihiro Usui,<br />

Deanna J. Campbell<br />

1994<br />

Oxysterols Promote Malignant Transformation of Cholangiocytes<br />

in a Hedgehog Signaling-dependent Manner<br />

Nataliya Razumilava 2 , Anuradha Krishnan 1 , Steven P. O’Hara 1 ,<br />

Gregory J. Gores 1 ; 1 Mayo Clinic, Rochester, MN; 2 Division of<br />

Gastroenterology and Hepatology, University of Michigan, Ann<br />

Arbor, MI<br />

Background: Cholangiocarcinoma is a highly lethal neoplasm.<br />

It is often observed in a milieu of biliary tract inflammation<br />

1995<br />

A Randomized Prospective Open-label Trial Comparing<br />

Peginterferon Plus Adefovir or Tenofovir Combination<br />

Therapy Versus No Treatment in HBeAg-Negative<br />

Chronic Hepatitis B Patients with a Low Viral Load<br />

Annikki de Niet 1 , Louis Jansen 1 , Femke Stelma 1 , Sophie Willemse 1 ,<br />

Sjoerd Kuiken 2 , Sebastiaan Weijer 3 , Carin M. Van Nieuwkerk 4 ,<br />

Hans L. Zaaijer 5,1 , Richard Molenkamp 1 , Bart Takkenberg 1 ,<br />

Maarten Koot 5 , Joanne Verheij 1 , Ulrich Beuers 1 , Hendrik W.<br />

Reesink 1 ; 1 Academic Medical Center, Amsterdam, Netherlands;<br />

2 Sint Lucas Andreas Hospital, Amsterdam, Netherlands; 3 Medical<br />

Center Zuiderzee, Lelystad, Netherlands; 4 VU Medical Center,<br />

Amsterdam, Netherlands; 5 Sanquin, Amsterdam, Netherlands<br />

Background and aims: Chronic hepatitis B (CHB) patients with<br />

a low viral load (LVL) are currently not eligible for antiviral<br />

treatment. However, they comprise the largest group of CHB<br />

patients and are still at increased risk of developing cirrhosis or<br />

hepatocellular carcinoma. Here, we present the final results of a<br />

randomized trial designed to determine HBsAg loss and decline<br />

in CHB patients with LVL receiving combination treatment of<br />

peginterferon alfa-2a (peg-IFN) and a nucleotide analogue or<br />

no treatment. Methods: 134 CHB patients (HBeAg-negative,<br />

HBV-DNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1179A<br />

peg-IFN + adefovir (arm I; n=46), peg-IFN + tenofovir (arm<br />

II; n=45) or no treatment (arm III; n=43) for 48 weeks (ITT<br />

population), and followed-up for an additional 24 weeks. Randomization<br />

was stratified by HBV genotype A (22%), non-A (B<br />

7%, C 4%, D 26%, E/F/G 20%), or indeterminable (21%). The<br />

median age was 43 years, 57% were male. Twelve patients<br />

discontinued the study before week 72 (5 in arm I, 6 in arm<br />

II, 1 in arm III). HBsAg loss (AxSYM


1180A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2008-2014 for >2 years (ETV group). We excluded patients<br />

who developed HCC within the first year of enrollment. The<br />

primary endpoint was the development of HCC, and secondary<br />

endpoints were the development of cirrhotic complications<br />

and mortality. Results As of Dec 31, 2014, we enrolled 1315<br />

patients in the ETV group and 367 untreated patients before the<br />

development of HCC. Compared to untreated group, patients in<br />

the ETV group were significantly older and had more advanced<br />

liver disease. The mean duration of follow-up was 7 and 4<br />

years in the untreated and ETV groups, respectively. Compared<br />

with the untreated group, ETV therapy was associated with a<br />

significantly lower lifetime risk for HCC (Logrank P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1181A<br />

3.7 years and the median time from HBsAg loss to NA cessation<br />

was 0.6 years. At NA cessation, 61% of the patients<br />

were anti-HBs positive for a median duration of 0.5 years.<br />

After NA cessation, 4 (9%) patients relapsed. Three of them<br />

developed detectable HBV DNA (23, 223 and 1320 IU/mL)<br />

within the first 3 months. All 3 patients remained HBsAg negative,<br />

one was retreated while the other 2 returned to HBV<br />

DNA 1<br />

year and did not have significant co-morbidities were enrolled.<br />

Fasting serum, spot and 24 hour urine samples were collected.<br />

Fractional excretion of potassium (FEK + ), phosphate (FEPO 4<br />

),<br />

uric acid (FEUA), urinary β2MG/creatinine (Cr) and urine protein/creatinine<br />

ratio (UPCR) were calculated. The criteria of<br />

renal loss were defined as follow: proteinuria (24-hour urine<br />

protein >150 mg or positive protein dipstick), glucosuria while<br />

having normoglycemia, phosphate loss (FEPO 4<br />

>18%), uric<br />

acid loss (FEUA>15%), hypokalemia with renal K + loss, normal<br />

gap metabolic acidosis. Subclinical and definite proximal RTD<br />

was defined when there was a presence of 2 and ≥3 criteria,<br />

respectively. Receiver Operating Characteristic (ROC) was<br />

analyzed. The comparison of areas under the ROC was done.<br />

Results: Of 92 patients, subclinical and proximal RTD were<br />

seen in 15 (16.3%) and 9 (9.8%) patients. Median (range)<br />

duration of nucleotide analogue usage in normal, subclinical<br />

and definite proximal RTD were 45 (12-116), 66 (15-114),<br />

81 (42-108) months. The performance of spot urine protein,<br />

UPCR, urinary β2MG, urinary β2MG/Cr, FEPO 4<br />

and FEUA<br />

for detecting proximal RTD were showed in Table 1. Urinary<br />

β2MG yielded the highest AUROC [0.888 (0.804-0.972)]. At<br />

the cutoff of 300 mcg/L, urinary β2MG was 91.67% sensitivity<br />

and 76.47% specificity. There was no difference of diagnostic<br />

accuracy among tests (P=0.23). For detecting subclinical<br />

proximal RTD, urinary β2MG had the highest AUROC [0.822<br />

(0.699-0.946)]. At the cutoff of 300 mcg/L, urinary β2MG<br />

was 86.67% sensitivity and 77.27% specificity. Conclusions:<br />

16.3% and 9.8% CHB patients on nucleotide analogues developed<br />

subclinical and definite proximal RTD. Urinary β2MG<br />

had the highest accuracy for detecting proximal RTD. During<br />

nucleotide analogue treatment, urinary β2MG or UPCR should<br />

be monitored.<br />

Table 1. Receiver Operating Characteristic (ROC) analysis for<br />

detecting proximal RTD


1182A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Abhasnee Sobhonslidsuk - Grant/Research Support: Merck Sharp & Dohme<br />

corp, Eisai, Biotron; Speaking and Teaching: DKSH<br />

The following authors have nothing to disclose: Jirachaya Wanichanuwat, Bunyong<br />

Phakdeekitcharoen, Areepan Sophonsritsuk, Supanna Petraksa, Saowanee<br />

Kajanachumpol, Alongkorn Pugasab, Paisan Jittorntam, Anucha Kongsomgan,<br />

Pawin Numthavaj, Sittiruk Roytrakul<br />

2000<br />

Excellent outcomes of hepatitis B core antibody positive<br />

donors with oral antiviral agent alone in liver transplantation<br />

Tiffany C. Wong, James Fung, Kenneth S. Chok, See-Ching Chan,<br />

Chung Mau Lo; Surgery, The University of Hong Kong, Hong<br />

Kong, Hong Kong<br />

Background The use of hepatitis B core antibody (HBcAb) positive<br />

liver graft is controversial as antiviral protocol varies and<br />

the reported risk of de novo hepatitis B infection (HBV) ranged<br />

from 0 to 14%. Purpose To evaluate the risk of de novo HBV<br />

infection from HBcAb positive donor to Hepatitis B surface<br />

antigen (HBsAg) negative recipients in liver transplantation<br />

(LT). Method This was a retrospective review of all LT from<br />

2000 to 2014 in Queen Mary Hospital, Hong Kong. All data<br />

were retrieved from a prospective collected database. Results<br />

During study period, there were 1,034 LT at our center and<br />

128(12.4%) were from HBcAb positive donors to HBsAg negative<br />

recipients. Baseline demographics were listed in table<br />

1. All patients received antiviral prophylaxis only; lamivudine<br />

(n=117) and entecavir (n=11) with no hepatitis B immunoglobulin<br />

(HBIG) at any time point. De novo HBV as defined<br />

by HBsAg seropositive were seen in 3 (2.3%) recipients. They<br />

were 2 pediatric and 1 adult recipients. All had detectable<br />

HBV DNA (range from 179.3 to 373.6x10^6 copies/ml) at<br />

time of diagnosis and 2 had YMDD mutation. Liver biopsies<br />

were done in 2 of them and showed no evidence of fibrosis<br />

and immunostains for HBsAg and hepatitis B core antigen<br />

(HBcAg) werenegative. The median time from LT to de novo<br />

HBV was 59 (9-60) months. Tenofovir were added to the 2<br />

pediatric patients and adevofir was added to the adult patient.<br />

All patients with de novo HBV had undetectable HBV DNA after<br />

treatment. The median follow-up time was 71.1 (0.3 to 202.3)<br />

months. The 1-year and 5-year overall and graft survival were<br />

95.6%, 90.3% and 92.8% and 87.0% respectively. None had<br />

graft failure or mortality due to HBV infection. Conclusion The<br />

risk of de novo HBV infection is very low and had no impact on<br />

long-term patient and graft survival. Oral antiviral agent alone<br />

is effective and adequate to prevent de novo HBV in HBsAg<br />

negative patients who received a HBcAB positive graft.<br />

Table 1. Baseline demographic of our cohort<br />

Disclosures:<br />

The following authors have nothing to disclose: Tiffany C. Wong, James Fung,<br />

Kenneth S. Chok, See-Ching Chan, Chung Mau Lo<br />

2001<br />

Early prediction of response to peginterferon for<br />

HBeAg-positive chronic hepatitis B using HBsAg levels at<br />

week 4 and week 8 of therapy<br />

Milan J. Sonneveld 1 , Willem Pieter Brouwer 1 , Stefan Zeuzem 2 , Yilmaz<br />

Cakaloglu 3 , Fehmi Tabak 4 , Krzysztof. Simon 5 , Bettina E. Hansen<br />

1 , Harry L. Janssen 1,6 ; 1 Erasmus MC, University Medical Center<br />

Rotterdam, Rotterdam, Netherlands; 2 Medical Clinic 1, Johan<br />

Wolfgang Goethe University Medical Center, Frankfurt, Germany;<br />

3 Gastroenterohepatology, Istanbul Medical University, Istanbul,<br />

Turkey; 4 Hepatology, Cerrahpasa Medical School, Istanbul, Turkey;<br />

5 Infectious Disease, Medical University Wroclaw, Wroclaw,<br />

Poland; 6 Gastroenterology, University Health Network, Toronto,<br />

ON, Canada<br />

Background & Aims. Serum HBsAg levels can predict response<br />

to peginterferon (PEG-IFN) at week 12 of treatment; the performance<br />

of prediction-rules at earlier time-points is currently<br />

unknown. Methods. HBeAg-positive patients treated with PEG-<br />

IFN alfa-2b alone or in combination with lamivudine in a global<br />

randomized trial were enrolled if they had available outcome<br />

data and available HBsAg levels at week 4 and week 8 of<br />

therapy. Relationship between HBsAg at week 4 and week 8<br />

of treatment and response (HBeAg loss with HBV DNA


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1183A<br />

2002<br />

Final Results of Peginterferon Alfa-2b Add-on<br />

during Long-term Nucleos(t)ide Analogue Therapy in<br />

HBeAg-positive Patients – a Multicenter Randomized<br />

Controlled Trial (PEGON Study)<br />

Heng Chi 1 , Qing Xie 2 , Ning-Ping Zhang 3 , Xun Qi 4 , Chen Liang 4 ,<br />

Simin Guo 2 , Qing Guo 2 , Pauline Arends 1 , Jiyao Wang 3 , Elke Verhey<br />

1 , Robert J. de Knegt 1 , Bettina E. Hansen 1 , Harry L. Janssen 1,5 ;<br />

1 Gastroenterology and Hepatology, Erasmus MC University Medical<br />

Center Rotterdam, Rotterdam, Netherlands; 2 Department of<br />

Infectious Diseases, Ruijin Hospital, Jiaotong University, Shanghai,<br />

China; 3 Department of Gastroenterology and Hepatology, Zhongshan<br />

hospital, Fudan University, Shanghai, China; 4 Department of<br />

Hepatitis Disease, Shanghai Public Health Clinical Center, Fudan<br />

University, Shanghai, China; 5 Toronto Centre for Liver Disease,<br />

Toronto General & Western Hospital, University of Toronto, Shanghai,<br />

ON, Canada<br />

BACKGROUND: Nucleos(t)ide analogues (NA) are potent<br />

inhibitors of viral replication in chronic hepatitis B (CHB)<br />

patients. However, sustained off-treatment response is infrequently<br />

achieved indicating the necessity of long-term therapy.<br />

Peginterferon (PEG-IFN) add-on in HBeAg-positive patients on<br />

long-term NA may facilitate serological responses and finite<br />

therapy. METHODS: In this investigator-initiated randomized<br />

trial conducted in Europe and China, 82 HBeAg-positive<br />

patients were treated for at least 12 months with entecavir<br />

(ETV) or tenofovir (TDF) with suppressed HBV DNA and HBeAg<br />

positivity at randomization. Patients were randomized to 48<br />

weeks of PEG-IFN alfa-2b add-on, or continued NA monotherapy.<br />

Response (HBeAg seroconversion with HBV DNA


1184A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

improving PAGE-B scores developed a clinical event, whereas<br />

8/224 patients with stable and 3/71 patients with worsening<br />

PAGE-B scores did. Conclusion: In this large real-life European<br />

cohort of Caucasians and Asians treated with ETV, on-treatment<br />

PAGE-B score could predict HCC, as well as other liver-related<br />

and clinical events. Hence, the PAGE-B score obtained during<br />

ETV/TDF treatment could predict adverse clinical outcomes.<br />

Disclosures:<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

Ashley S. Brown - Advisory Committees or Review Panels: MSD, Roche, Bristol-Myers-Squibb,<br />

Gilead, Janssen, Abbvie, Achillion; Speaking and Teaching:<br />

MSD, Roche, Bristol-Myers Squibb, Gilead, Janssen, Abbvie<br />

Ivana Carey - Grant/Research Support: Gilead, Roche; Speaking and Teaching:<br />

BMS<br />

David J. Mutimer - Advisory Committees or Review Panels: Gilead Sciences,<br />

AbbVie, Janssen, MSD, BMS; Speaking and Teaching: Gilead Sciences, AbbVie,<br />

Janssen, BMS<br />

Florian van Bömmel - Advisory Committees or Review Panels: Gilead Sciences<br />

Inc., Roche Pharmaceutics; Grant/Research Support: Biopredictive; Speaking<br />

and Teaching: Bristol-Myers Squibb, Janssen, Roche Pharmaceutics, Gilead Sciences<br />

Inc., Abbvie<br />

Robert J. de Knegt - Advisory Committees or Review Panels: Roche, Norgine,<br />

Janssen Cilag, AbbVie; Grant/Research Support: Roche, Janssen Cilag, BMS,<br />

AbbVie; Speaking and Teaching: Gilead, Roche, Janssen Cilag, AbbVie<br />

Tania M. Welzel - Advisory Committees or Review Panels: Novartis, Janssen,<br />

Gilead, Abbvie, Boehringer-Ingelheim+, BMS<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Fabien Zoulim - Advisory Committees or Review Panels: Janssen, Gilead, Novira,<br />

Abbvie, Tekmyra, Transgene; Consulting: Roche; Grant/Research Support:<br />

Novartis, Gilead, Scynexis, Roche, Novira, Assemblypharm; Speaking and<br />

Teaching: Bristol Myers Squibb, Gilead<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Heng Chi, Willem Pieter Brouwer,<br />

Massimo Fasano, Katja Deterding, Ye H. Oo, Pauline Arends, Teresa A. Santantonio,<br />

Pierre Pradat, Bettina E. Hansen<br />

2004<br />

Long Term Efficacy and Safety of Tenofovir DF (TDF) in<br />

Chronic Hepatitis B Patients (CHB) with Documented<br />

Lamivudine Resistance (LAM-R): 5 Year Results From a<br />

Randomized, Controlled Trial<br />

Scott Fung 1 , Hie-Won Hann 2 , Magdy Elkhashab 3 , Thomas Berg 4 ,<br />

Milotka J. Fabri 5 , Andrzej Horban 6 , Mijomir Pelemis 7 , Ioan<br />

Sporea 8 , John F. Flaherty 9 , Benedetta Massetto 9 , Kyungpil Kim 9 ,<br />

Kathryn M. Kitrinos 9 , Mani Subramanian 9 , Cihan Yurdaydin 10 ,<br />

Edward J. Gane 11 ; 1 Toronto General Hospital, Toronto, ON,<br />

Canada; 2 Thomas Jefferson University Hospital, Philadelphia, PA;<br />

3 Toronto Liver Centre, Toronto, ON, Canada; 4 University Hospital<br />

Leipzig, Leipzig, Germany; 5 Clinic for Infectious Diseases,<br />

Novi Sad, Serbia; 6 Medical University of Warsaw, Warsaw,<br />

Poland; 7 Clinic for Infectious and Tropical Diseases,, Belgrade,<br />

Serbia; 8 University of Medicine and Pharmacy Timisoara, Timisoara,<br />

Romania; 9 Gilead Sciences, Inc., Foster City, CA; 10 Ankara<br />

Üniversitesi, Ankara, Turkey; 11 Auckland General Hospital, Auckland,<br />

New Zealand<br />

Background: In CHB patients with LAM-resistance (LAM-R), TDF<br />

has shown efficacy comparable to FTC/TDF and no detectable<br />

TDF resistance at 2 years (Gastroenterology 2014;146:980-<br />

88). The final 5 year efficacy and safety results from this trial<br />

are presented. Methods: CHB patients on LAM with HBV DNA<br />

≥3 log 10<br />

IU/mL and with documented LAM-R (INNO-LiPA<br />

Multi-DR, v3) were randomized (1:1) to TDF or FTC/TDF and<br />

followed in a blinded fashion for 5 years. Results: Two hundred<br />

eighty patients were randomized; 239 (85%) completed 5<br />

years of treatment. At baseline, mean age was 47 years, most<br />

were male (75%) and non-Asian (66%); 53% were HBeAgnegative,<br />

and HBV genotype distribution (A-D) was 22%, 13%,<br />

19%, and 43%, respectively. Mean (SD) HBV DNA was 5.7<br />

(1.9) log 10<br />

IU/mL, and 42% had ALT ≤ULN at baseline. At<br />

Year 5, virologic, serologic, and biochemical responses were<br />

similar among groups, and remained stable from Year 2 to 5<br />

(Table). Nine patients (4-TDF, 5-FTC/TDF) discontinued due to<br />

an adverse event, including increased serum creatinine in 1<br />

patient. Hepatocellular carcinoma was reported in 4 (1.4%)<br />

patients. Confirmed renal safety endpoints (both groups combined)<br />

over 5 years were: CrCL


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1185A<br />

Disclosures:<br />

Scott Fung - Advisory Committees or Review Panels: Gilead Sciences, Merck,<br />

Vertex, Roche; Speaking and Teaching: BMS, Gilead, AbbVie, Janssen, Merck<br />

Hie-Won Hann - Advisory Committees or Review Panels: Gilead; Grant/Research<br />

Support: Gilead, Bristol-Myers Squibb<br />

Magdy Elkhashab - Advisory Committees or Review Panels: GSK, Gilead, Merk,<br />

BMS, ABBVIE; Grant/Research Support: EISAI, GENENTECH, GSK, GILEAD,<br />

ABBVIE, Merk, BMS<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

John F. Flaherty - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Benedetta Massetto - Employment: Gilead Sciences, Inc.; Stock Shareholder:<br />

Gilead Sciences, Inc<br />

Kathryn M. Kitrinos - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Mani Subramanian - Employment: Gilead Sciences<br />

Cihan Yurdaydin - Advisory Committees or Review Panels: Janssen, Roche,<br />

Merck, Gilead, AbbVie; Speaking and Teaching: BMS<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

The following authors have nothing to disclose: Milotka J. Fabri, Andrzej Horban,<br />

Mijomir Pelemis, Ioan Sporea, Kyungpil Kim<br />

2005<br />

Plasma MicroRNA Levels are Associated with Therapy<br />

Response in Chronic Hepatitis B Patients Treated with<br />

Peginterferon and Adefovir<br />

Meike H. van der Ree 1,2 , Louis Jansen 1,2 , Karel A. van Dort 2 , Bart<br />

Takkenberg 1 , Neeltje A. Kootstra 2 , Hendrik W. Reesink 1,2 ; 1 Gastroenterology<br />

and Hepatology, AMC, Amsterdam, Netherlands;<br />

2 Experimental Immunology, AMC, Amsterdam, Netherlands<br />

Background and Aims: MicroRNAs (miRNAs) are small,<br />

non-coding RNA molecules involved in various cellular processes<br />

of post-transcriptional regulation of gene expression.<br />

Circulating miRNAs have been linked to hepatitis B virus (HBV)<br />

infection. The aim of this study was to identify plasma miRNAs<br />

signatures at baseline that predict response to peginterferon<br />

based therapy in chronic hepatitis B (CHB) patients. Methods:<br />

We included 86 CHB patients who participated in an investigator-initiated<br />

study and completed 48 weeks of peginterferon<br />

alfa-2a and adefovir combination therapy, followed by<br />

a treatment-free follow-up of 2 years (week 144). RNA was<br />

isolated from baseline plasma samples with the miRCURY RNA<br />

isolation kit (Exiqon). A miRNA profile (qPCR panel of 179<br />

miRNAs, Exiqon) was determined in an identification cohort<br />

of 12 HBeAg-positive and 12 HBeAg-negative patients, both<br />

consisting of 6 non-responders (NR) and 6 combined responders<br />

(CR) (HBeAg negativity, HBV DNA ≤2,000 IU/mL, and ALT<br />

normalization) of which 3 patients had HBsAg loss. Then, a<br />

selection of 28 miRNAs (>1.5-fold difference (p2-fold difference (p


1186A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ings may be significant enough to even warrant a Black Box<br />

Warning by the FDA<br />

Disclosures:<br />

John R. Lake - Advisory Committees or Review Panels: BMS; Consulting: Vital<br />

Therapies, Novartis; Grant/Research Support: Gilead, Salix, Ocera, Essai<br />

The following authors have nothing to disclose: Amar Mahgoub, Tyson Sievers<br />

2007<br />

TKM-HBV, a Novel RNA Interference Treatment for<br />

Chronic Hepatitis B, Rapidly Reduces Surface Antigen<br />

and other Viral Proteins in both Intrahepatic and Peripheral<br />

Compartments<br />

Amy C. Lee, Emily P. Thi, Luying Pei, Xin Ye, Jennifer Cross,<br />

Sandie Y. Du, Ammen P. Dhillon, Janet R. Phelps, Kevin McClintock,<br />

Michael Abrams, Ian MacLachlan; Tekmira Pharmaceuticals,<br />

Burnaby, BC, Canada<br />

TKM-HBV is a novel RNA interference (RNAi) therapeutic intervention<br />

against hepatitis B virus (HBV) and currently in Phase 1<br />

clinical development. It is designed to reduce the viral antigen<br />

load in chronically infected patients and allow the body to<br />

escape the state of immune repression imposed by the virus.<br />

A mixture of three different oligonucleotide duplexes encapsulated<br />

within a lipid nanoparticle delivery system, TKM-HBV acts<br />

directly on all HBV RNAs (pregenomic RNA as well as viral<br />

mRNA) via nucleotide sequence-specific cleavage. Because it<br />

prevents the synthesis of viral proteins and reduces the overall<br />

antigen load in the body, this mode of drug action may<br />

present advantages over other approaches that seek to block<br />

viral protein secretion into the bloodstream thereby potentially<br />

causing intracellular build-up and ER stress. A hydrodynamic<br />

injection (HDI) immunodeficient NOD.CB17-Prkdc scid /J mouse<br />

model of HBV was used to characterize viral RNA targeting<br />

and the time course of viral antigen reduction following a single<br />

administration of TKM-HBV and its effective distribution<br />

to its target organ, the liver. HBV RNA was measured using<br />

branched DNA assays, HBV DNA was quantitated via QPCR,<br />

and HBV proteins were detected using ELISA and immunohistochemistry<br />

methods. Over 92% reduction of total HBV RNA in<br />

the liver was observed as soon as 2 days after treatment, with<br />

a slightly more moderate effect on the 3.5 kb species alone<br />

suggesting this molecule is less accessible to the cellular RNAi<br />

machinery than other HBV transcripts. Onset of serum HBV<br />

DNA reduction was also highly rapid, with the nadir of >98%<br />

inhibition detected at Day 2. Maximal treatment effects were<br />

essentially reached at Day 4 for intrahepatic and serum HBsAg<br />

(98% and 97%, respectively) whereas reduction of intrahepatic<br />

HBcAg appeared to be a more gradual process, continuing to<br />

decrease beyond Day 4 with a nadir reached at Day 7. Taken<br />

together, these data suggest that the peripheral compartment is<br />

a dynamic environment with rapid turnover of viral markers in<br />

the blood stream whereas the turnover of viral proteins within<br />

the intracellular compartment of liver cells occurs more slowly.<br />

The kinetics of viral marker turnover may differ in a more<br />

immunocompetent system. In summary, TKM-HBV effectively<br />

removed viral antigens from both the intrahepatic and peripheral<br />

compartments within days after a single treatment dose<br />

in HDI mice. These viral elements include immunomodulatory<br />

surface and core proteins which are implicated in mediating<br />

the immune-repressed condition of chronic HBV infection.<br />

Disclosures:<br />

Amy C. Lee - Employment: Tekmira<br />

Emily P. Thi - Employment: Tekmira Pharmaceuticals<br />

Xin Ye - Employment: Tekmira Pharmaceuticals Corporation<br />

Kevin McClintock - Employment: Tekmira Pharmaceuticals, Tekmira Pharmaceuticals,<br />

Tekmira Pharmaceuticals, Tekmira Pharmaceuticals<br />

The following authors have nothing to disclose: Luying Pei, Jennifer Cross, Sandie<br />

Y. Du, Ammen P. Dhillon, Janet R. Phelps, Michael Abrams, Ian MacLachlan<br />

2008<br />

Long-Term Clinical And Serological Outcome Of Hepatitis<br />

B E-Antigen Negative Chinese-Pegylated Interferon<br />

Versus Prolonged Tenofovir Disoproxil Fumarate Or<br />

Entecavir-Eight Years Follow-Up<br />

Guofeng Chen 1 , Qing Shao 1 , Dong Ji 1 , Fan Li 1 , Bing Li 1 , Zhongbin<br />

Li 1 , Jialiang Liu 1 , Xiaoxia Niu 1 , Yudong Wang 2 , Cheng Wang 2 ,<br />

Jing Chen 2 , Vanessa Wu 2 , George K. Lau 1,2 ; 1 Liver Fibrosis Diagnosis<br />

and Treatment Center, 302 Hospital, Beijing, China; 2 Gastroenterology<br />

& Hepatology, Humanity & Health Medical Group,<br />

Hong Kong, China<br />

Background and Aims Long-term follow-up clinical and serological<br />

data after a finite course of 48 wks pegylated interferon-based<br />

therapy as compared with prolonged nucleos(t)<br />

ide analogues (NUCs) therapy in HBeAg negative Chinese<br />

patients, is lacking. Methods In a single center, 121 HBeAg<br />

negative Chinese patients treated with a 48 wks course of<br />

pegylated interferon (pIFN) were compared with 346 nucleos(t)ide-naïve<br />

patients treated with prolonged TDF (n=158) or<br />

ETV (n=188) for over eight years. Baseline liver fibrosis was<br />

assessed by either liver biopsy or by Fibroscan (Echosens).<br />

All patients were followed up at 3-6 monthly intervals, with<br />

serial measurement of liver function test, serum HBV DNA,<br />

HBsAg, anti-HBs (if HBsAg turn negative) and liver stiffness<br />

by fibroscan. Quantitative HBsAg (qHBsAg), with the use of<br />

either Architect HBsAg (Abbott Laboratories, Ireland) or Elecsys<br />

HBsAg (Roche Diagnostics, Germany) was also performed on<br />

all serum samples available. Multivariable regression analysis<br />

was performed to identify the independent predictors of HBsAg<br />

loss, liver fibrosis regression and development of hepatocellular<br />

carcinoma. Results Twenty-four (19.8%) pIFN-treated as<br />

compared to 9 (2.6%) NUCs-treated patients had serological<br />

clearance of HBsAg (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1187A<br />

2009<br />

Functional Activation of NK and CD8 + T Cells In Vitro by<br />

the Toll-Like Receptor 7 Agonist GS-9620<br />

Holly Micolochick Steuer 1 , Stephane Daffis 1 , Sophie M. Lehar 1 ,<br />

Adam Palazzo 1 , Hugo Tharinger 1 , Christian Frey 1 , Stefan Pflanz 1 ,<br />

Congrong Niu 1 , Christine Y. Chang 2 , Michelle Q. Jin 2 , Vincent L.<br />

Chen 3 , William E. Delaney 1 , Leanne Peiser 1 , Simon P. Fletcher 1 ,<br />

Mindie H. Nguyen 2 ; 1 Gilead Sciences, Foster City, CA; 2 Division<br />

of Gastroenterology and Hepatology, Stanford University Medical<br />

Center, Palo Alto, CA; 3 Department of Medicine, Stanford University<br />

Medical Center, Palo Alto, CA<br />

Background & Aims: GS-9620, an oral agonist of toll-like receptor<br />

7 (TLR7), is in phase 2 <strong>studies</strong> for the treatment of chronic<br />

hepatitis B (CHB). We previously demonstrated that reduction<br />

of HBsAg levels by GS-9620 in HBV-infected chimpanzees was<br />

associated with the intrahepatic induction of NK and CD8 + T<br />

cell transcriptional signatures, together with biochemical and<br />

immunohistochemical evidence of hepatocyte cell death. To<br />

further explore the cellular mechanisms of antiviral response to<br />

GS-9620, we characterized the effect of GS-9620 on primary<br />

human NK and CD8 + T cells in vitro. Methods: Peripheral blood<br />

mononuclear cells (PBMC) obtained from healthy volunteers<br />

(HV) and CHB patients (HBeAg-positive and HBeAg-negative<br />

patients, treatment naïve and virally suppressed with nucleos(t)<br />

ide therapy) were treated with 1-100 nM GS-9620. NK cell<br />

activation and effector function in HV PBMCs was analyzed<br />

by flow cytometry, and cytolytic activity was measured with a<br />

fluorescent cell-based cytotoxicity assay. CD8 + T cell function<br />

was assessed at recall by flow cytometry after expansion of HV<br />

or CHB PBMCs for 6-9 days with CEFT (CMV, EBV, Influenza<br />

and Tetanus toxin) or HBV core peptides, respectively. Results:<br />

GS-9620 potently induced expression of the activation marker<br />

CD69 on NK cells from HV PBMCs, and augmented NK cell<br />

cytotoxicity against HepG2 cells and HBV-infected primary<br />

human hepatocytes (PHH). In contrast, GS-9620 did not induce<br />

intracellular IFN-γ production in NK cells. GS-9620 induced<br />

comparable IFN-α levels in HV (n=5) and CHB PBMCs (n=11)<br />

(mean 1403 vs. 1075 pg/mL, respectively, p=0.44). This was<br />

notable because receptor blocking experiments demonstrated<br />

that GS-9620-dependent NK cell activation was mediated<br />

through type I IFN. GS-9620 also increased the frequency of<br />

CD8 + T cells expressing the cytotoxic marker CD107a after<br />

HBV core peptide expansion of CHB PBMCs (n=11, mean<br />

1.9-fold increase, p=0.03). Similarly, GS-9620 increased the<br />

frequency of CD107a + CD8 + T cells in HV PBMCs after CEFT<br />

peptide expansion (n=9, mean 2.1-fold increase, p=0.03).<br />

In contrast, GS-9620 did not alter the frequency of IFN-γ or<br />

TNF-α secreting CD8 + T cells in either HV or CHB PBMCs after<br />

peptide expansion. Conclusion: GS-9620 induced a cytotoxic<br />

phenotype but not antiviral cytokine production in human NK<br />

and CD8 + T cells in vitro. These in vitro functional data suggest<br />

that the antiviral response to GS-9620 in vivo is likely mediated,<br />

at least in part, by the cytolytic activity of NK and/or<br />

CD8 + T cells.<br />

Disclosures:<br />

Holly Micolochick Steuer - Employment: Gilead Sciences, Pharmasset, Inc, Pharmasset,<br />

Inc, Pharmasset, Inc<br />

Stephane Daffis - Employment: Gilead Sciences<br />

Adam Palazzo - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Christian Frey - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Stefan Pflanz - Employment: Gilead Sciences<br />

Congrong Niu - Employment: Gilead Science<br />

William E. Delaney - Employment: Gilead Sciences; Patent Held/Filed: Gilead<br />

Sciences; Stock Shareholder: Gilead Sciences<br />

Leanne Peiser - Employment: Gilead Sciences, Merck; Stock Shareholder: Gilead<br />

Sciences, Merck<br />

Simon P. Fletcher - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Sophie M. Lehar, Hugo Tharinger,<br />

Christine Y. Chang, Michelle Q. Jin, Vincent L. Chen<br />

2010<br />

The incidence and predictors of HBsAg loss and hepatocellular<br />

carcinoma development after the cessation of<br />

lamivudine and entecavir treatment in chronic hepatitis<br />

B patients<br />

Chien-Hung Chen, Chuan-Mo Lee, Chao-Hung Hung, JIng-Houng<br />

Wang, Sheng-Nan Lu, Tsung-Hui Hu; Chang Gung Memorial Hospital-Kaohsiung<br />

Medical Center, Kaohsiung, Taiwan<br />

Background: It remained unclear that the incidence and predictors<br />

of hepatitis B surface antigen (HBsAg) loss and hepatocellular<br />

carcinoma (HCC) development after the cessation of<br />

nucleos(t)ide analogs (NA) therapy. Aims: To investigate the<br />

incidence and predictors of HBsAg loss and HCC developement<br />

after stopping NA treatment in chronic hepatitis B (CHB)<br />

patients. Patients and Methods: From 2004 to 2012, a total<br />

of 450 CHB patients (169 HBeAg-positive, 281 HBeAg-negative)<br />

(84 cirrhosis) received lamivudine (n=198) or entecavir<br />

(n=252) treatment and have stopped the treatment at least 12<br />

months were recruited. All patients fulfilled the stopping criteria<br />

of the APASL 2008 or 2012. Results: In 169 HBeAg-positive<br />

patients, the cumulative rates of HBsAg loss in years 3, 5 and<br />

8 were 3.6%, 9.9% and 18% respectively, after stopping NA<br />

treatment. Cox regression analysis showed that old age, lower<br />

baseline HBV DNA and end-of-treatment qHBsAg were independent<br />

predictors for HBsAg loss. In 281 HBeAg-negative<br />

patients, the cumulative rates of HBsAg loss in years 3, 5 and<br />

8 were 14.3%, 26.8% and 45.4% respectively, after stopping<br />

NA treatment. Cox regression analysis showed that male, cirrhosis,<br />

lower baseline ALT and end-of-treatment qHBsAg levels<br />

were independent predictors for HBsAg loss. Of the 56 patients<br />

with HBsAg loss, 8 and 4 experienced virological and clinical<br />

relapse respectively, before HBsAg loss. For all patients, the<br />

cumulative rates of HBsAg loss in patients who had end-oftreatment<br />

qHBsAg 1000 IU/mL<br />

in year 7 were 71.9%, 45.1%, 27.3% and 7.5% respectively.<br />

For all patients, the cumulative rates of new HCC development<br />

in year 10 (follow-up period included re-treatment duration) in<br />

non-cirrhotic and cirrhotic patients were 2% and 33% respectively.<br />

Cox regression analysis showed that old age, cirrhosis<br />

and longer treatment duration were independent predictors for<br />

new HCC development. HBsAg loss after stopping NA treatment<br />

was not a significant factor for HCC development. Five<br />

patients (4 cirrhosis) who experienced HBsAg loss developed<br />

new HCC. Conclusions: End-of-treatment qHBsAg level was a<br />

useful predictor for HBsAg loss after stopping lamivudine and<br />

entecavir treatment in CHB patient. HBsAg loss after stopping<br />

NA treatment did not completely eliminate the risk of HCC,<br />

especially in cirrhotic patient.<br />

Disclosures:<br />

The following authors have nothing to disclose: Chien-Hung Chen, Chuan-Mo<br />

Lee, Chao-Hung Hung, JIng-Houng Wang, Sheng-Nan Lu, Tsung-Hui Hu


1188A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2011<br />

The combination of IFN-Lambda-4 and HLA-DPB1 polymorphisms<br />

accurately predict HBsAg loss in interferon<br />

treated genotype D patients with HBeAg-negative<br />

chronic hepatitis B<br />

Pietro Lampertico 1 , Enrico Galmozzi 1 , Floriana Facchetti 1 , Cristina<br />

Cheroni 2 , Federica Invernizzi 1 , Vincenza Valveri 2 , Roberta<br />

Soffredini 1 , Mauro Viganò 3 , Sergio Abrignani 2 , Raffaele De Francesco<br />

2 , Massimo Colombo 1 ; 1 Division of Gastroenterology and<br />

Hepatology, Fondazione IRCCS Ca’ granda Ospedale Maggiore<br />

policlinico, Università degli Studi di Milano, Milan, Italy; 2 INGM<br />

- Istituto Nazionale Genetica Molecolare “Romeo ed Enrica Invernizzi”,<br />

Milan, Italy; 3 Division of Hepatology, Ospedale San<br />

Giuseppe, Università degli Studi di Milano, Milan, Italy<br />

Background. HLA-DP polymorphisms have recently been associated<br />

to spontaneous hepatitis B virus (HBV) clearance in<br />

Asians patients whereas polymorphisms mapping the interferon<br />

lambda 3 and 4 (IFNL3 and 4) genes predicted the outcome<br />

of IFN-based therapy of HCV infection. Whether the latter<br />

applies also to HBsAg clearance in IFN treated difficult to cure<br />

HBeAg-negative genotype D chronic hepatitis B Caucasian<br />

patients, is unknown. Methods. 126 such patients with compensated<br />

disease (46 years, 82% males, 90% genotype D,<br />

HBV DNA 6.2 log cp/ml, ALT 132 IU/L, 40% with cirrhosis)<br />

received (Peg)IFN alfa for 22 (6-48) months and were followed<br />

for 11 (1-23) years with HBsAg clearance as a primary endpoint.<br />

HLA-DPA1 (rs3077 A/G), HLA-DPB1 (rs9277534 A/G<br />

and rs9277535 A/G), IFNL3 (rs8099917 T/G) and IFNL4<br />

(rs12979860 C/T, rs368234815 TT/ΔG and rs117648444<br />

C/T) polymorphisms were assessed by both TaqMan SNP<br />

Genotyping Assays and Sanger sequencing. Results. Twenty-eight<br />

patients ultimately cleared HBsAg with a 18-year<br />

cumulative probability of 30%, with a preference for patients<br />

with lower HBV DNA levels and higher ALT levels at baseline,<br />

HLA-DPB1 AG/GG, HLA-DPA1 AG/GG, IFNL3 rs8099917<br />

TT genotypes and the combination of IFNL4 rs368234815<br />

TT/TT and rs117648444 CT/TT genotypes. At multivariate<br />

analysis, HLA-DPB1 rs9277535 AG/GG genotype was the<br />

strongest independent baseline predictor of HBsAg seroclearance<br />

(HR: 6.61; 95%CI 2.6-16, p5 yrs<br />

(mean follow-up:6.4, median:6.2). Results: HCCs have been<br />

diagnosed in 90/1946 (4.6%) patients within the first 5 yrs<br />

and 7/794 (0.9%) patients remaining at risk beyond the first 5<br />

yrs of therapy. The 7 HCCs diagnosed after yr-5 (at 5.4-6.7 yrs<br />

after ETV/TDF onset) developed in 4/531 (0.8%) non-cirrhotic<br />

and 3/224 (1.3%) cirrhotic male patients [50-75 yrs old at<br />

ETV (n=2) or TDF (n=5) onset] who had remained in on-therapy<br />

virological remission since the first yr (the 75-yr old patient had<br />

discontinued TDF at 5 yrs and had normal ALT and HBVDNA<br />


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1189A<br />

yrs:1.07%, p=0.046). HCC development beyond yr-5 was<br />

associated with older age (RH per yr:1.069, 95% CI:0.997-<br />

1.146; p=0.062) or age ≥55 yrs at ETV/TDF start (p=0.029<br />

by log-rank), but not with gender (p=0.340) or presence of<br />

cirrhosis at baseline (p=0.371). Conclusions: The HCC risk<br />

seems to be decreasing after the first 5 yrs of ETV/TDF therapy<br />

in CHB patients, especially in those with compensated cirrhosis<br />

at baseline. Older age at treatment initiation, particularly age<br />

≥55 yrs, appears to represent the main risk factor associated<br />

with late HCC development.<br />

Disclosures:<br />

George V. Papatheodoridis - Advisory Committees or Review Panels: Merck<br />

Sharp & Dohme, Novartis, Abbvie, Boerhinger Ingelheim, Bristol-Meyer Squibb,<br />

Gilead, Roche, Janssen, GlaxoSmith Kleine; Grant/Research Support: Roche,<br />

Gilead, Bristol-Meyer Squibb, Abbvie, Janssen; Speaking and Teaching: Merck<br />

Sharp & Dohme, Bristol-Meyer Squibb, Gilead, Roche, Janssen, Abbvie<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

Florian van Bömmel - Advisory Committees or Review Panels: Gilead Sciences<br />

Inc., Roche Pharmaceutics; Grant/Research Support: Biopredictive; Speaking<br />

and Teaching: Bristol-Myers Squibb, Janssen, Roche Pharmaceutics, Gilead Sciences<br />

Inc., Abbvie<br />

Ioannis Goulis - Consulting: Gilead Sciences, Abbvie, Janssen-Cilag, Janssen-Cilag,<br />

BMS; Grant/Research Support: BMS; Speaking and Teaching: BMS, Gilead<br />

Sciences, Janssen-Cilag<br />

Spilios Manolakopoulos - Advisory Committees or Review Panels: NOVARTIS,<br />

ABBVIE, MSD, BMS, GILEAD, JANNSEN; Consulting: GILEAD, BMS, ABBVIE,<br />

JANNSEN, MSD; Grant/Research Support: BMS, GILEAD; Speaking and Teaching:<br />

ABBVIE, MSD, GILEAD, BMS, GSK, JANNSEN<br />

Cihan Yurdaydin - Advisory Committees or Review Panels: Janssen, Roche,<br />

Merck, Gilead, AbbVie; Speaking and Teaching: BMS<br />

Thomas Berg - Advisory Committees or Review Panels: Gilead, BMS, Roche,<br />

Tibotec, Vertex, Jannsen, Novartis, Abbott, Merck, Abbvie; Consulting: Gilead,<br />

BMS, Roche, Tibotec; Vertex, Janssen; Grant/Research Support: Gilead, BMS,<br />

Roche, Tibotec; Vertex, Jannssen, Merck/MSD, Boehringer Ingelheim, Novartis,<br />

Abbvie; Speaking and Teaching: Gilead, BMS, Roche, Tibotec; Vertex, Janssen,<br />

Merck/MSD, Novartis, Merck, Bayer, Abbvie<br />

Massimo Colombo - Advisory Committees or Review Panels: BRISTOL-MEY-<br />

ERS-SQUIBB, SCHERING-PLOUGH, ROCHE, GILEAD, BRISTOL-MEYERS-SQUIBB,<br />

SCHERING-PLOUGH, ROCHE, GILEAD, Janssen Cilag, Achillion; Grant/<br />

Research Support: BRISTOL-MEYERS-SQUIBB, ROCHE, GILEAD, BRISTOL-MEY-<br />

ERS-SQUIBB, ROCHE, GILEAD; Speaking and Teaching: Glaxo Smith-Kline,<br />

BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH, ROCHE, NOVARTIS, GILEAD,<br />

VERTEX, Glaxo Smith-Kline, BRISTOL-MEYERS-SQUIBB, SCHERING-PLOUGH,<br />

ROCHE, NOVARTIS, GILEAD, VERTEX, Sanofi<br />

Rafael Esteban - Speaking and Teaching: MSD, BMS, Novartis, Gilead, Glaxo,<br />

MSD, BMS, Novartis, Gilead, Glaxo, Janssen<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Pietro Lampertico - Advisory Committees or Review Panels: Bayer, Bayer; Speaking<br />

and Teaching: Bristol-Myers Squibb, Roche, GlaxoSmithKline, Novartis, Gilead,<br />

Bristol-Myers Squibb, Roche, GlaxoSmithKline, Novartis, Gilead<br />

The following authors have nothing to disclose: Ramazan Idilman, George N.<br />

Dalekos, Heng Chi, José Luís Calleja, Vana Sypsa, Federica Invernizzi, Spyros I.<br />

Siakavellas, Onur Keskin, Nikolaos Gatselis, Bettina E. Hansen, Maria Lehretz,<br />

Juan de la Revilla, Savvoula Savvidou, Anastasia Kourikou, Floriana Facchetti,<br />

Jiannis Vlachogiannakos, Kostas Galanis<br />

2013<br />

Week 24 HBsAg levels predict response to peginterferon<br />

alfa-2b in HBeAg-positive genotype B or C<br />

chronic hepatitis B patients not meeting a week 12 stopping-rule<br />

Milan J. Sonneveld 1 , Jun Cheng 2 , Qing Xie 3 , Bettina E. Hansen 1 ,<br />

Jin-Lin Hou 4 , Fehmi Tabak 5 , Stefan Zeuzem 6 , Harry L. Janssen 1,7 ;<br />

1 Erasmus MC, University Medical Center Rotterdam, Rotterdam,<br />

Netherlands; 2 Capital Medical University, Ditan Hospital, Beijing,<br />

China; 3 Infectious Diseases, Ruijin Hospital, Shanghai, China;<br />

4 Nanfang Hospital, Infectious Diseases, Guangzhou, China;<br />

5 Istanbul Medical School, Hepatology, Istanbul, Turkey; 6 Medical<br />

clinic 1, Goethe University, Frankfurt, Germany; 7 Gastroenterology,<br />

University Health Network, Toronto, ON, Canada<br />

Background & Aims. Serum levels of HBsAg (qHBsAg) predict<br />

response in HBeAg-positive patients with HBV genotypes B<br />

and C treated with peginterferon (PEG-IFN). A stopping-rule<br />

has been proposed based on qHBsAg >20,000 at week<br />

12 or 24. The performance of these stopping-rules has not<br />

been sufficiently validated in patients treated with PEG-IFN<br />

alfa-2b. Similarly, the value of measuring qHBsAg at week 24<br />

in patients not meeting the week 12 stopping-rule is unknown.<br />

Methods. HBeAg-positive patients with HBV genotypes B or<br />

C treated with PEG-IFN alfa-2b monotherapy for one year in<br />

2 global randomized trials (HBV 99-01 and P05170) were<br />

enrolled if they had available qHBsAg and a known therapy<br />

outcome. We assessed (1) the performance of the proposed<br />

stopping-rules based on qHBsAg >20,000 IU/mL and (2)<br />

assessed predictive performance of qHBsAg at week 24 in<br />

patients not meeting the week 12 stopping-rule. Response was<br />

defined as HBeAg loss with HBV DNA 20,000 at week<br />

12 was observed in 53 (26%) patients, of whom 4 (7.5%)<br />

achieved a response (negative predictive value [NPV] 92%).<br />

At week 24, 33 (16%) patients had qHBsAg >20,000, of<br />

whom 0 (0%, NPV 100%) achieved a response (p20,000<br />

IU/mL) at week 12 or 24 of PEG-IFN alfa-2b therapy identifies<br />

genotype B and C patients with a low probability of response.<br />

Among patients not meeting the week 12 stopping-rule, qHBsAg<br />

at week 24 predicts treatment outcome.<br />

Disclosures:<br />

Milan J. Sonneveld - Advisory Committees or Review Panels: Roche; Speaking<br />

and Teaching: Roche, BMS<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen


1190A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

The following authors have nothing to disclose: Jun Cheng, Qing Xie, Bettina E.<br />

Hansen, Jin-Lin Hou, Fehmi Tabak<br />

not achieve significant HBsAg declines early in treatment and<br />

highlights the need for new therapies to more potently inhibit<br />

cccDNA transcription or persistence.<br />

2014<br />

Early Decline in Serum HBsAg at Week 4 and 12<br />

of Therapy with Tenofovir Disoproxil Fumarate and<br />

Pegylated Interferon is Predictive of HBsAg Loss<br />

Henry Lik-Yuen Chan 1 , Sang Hoon Ahn 2 , Wan-Long Chuang 3 ,<br />

Fehmi Tabak 4 , Rajiv Mehta 5 , Joerg Petersen 6 , George Wu 7 , Eduardo<br />

B. Martins 7 , Leland J. Yee 7 , Anuj Gaggar 7 , Mani Subramanian<br />

7 , Seng Gee Lim 8 , Scott Fung 9 , Graham R. Foster 10 , Maria<br />

Buti 11 , Giovanni B. Gaeta 12 , Aric Josun Hui 13 , George V. Papatheodoridis<br />

14 , Robert Flisiak 15 , Patrick Marcellin 16 ; 1 The Chinese<br />

University of Hong Kong, Hong Kong, China; 2 Yonsei University<br />

College of Medicine, Seoul, Korea (the Republic of); 3 Kaohsiung<br />

Medical University Hospital, Kaohsiung Medical University,<br />

Kaohsiung, Taiwan; 4 Istanbul University Cerrahpasa Faculty of<br />

Medicine, Istanbul, Turkey; 5 Liver Clinic, Surat, India; 6 University<br />

of Hamburg, Hamburg, Germany; 7 Gilead Sciences, Inc., Foster<br />

City, CA; 8 Yong Loo Lin School of Medicine, National University of<br />

Singapore, Singapore, Singapore; 9 University of Toronto, Toronto<br />

General Hospital, Toronto, ON, Canada; 10 Queen Marys University<br />

of London, London, United Kingdom; 11 Hospital Universitari<br />

Vall d’Hebron, Barcelona, Spain; 12 Second University of Naples,<br />

Naples, Italy; 13 The Chinese University of Hong Kong, Alice Ho<br />

Miu Ling Nethersole Hospital, Hong Kong, China; 14 Athens University<br />

Medical School, Athens, Greece; 15 Medical University of<br />

Bialystok, Bialystok, Poland; 16 Hôpital Beaujon, University of Paris,<br />

Paris, France<br />

Introduction: Combination treatment with pegylated interferon<br />

(PEG) and tenofovir disproxil fumarate (TDF) has increased<br />

rates of HBsAg loss as compared to either treatment alone.<br />

Here we evaluate the early HBsAg kinetics in patients treated<br />

with monotherapy (either PEG or TDF) or combination treatment<br />

of PEG+TDF. Methods: 740 CHB patients without advanced<br />

disease were randomized to receive (TDF+PEG) x48 weeks<br />

(arm A); (PEF+TDF) x16 weeks followed by TDF x32 weeks<br />

(arm B); TDF (arm C); PEG x48 weeks (arm D). Quantitative<br />

HBsAg was measured (Abbot Architect Assay) at baseline,<br />

Wks 4 and 12. Baseline predictors for HBsAg loss by Wk 72<br />

were evaluated using logistic regression analyses. Results: The<br />

proportion of patients with HBsAg declines of >0.5log 10<br />

, and<br />

>1log 10<br />

, at Wk 4 or 12 were higher among patients receiving<br />

PEG+TDF (arms A and B) as compared to monotherapy with<br />

PEG or TDF. By Wk 4, among patients receiving combination<br />

therapy (N=364), 5.5% had >1log 10<br />

decline and 11%<br />

had >0.5log 10<br />

decline, which increased by Wk 12 to 13.9%<br />

and 28.1% respectively (figure). Multivariate analyses of arms<br />

A+B showed that patients with >0.5log 10<br />

decline at Wk 4 or<br />

>1log 10<br />

decline at Wk 12 were more likely to be Genotype<br />

B and C with elevated ALT and HBV DNA. HBsAg declines in<br />

both monotherapy arms were also associated with similar baseline<br />

virologic characteristics. Within Arm A, sixteen patients<br />

experienced serum HBsAg loss by Wk 72. HBsAg declines<br />

of 1log 10<br />

at Wk 12 had a PPV of 0.31 and 0.43,<br />

respectively. Conclusions: Early kinetics of HBsAg with combination<br />

treatment of TDF and PEG as early as 4 weeks after<br />

treatment can predict eventual HBsAg loss with high NPV and<br />

modest PPV. The majority of patients treated, however, does<br />

Disclosures:<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Joerg Petersen - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Gilead, Novartis, Merck, Bristol-Myers Squibb, Gilead, Novartis, Merck; Grant/<br />

Research Support: Roche, GlaxoSmithKline, Roche, GlaxoSmithKline; Speaking<br />

and Teaching: Abbott, Tibotec, Merck, Abbott, Tibotec, Merck<br />

George Wu - Employment: Gilead Bioscience<br />

Eduardo B. Martins - Employment: Gilead Sciences, Inc.; Stock Shareholder:<br />

Gilead Sciences, Inc.<br />

Leland J. Yee - Employment: Gilead Science<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

Seng Gee Lim - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals, Gilead<br />

Pharmaceuticals, Roche Pharmaceuticals; Speaking and Teaching: Bristol-Myers<br />

Squibb, Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals,<br />

Gilead Pharmaceuticals<br />

Scott Fung - Advisory Committees or Review Panels: Gilead Sciences, Merck,<br />

Vertex, Roche; Speaking and Teaching: BMS, Gilead, AbbVie, Janssen, Merck<br />

Graham R. Foster - Advisory Committees or Review Panels: GlaxoSmithKline,<br />

Novartis, Boehringer Ingelheim, Tibotec, Chughai, Gilead, Janssen, Idenix,<br />

GlaxoSmithKline, Novartis, Roche, Tibotec, Chughai, Gilead, Merck, Janssen,<br />

Idenix, BMS; Board Membership: Boehringer Ingelheim; Grant/Research Support:<br />

Chughai, Roche, Chughai; Speaking and Teaching: Roche, Gilead, Tibotec,<br />

Merck, BMS, Boehringer Ingelheim, Gilead, Janssen<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

Giovanni B. Gaeta - Advisory Committees or Review Panels: Janssen, Merck,<br />

Abbvie, Roche; Speaking and Teaching: BMS, Gilead<br />

George V. Papatheodoridis - Advisory Committees or Review Panels: Merck<br />

Sharp & Dohme, Novartis, Abbvie, Boerhinger Ingelheim, Bristol-Meyer Squibb,<br />

Gilead, Roche, Janssen, GlaxoSmith Kleine; Grant/Research Support: Roche,<br />

Gilead, Bristol-Meyer Squibb, Abbvie, Janssen; Speaking and Teaching: Merck<br />

Sharp & Dohme, Bristol-Meyer Squibb, Gilead, Roche, Janssen, Abbvie<br />

Robert Flisiak - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, Janssen, Merck, Roche, Abbvie; Consulting: Janssen, Bristol Myers<br />

Squibb, Gilead, Merck, Abbvie; Grant/Research Support: Roche; Speaking and<br />

Teaching: Janssen, Roche, Bristol Myers Squibb, Gilead, Merck, Abbvie<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

The following authors have nothing to disclose: Sang Hoon Ahn, Fehmi Tabak,<br />

Rajiv Mehta, Aric Josun Hui


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1191A<br />

2015<br />

Safety and Efficacy of GS-4774 in Patients with Chronic<br />

Hepatitis B on Oral Antiviral Therapy<br />

Anna S. Lok 1 , Calvin Q. Pan 2 , Steven-Huy B. Han 3 , Huy N.<br />

Trinh 4 , Jeffrey Fessel 5 , Tim Rodell 6 , Benedetta Massetto 7 , Anh-Hoa<br />

Nguyen 7 , Anuj Gaggar 7 , Mani Subramanian 7 , John G. McHutchison<br />

7 , Carlo Ferrari 8 , Hannah Lee 9 , Stuart C. Gordon 10 , Edward<br />

J. Gane 11 ; 1 University of Michigan Health System, Ann Arbor,<br />

MI; 2 Medical Procare, PLLC, Flushing, NY; 3 Ronald Reagan UCLA<br />

Medical Center, Los Angeles, CA; 4 San Jose Gastroenterology,<br />

San Jose, CA; 5 Kaiser Permanente, San Francisco, CA; 6 Globeimmune,<br />

Louisville, CO; 7 Gilead Sciences, Inc., Foster City, CA; 8 University<br />

of Parma, Parma, Italy; 9 Tufts Medical Center, Boston, MA;<br />

10 Henry Ford Hospital, Detroit, MI; 11 Auckland General Hospital,<br />

Auckland, New Zealand<br />

Background: Chronic HBV infection is associated with a<br />

dysfunctional T-cell response. GS-4774 is a heat-inactivated<br />

yeast-based therapeutic T-cell vaccine expressing a recombinant<br />

protein containing HBV core, surface, and X proteins to<br />

elicit an HBV-specific T-cell response. This Phase 2 study evaluated<br />

GS-4774 in virally-suppressed chronic hepatitis B (CHB)<br />

patients. Methods: In this multicenter study, 178 non-cirrhotic<br />

patients with CHB were enrolled. All patients were virally suppressed<br />

on an oral antiviral (OAV) for >1 year. Patients were<br />

randomized to continue OAV alone or to continue OAV and<br />

receive GS-4774 every 4 weeks for 6 doses at either 2 yeast<br />

units (YU), 10 YU, or 40 YU [1 YU=10^7 yeast cells]. The<br />

primary endpoint was decline from baseline in quantitative<br />

serum HBsAg (Abbot Architect Assay) at Week 24. Safety<br />

assessments include monitoring of injection site reactions,<br />

hepatic flares, and virologic breakthrough. Results: Baseline<br />

demographics were similar across all groups. The mean age<br />

of patients was 45 to 50 years and 62-74% were men. The<br />

majority (68-80%) of patients was Asian, 24-26% were HBeAg<br />

positive, and the mean duration of prior OAV treatment was<br />

4.4-4.8 years. Grade 3 or 4 adverse events (AEs) occurred in<br />

30%<br />

reductions in HBsAg (observed variability in the OAV arm)<br />

at Week 48 (Panel B) at all timepoints. No patients achieved<br />

HBsAg loss. Five of 32 (16%) HBeAg positive patients treated<br />

with GS-4774 experienced HBeAg loss with 4 having HBeAg<br />

seroconversion. No virologic breakthrough was observed. Conclusions:<br />

GS-4774 was safe and well-tolerated in CHB patients<br />

with no patients achieving HBsAg loss during the study. Modest<br />

reductions in HBsAg were observed in some patients treated<br />

with GS-4774.<br />

Disclosures:<br />

Anna S. Lok - Advisory Committees or Review Panels: Gilead, MYR, Tekmira;<br />

Consulting: GSK, Merck; Grant/Research Support: AbbVie, BMS, Gilead, Idenix<br />

Calvin Q. Pan - Advisory Committees or Review Panels: Gilead; Consulting:<br />

BMS, Gilead, Abbvie, Janssen ; Grant/Research Support: BMS, Gilead, Merck;<br />

Speaking and Teaching: BMS, Gilead, Abbvie<br />

Steven-Huy B. Han - Grant/Research Support: Bristol Myers Squibb, Gilead<br />

Huy N. Trinh - Advisory Committees or Review Panels: BMS, Gilead; Grant/<br />

Research Support: BMS, Gilead; Speaking and Teaching: BMS, Gilead; Stock<br />

Shareholder: Gilead<br />

Jeffrey Fessel - Grant/Research Support: gilead, bms, abbvie, gsk, johnson &<br />

johnson<br />

Benedetta Massetto - Employment: Gilead Sciences, Inc.; Stock Shareholder:<br />

Gilead Sciences, Inc<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Carlo Ferrari - Advisory Committees or Review Panels: Gilead, Roche, Abbvie,<br />

BMS, Merck, Arrowhead; Grant/Research Support: Gilead, Roche, Janssen<br />

Stuart C. Gordon - Advisory Committees or Review Panels: Janssen; Consulting:<br />

Merck, Gilead, BMS, CVS Caremark, Amgen, AbbVie; Grant/Research Support:<br />

Merck, Gilead, AbbVie, Intercept Pharmaceuticals, Exalenz Sciences, Inc., BMS<br />

Edward J. Gane - Advisory Committees or Review Panels: Novira, AbbVie, Janssen,<br />

Gilead Sciences, Janssen Cilag, Achillion, Merck, Tekmira; Speaking and<br />

Teaching: AbbVie, Gilead Sciences, Merck<br />

The following authors have nothing to disclose: Tim Rodell, Anh-Hoa Nguyen,<br />

Hannah Lee<br />

2016<br />

Entecavir and Tenofovir More Effectively Reduce the<br />

Recurrence of Hepatitis B Virus-related Hepatocellular<br />

Carcinoma after Curative Treatment than Low-potent<br />

Antivirals<br />

Hongkeun Ahn 1 , Jeong-Hoon Lee 1 , YOUNG CHANG 1 , Joon Yeul<br />

Nam 1 , Hyeki Cho 1 , Young Youn Cho 1 , Minjong Lee 1 , Jeong-Ju<br />

Yoo 1 , Yuri Cho 1 , Donghyeon Lee 1 , Eun Ju Cho 1 , Su Jong Yu 1 ,<br />

Yoon Jun Kim 1 , Jung-Hwan Yoon 1 , June Sung Lee 2 ; 1 Department of<br />

Internal Medicine and Liver Research Institute, Seoul National University<br />

Hospital, Seoul, Korea (the Republic of); 2 Internal Medicine,<br />

Inje University, Ilsan Paik Hospital, Goyang, Korea (the Republic<br />

of)<br />

Backgroud and Aim: The effect of antiviral potency of nucleos(t)ide<br />

analogues (NAs) on the recurrence of hepatocellular<br />

carcinoma (HCC) in chronic hepatitis B (CHB) patients has<br />

not been well investigated. This study aimed to evaluate the<br />

association of the potency of antivirals and the risk of hepatitis<br />

B virus (HBV)-related HCC recurrence after potentially curative<br />

treatment. Methods: We performed a retrospective analysis of<br />

data from 473 consecutive hepatitis B virus (HBV)-related HCC<br />

patients treated with radio-frequency ablation (RFA, n=261)<br />

or resection (n=212). According to the potency of antiviral<br />

treatments, total population was divided into three groups:<br />

group A (no antiviral treatment, n=224), group B (low-potent<br />

NA: telbivudine, adefovir, clevudine, and lamivudine; n=76),<br />

group C (high-potent NA: entecavir [ETV] and tenofovir disoproxil<br />

fumarate [TDF]; n=173). Recurrence-free survival (RFS)<br />

in three groups were analyzed by Kaplan-Meier curve analysis<br />

and Cox proportional hazards model. Results: The median<br />

follow-up duration was 56.9 months. The median RFS was<br />

38.3, 43.6, and 92.9 months in group A, group B, and group<br />

C, respectively (P=0.012 by log-rank test). In multivariate<br />

analysis, group C had significantly prolonged RFS (hazard<br />

ratio [HR]=0.55, 95% confidence interval [CI]=0.42–0.74;<br />

P


1192A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

cant effect on RFS between groups B and C (HR=1.17, 95%<br />

CI=0.63–2.18; P=0.620). Conclusion: High-potent antiviral<br />

agents (ETV or TDF) reduce the risk of HCC recurrence compared<br />

to low-potent antivirals as well as no antivirals, especially<br />

in patients with high viral titer.<br />

Disclosures:<br />

Yoon Jun Kim - Grant/Research Support: Bristol-Myers Squibb, Roche, JW Creagene,<br />

Bukwang Pharmaceuticals, Handok Pharmaceuticals, Hanmi Pharmaceuticals,<br />

Yuhan Pharmaceuticals; Speaking and Teaching: Bayer HealthCare<br />

Pharmaceuticals, Gilead Science, MSD Korea, Yuhan Pharmaceuticals, Samil<br />

Pharmaceuticals, CJ Pharmaceuticals, Bukwang Pharmaceuticals, Handok Pharmaceuticals<br />

The following authors have nothing to disclose: Hongkeun Ahn, Jeong-Hoon Lee,<br />

YOUNG CHANG, Joon Yeul Nam, Hyeki Cho, Young Youn Cho, Minjong Lee,<br />

Jeong-Ju Yoo, Yuri Cho, Donghyeon Lee, Eun Ju Cho, Su Jong Yu, Jung-Hwan<br />

Yoon, June Sung Lee<br />

2017<br />

Long term follow up of chronic hepatitis B patients after<br />

oral antiviral treatment discontinuation<br />

Seong Hee Kang, Jong Eun Yeon, Young-Sun Lee, Tae Suk Kim,<br />

Yang Jae Yoo, Ji Hoon Kim, Kwan Soo Byun; Gastroenterology &<br />

Hepatology, Korea University Medical Center, Seoul, Korea (the<br />

Republic of)<br />

Introduction: In the current guidelines, 6-12 months of consolidation<br />

after the HBeAg seroconversion is recommended for<br />

HBeAg-positive patients. And the ideal treatment duration of<br />

HBeAg negative chronic hepatitis B (CHB) is not well known.<br />

We evaluated long term clinical courses in patient after antiviral<br />

treatment cessation. Methods: A total of 118 HBeAg-positive<br />

and 60 HBeAg-negative CHB patients who discontinued<br />

lamivudine between 1997 and 2014 were retrospectively analyzed.<br />

Results: In HBeAg-positive patients, the mean age of the<br />

patients at treatment initiation was 32.5 years. The mean duration<br />

of lamivudine treatment in was 32.3 months and the mean<br />

follow-up period after discontinuation was 72.4 months. The<br />

cumulative probability of a composite relapse (≥ 10 4 copies/<br />

mL) at 1, 6, 12, 24, 48, 96, 180 months were 11.9%, 28.8%,<br />

38.1%, 51.7%, 59.3%, 63.6%, 64.4% respectively. The mean<br />

duration of lamivudine treatment were 31.9 months and 32.9<br />

months in the relapse and non-relapse group (p=0.79). The<br />

mean duration of consolidation period after HBeAg seroconversion<br />

were 10.7 months and 11.4 months were not different<br />

(p=0.792). Multivariate analysis revealed pretreatment age ≤<br />

34 years [HR 0.324, 95% CI: 0.148-0.710 p=0.005] and<br />

undetectable HBV DNA [HR 0.261, 95% CI: 0.107-0.638<br />

p=0.003] within the three month after treatment discontinuation<br />

were the predictive factors of non-relapse. In HBeAg-negative<br />

patients, the mean age of the patients was 39.2 years.<br />

The cumulative probability of a composite relapse at 1, 6, 12,<br />

24, 48, 96, 180 months were 25.0%, 33.3%, 35.0%, 41.7%,<br />

43.3%, 46.7%, 48.3% respectively. Mean time to relapse after<br />

off-treatment was 13.6 months. The mean duration of lamivudine<br />

treatment were 29.8 months and 33.5 months in the<br />

relapse and non-relapse group (p=0.447). Conclusions: Even<br />

with the long term consolidation after HBeAg seroconversion,<br />

most HBeAg positive CHB patients showed relapse. And half<br />

of patients with HBeAg-negative CHB had relapsed after treatment<br />

cessation. Antiviral therapy should be maintained for long<br />

term period or until HBsAg was lose.<br />

Disclosures:<br />

Kwan Soo Byun - Grant/Research Support: Gilead, BMS<br />

The following authors have nothing to disclose: Seong Hee Kang, Jong Eun Yeon,<br />

Young-Sun Lee, Tae Suk Kim, Yang Jae Yoo, Ji Hoon Kim<br />

2018<br />

The impact of timely initial injection of two different<br />

doses of hepatitis B virus (HBV) vaccine plus hepatitis B<br />

immunoglobulin (HBIG) on preventing perinatal transmission<br />

of HBV infection<br />

chong wang 1 , Chuan Wang 2 , Jie Li 3 , xing Wu 2 , Fei Kong 1 , Simin<br />

Wen 2 , Jing Jiang 2 , Junqi Niu 1 ; 1 hepatology, The First Hospital of<br />

Jilin University, Changchun, China; 2 The First Hospital of Jilin University,<br />

Changchun, China; 3 Department of microbiology, Beijing<br />

university, Beijing, China<br />

Background & aims: Routine immunoprophylaxis reduces HBV<br />

transmission. However, it does not have individual vaccine<br />

dose to babies whose mothers with different hepatitis B virus e<br />

antigen (HBeAg) status. this study was to investigate the effectiveness<br />

of an immunoprophylaxis protocol with two different<br />

vaccine doses according to mothers’HBeAg(+/-) status and<br />

timely initial injection on blocking mother-to-infant transmission<br />

of HBV, and to evaluate potential risk factors associated with<br />

immunoprophylaxis failure and protective antibody titers. Methods:<br />

A population-based prospective study was conducted from<br />

July 2012 to April 2015. A total of 863 hepatitis B virus surface<br />

antigen(HBsAg)-positive mothers and their 871 infants (8<br />

pairs of twins) were included in the study. The initial injection<br />

of HBV vaccine and HBIG(100IU) were carried out within 2<br />

hours of birth. 20mg and 10mg HBV vaccine were given to<br />

babies born to HBeAg+ and HBeAg- mothers respectively. We<br />

analyze the association of multiple risk factors with immunoprophylaxis<br />

failure and vaccine response. Results: At 7 months,<br />

no immunoprophylaxis failure was observed in 565 babies<br />

born to HBeAg- mothers.Among 306 infants born to HBeAg+<br />

mothers, 16 (5.2%) babies were immunoprophylaxis failure<br />

(HBsAg-positive) at 7 months, with high HBV DNA load ≥8<br />

log10 IU/ml (Crude OR:3.69, 95%CI:1.03-13.24; Adjusted<br />

OR:4.53, 95%CI a :1.19-17.34; P a =0.027) and Non timely<br />

or inadequate initial injection (Crude OR:3.42, 95%CI:1.12-<br />

10.49; Adjusted OR:5.23, 95%CI a :1.54-17.82; P a =0.008)<br />

were strongly associated with immunoprophylaxis failure,<br />

while other factors including delivery mode, feeding pattern,<br />

infant gender and birth weight showed no significant association.<br />

Of 855 babies for whom blocking was successful,<br />

825(96.5%) had produced adequate titers of antibody to<br />

HBsAg (anti-HBs) (≥ 100 mIU/mL). For full-term babies, birth<br />

weight< 3000 g was correlated with low immune response<br />

to vaccination (Crude OR:0.37, 95%CI:0.17-0.84; Adjusted<br />

OR:0.37, 95%CI a :0.16-0.84; P a =0.018). Conclusions: Application<br />

of timely initial injection of two different doses of HBV<br />

vaccine plus HBIG to infants born to mothers with different<br />

HBeAg status is effective for preventing perinatal transmission<br />

of HBV, particularly for those whose mothers are HBeAg-negative.<br />

The overwhelming majority of babies produced adequate<br />

levels of protective anti-HBs titers (>100 mIU/mL) after immunoprophylaxis.<br />

high HBV DNA load, and non-timely or inadequate<br />

initial injection were associated with immunoprophylaxis<br />

failure. Relative low birth weight of full-term babies might be<br />

correlated with weak immune response to vaccination.<br />

Disclosures:<br />

The following authors have nothing to disclose: chong wang, Chuan Wang, Jie<br />

Li, xing Wu, Fei Kong, Simin Wen, Jing Jiang, Junqi Niu


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1193A<br />

2019<br />

Discontinuation of Effective Long-Term Nucleos(t)ide<br />

Analogue(s) [NA(s)] Treatment in HBeAg-Negative<br />

Chronic Hepatitis B (CHBe-) Patients: A Multicenter<br />

Ongoing Prospective Study<br />

Spilios Manolakopoulos 1 , Hariklia Kranidioti 1 , Anastasia<br />

Kourikou 1 , Emilia Hadziyannis 1 , Georgios Kontos 1 , Melanie<br />

Deutsch 1 , Alexandra Alexopoulou 1 , Emanuel K. Manesis 1 , George<br />

V. Papatheodoridis 2,1 ; 1 2nd Department of Internal Medicine, University<br />

of Athens, Hippokration General Hospital, Athens, Greece;<br />

2 Department of Gastroenterology, University of Athens, Laiko General<br />

Hospital, Athens, Greece<br />

Background/Aim The durability of response to NA(s) after<br />

their cessation, has not been adequately investigated in<br />

CHBe- patients. This prospective study aims to investigate the<br />

outcome of CHBe- patients who discontinue NA(s) after longterm<br />

virological suppression and to explore factors that may<br />

predict relapse. Methods Non-cirrhotic CHBe- patients with<br />

virological remission (undetectable HBV-DNA) under NA(s)<br />

for ≥4 years who could remain under close follow-up (every<br />

month for the first 3 months and every 3 months thereafter)<br />

were invited to consent to treatment discontinuation. Τhe criteria<br />

for NA(s) retreatment were: ALT>2xULN & Bilirubin>2mg/<br />

dl, ALT>10xULN, ALT>5xULN in 2 monthly determinations,<br />

ALT>3xULN and HBV-DNA>100,000IU/mL, ALT>ULN & HBV-<br />

DNA>2,000IU/mL for ≥6 months. Results Until May 2015,<br />

43 patients [males: 74%, median age: 60 (22-86) ys, duration<br />

of NA(s): 7.9 (4-14) ys] had >6 months follow-up [median: 27<br />

(6-45) months] and were included in the analysis. Off-treatment<br />

virological relapse (HBV-DNA >2000IU/mL) was developed<br />

in 40 (93%) and biochemical relapse (ALT >ULN) in 29 (67%)<br />

patients. Ten patients (23%) experienced hepatitis flare (ALT<br />

>10 ULN) at a median of 3 (1-9) months after NA(s) discontinuation,<br />

without evidence of liver decompensation. Retreatment<br />

was required in 18 (42%) patients at a median of 6.5 (1-33)<br />

months. Of the 25 patients who remained untreated, 14 (56%)<br />

had ALT persistently


1194A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Tetsuya Hosaka, Masahiro<br />

Kobayashi, Shunichiro Fujiyama, Hideo Kunimoto, Yushi Sorin, Yusuke<br />

Kawamura, Hitomi Sezaki, Satoshi Saitoh, Yasuji Arase, Mariko Kobayashi<br />

2021<br />

Antiviral Activity of Tenofovir Alafenamide (TAF) Against<br />

Drug Resistant HBV Isolates In Vitro<br />

Yang Liu, Ben C. Mitchell, Phillip Dinh, Michael D. Miller, Kathryn<br />

M. Kitrinos; Gilead Sciences, Foster City, CA<br />

Background: Tenofovir alafenamide (TAF) is a next generation<br />

prodrug of tenofovir that is being evaluated for the treatment<br />

of hepatitis B virus (HBV) and human immunodeficiency virus<br />

(HIV) infections. The aim of this study was to evaluate the in<br />

vitro antiviral activity of TAF against adefovir-resistant (ADV-<br />

R), lamivudine-resistant (LAM-R), and entecavir-resistant (ETV-<br />

R) isolates. Methods: A full length genotype D HBV genome<br />

was amplified from a treatment-naïve subject and cloned<br />

into an expression vector under the direction of a CMV promoter.<br />

Eleven site-directed mutants were created: 5 ADV-R<br />

(rtA181T/sW172*, rtA181T/sW172L, rtA181V, rtN236T,<br />

rtA181V+N236T), 3 LAM-R (rtM204I, rtL180M+rtM204V,<br />

rtV173L+rtL180M+rtM204V), and 3 ETV-R (rtL180M+rtM-<br />

204V+rtT184G, rtL180M+rtM204V+rtS202G, rtL180M+rt-<br />

M204V+rtM250V) mutants. All constructs were transiently<br />

transfected into HepG2 cells and treated with TAF and tenofovir<br />

(TFV). Drug susceptibility was assessed using quantitative<br />

PCR of intracellular HBV DNA. Fold change (FC) values were<br />

generated by comparing the average EC 50<br />

values obtained<br />

for the mutant viruses to the wild-type virus. Results: Transient<br />

transfection of all constructs resulted in efficient HBV DNA replication<br />

in HepG2 cells. The TAF susceptibility for all the drug<br />

resistant isolates demonstrated mean EC 50<br />

values ranging from<br />

85.4 to 364.7 nM, while the wild-type TAF EC 50<br />

was 99.1<br />

nM. The mean FC values for the drug resistant isolates ranged<br />

from 0.9-3.7, with all but one isolate remaining sensitive to TAF<br />

(FC < 2). The ADV-R isolate rtA181V+rtN236T demonstrated<br />

low-level TAF reduced susceptibility (FC = 3.7). The TFV susceptibility<br />

for all the drug resistant isolates demonstrated mean<br />

EC 50<br />

values ranging from 7.8 to 40.5 mM, the mean FC values<br />

for the drug resistant isolates ranged from 0.6 to 2.8, with all<br />

but one isolate remaining sensitive to TFV (FC < 2). The ADV-R<br />

isolate rtA181V+rtN236T demonstrated low-level reduced susceptibility<br />

to TFV (FC = 2.8). Despite the difference in EC 50<br />

values,<br />

the FC values obtained for each isolate with TAF and TFV<br />

were consistent. Conclusions: All the ADV-R, LAM-R and ETV-R<br />

mutant isolates were sensitive to TAF with the exception of the<br />

ADV-R double mutant rtA181V+rtN236T, which exhibited low<br />

level reduced susceptibility to TAF. The FC values observed<br />

with TAF are similar to what has been previously observed with<br />

TFV. These data supports the on-going TAF phase 3 clinical<br />

trial program in treatment naive and treatment experienced<br />

HBV patients.<br />

Disclosures:<br />

Yang Liu - Employment: Gilead Sciences<br />

Phillip Dinh - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Michael D. Miller - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Kathryn M. Kitrinos - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

The following authors have nothing to disclose: Ben C. Mitchell<br />

2022<br />

Monthly dosing of ARC-520 in chronically hepatitis B<br />

virus infected chimpanzees produces rapid, deep and<br />

durable reductions in circulating viral antigens<br />

Christine I. Wooddell 1 , Deborah Chavez 3 , Jason E. Goetzmann 2 ,<br />

Lena M. Notvall-Elkey 3 , Helen Lee 3 , Bernadette Guerra 3 , Ryan M.<br />

Peterson 1 , Courtney N. Johnson 3 , Julia O. Hegge 1 , Robert Gish 4 ,<br />

Stephen Locarnini 5 , Christopher R. Anzalone 1 , Robert E. Lanford 3 ,<br />

David L. Lewis 1 ; 1 Arrowhead Madison, Arrowhead Research Corporation,<br />

Madison, WI; 2 New Iberia Research Center, University<br />

of Louisiana at Lafayette, New Iberia, LA; 3 Texas Biomedical<br />

Research Institute, San Antonio, TX; 4 Department of Medicine,<br />

Stanford University Medical Center, San Diego, CA; 5 Victorian<br />

Infectious Diseases Reference Laboratory, Melbourne, VIC, Australia<br />

Background: New therapeutic strategies are needed for<br />

chronic HBV infection. RNAi-based therapeutic ARC-520 that<br />

was designed to degrade viral transcripts and reduce HBV<br />

antigens and pgRNA is in clinical development. This study<br />

assessed multiple doses of ARC-520 in HBV infected chimps.<br />

Methods: 9 chimps (5 M, 4 F; 9-37 yrs) with chronic HBV were<br />

treated with NUCs for 8-24 weeks prior to receiving ARC-<br />

520. Following the NUC lead-in, ARC-520 was added and<br />

dosed once every 4 weeks (Q4W) for 6-11 doses. 5 chimps<br />

were HBeAg-positive (HBeAg+) and 4 were HBeAg-negative<br />

(HBeAg-) at start of study. Chimps received humane care as<br />

outlined in the Guide for the Care and Use of Laboratory Animals.<br />

Results: Prior to treatment, HBV serum DNA levels ranged<br />

from 1.1E+7 - 7.1E+8 IU/mL in HBeAg+ chimps and 20-660<br />

IU/mL in HBeAg- chimps. HBsAg levels ranged from 250-3190<br />

μg/mL in HBeAg+ chimps and from 1.2–200 μg/mL in HBeAgchimps.<br />

After the NUC lead-in, DNA levels were reduced to<br />

≤10 3 IU/mL in HBeAg+ chimps and to undetectable levels in<br />

HBeAg- chimps. HBsAg levels were only marginally decreased.<br />

All 9 chimps were then treated with 2, 3 or 4 mg/kg ARC-520.<br />

All procedures and treatments were well tolerated. Treatment<br />

initiation with ARC-520 resulted in strongly reduced viral antigens<br />

with the magnitude of the response in HBeAg+ chimps<br />

being greater than in HBeAg- chimps. Following the first injection<br />

of ARC-520 in HBeAg+ chimps, HBsAg was reduced at<br />

nadir by 0.5-1.4 log 10<br />

, while in HBeAg- chimps HBsAg was<br />

reduced by 0.4–0.7 log 10<br />

. Doses of 2, 3 and 4 mg/kg had<br />

similar efficacy. Levels of HBsAg tended to decrease upon further<br />

dosing. Following the final dose, HBsAg showed mean<br />

reduction at nadir of 1.7 ± 0.5 log 10<br />

in HBeAg+ chimps and<br />

0.7 ± 0.1 log 10<br />

in HBeAg- chimps. HBeAg in HBeAg+ chimps<br />

was reduced by 1.5– 2.0 log 10<br />

. Two HBeAg+ chimps seroconverted<br />

to HBeAg-, anti-HBe positive. Four months after all therapies<br />

ended, one of these has maintained 0.3 log 10<br />

reduction in<br />

HBsAg and greater than 2.0 log 10<br />

reduced serum DNA relative<br />

to baseline; the other has maintained 1.4 log 10<br />

reduced HBsAg<br />

and greater than 4.0 log 10<br />

reduced serum DNA. The latter had<br />

an on-treatment ALT flare that resolved. Conclusions: ARC-520<br />

administered Q4W on a background of daily NUCs was well<br />

tolerated and produced large reductions in circulating HBV<br />

antigens in chimpanzees. Responses were greater in HBeAg+<br />

chimps compared to HBeAg-; 2/5 HBeAg+ chimps serocleared<br />

HBeAg. ARC-520 administered to HBV infected chimpanzees<br />

produces rapid, deep and durable reductions in circulating<br />

HBsAg and HBeAg when dosed over a short defined period.<br />

Disclosures:<br />

Christine I. Wooddell - Employment: Arrowhead Research Corporation<br />

Ryan M. Peterson - Employment: Arrowhead Research Corporation<br />

Julia O. Hegge - Employment: Arrowhead Research Corp


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1195A<br />

Robert Gish - Advisory Committees or Review Panels: Gilead, AbbVie, Arrowhead;<br />

Consulting: Eiger, Isis, Genentech; Speaking and Teaching: Gilead, Abb-<br />

Vie; Stock Shareholder: Arrowhead<br />

Stephen Locarnini - Consulting: Gilead, Arrowhead; Employment: Melbourne<br />

Health<br />

Robert E. Lanford - Grant/Research Support: Arrowhead Research<br />

David L. Lewis - Employment: Arrowhead Research Corporation<br />

The following authors have nothing to disclose: Deborah Chavez, Jason E. Goetzmann,<br />

Lena M. Notvall-Elkey, Helen Lee, Bernadette Guerra, Courtney N. Johnson,<br />

Christopher R. Anzalone<br />

2023<br />

Efficacy and Safety of Peginterferon alfa-2b (40kD,<br />

Y Shape) (Pegberon) in HBeAg- positive CHB Patients<br />

—- A Phase III, Randomized, Multi-center, Positive-Controlled,<br />

Open-label Clinical Study<br />

Gui-Qiang Wang 1 , Feng-Qin Hou 1 , Wei Lu 2 , Jia Shang 3 , Guo-<br />

Zhong Gong 4 , Chen Pan 5 , Ming-Xiang Zhang 6 , Chi-Biao Yin 7 ,<br />

Qing Xie 8 , Yan-Zhong Peng 9 , Shi-Jun Chen 10 ; 1 Peking University<br />

First Hospital, Beijing, China; 2 Xiamen Amoytop Biotech Co.,Ltd,<br />

Xiamen, China; 3 Henan Provincial People’s Hospital, Zhengzhou,<br />

China; 4 The Second Xiangya Hospital of Central South University,<br />

Changsha, China; 5 Mengchao Hepatobiliary Hospital of Fujian<br />

Medical University, Fuzhou, China; 6 The Sixth People’s Hospital<br />

of Shenyang, Shenyang, China; 7 Guangzhou Eighth People’s<br />

Hospital, Guangzhou, China; 8 Rui Jin Hospital, Shanghai, China;<br />

9 Peking University Shenzhen Hospital, Shenzhen, China; 10 Jinan<br />

Infectious Diseases Hospital, Jinan, China<br />

Background: Chronic hepatitis B (CHB) is a life-threatening<br />

liver disease caused by the hepatitis B virus in the world including<br />

China. Peginterferon is one of the first-line drugs recommended<br />

for CHB in guidelines. Thusit is necessary to develop<br />

new peginterferon in China and construct higher-level basis of<br />

evidence-based medicine in Chinese CHB patients. Objective:<br />

Evaluate efficacy and safety of Pegberon (180μg), that developed<br />

in Chinese domestic company with gobal patent 40kD<br />

Y-shape peglated interferon α-2b, by the head-to-head, randomized,<br />

multi-center, positive controlled and open label clinical<br />

study with PEG IFNα-2a (180μg). Provide a higher level<br />

basis of evidence-based medicine for peginterferon treatment.<br />

Methods: Open-label Phase III study enrolled 538 patients for<br />

Pegberon group and 282 patients for PEG IFNα-2a group<br />

with 48 weeks treatment and 24 weeks follow up. The proportion<br />

of patients with HBeAg seroconversion and the safety<br />

were evaluated. HBV DNA analyzed by COBAS ® , Version<br />

2.0. Serum markers including quantitative HBsAg and HBsAb<br />

were analyzed by cobas e 411 analyzer, and HBeAg was<br />

analyzed by Elecsys HBeAg, HBeAb was analyzed by Elecsys<br />

anti-HBe. Results: Pegberon group and PEG IFNα-2a group<br />

had similar baseline virological and serological characteristics.<br />

30.82%(102/331) patients in Pegberon group and<br />

27.27%(48/176) patients in PEG IFNα-2a group achieved<br />

HBeAg seroconversion respectively in the 507 patients that<br />

completed the 72 weeks clinical trial at present. In intent-to<br />

treat populationsboth groups had similar trends of HBV DNA,<br />

HBeAg and HBsAg during 48-week treatment. At 48w, HBV<br />

DNA negative rate in Pegberon group was 11.90%(64/538)<br />

while 9.93%(28/282) in PEG IFNα-2a group. In addition,<br />

rate of HBV DNA10 6 IU/mL enrolled were<br />

treated with antiviral therapy from 24-32 weeks of gestation<br />

using NUCs,except for LDT+TDF at the first trimester, including<br />

330 in LdT group, 25 in TDF group, 27 in TDF+LdT group due<br />

to LdT resistance, respectively. 171 cases who were unwilling<br />

to take antiviral drugs served as controls. All infants were<br />

vaccinated with recombinant HBV vaccine and hepatitis B<br />

immune globulin (HBIG) according to standard immunoprophylaxis<br />

procedure. MTCT rate was determined by HBV markers<br />

including HBsAg and HBV DNA detection in 6 months after<br />

birth. Results: Significantly lower HBV DNA levels were noted<br />

in all the enrolled mothers when delivery. The rate of HBsAg<br />

positive and HBV DNA detectable in infants at 6 months was<br />

0% in each group. The incidence of undetectable cord blood<br />

HBV DNA levels has no significant difference among the three<br />

groups (LdT group, 99.2%; TDF group, 100%, LdT+TDF group,<br />

100%, P>0.05),which were significantly higher than the controls(61.5%),No<br />

severe adverse events or complications were<br />

observed in all the mothers and infants. Conclusions: LdT and<br />

TDF monotherapy or combotherapy were all effective and<br />

well-tolerated in HBeAg positive high viral load mothers and<br />

their infants on short term follow up, and these were associated<br />

with significant reduction of MTCT.<br />

Disclosures:<br />

The following authors have nothing to disclose: yanqiong zhang, Hongfei Huang,<br />

Quanxin Wu, Yuming Wang


1196A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2025<br />

Descriptive Comparison of Efficacy and Safety of Tenofovir<br />

Disoproxil Fumarate (TDF) plus Pegylated-Interferon<br />

α-2a (PEG) Combination in Asians and non-Asians<br />

with Chronic Hepatitis B (CHB)<br />

Patrick Marcellin 2 , Sang Hoon Ahn 3 , Won Young Tak 4 , Owen<br />

Tak Yin Tsang 8 , Aric Josun Hui 5 , Xiaoli Ma 6 , Wan-Long Chuang 7 ,<br />

Seng Gee Lim 9 , Rajiv Mehta 10 , Eduardo B. Martins 1 , Leland J.<br />

Yee 1 , Phillip Dinh 1 , Lanjia Lin 1 , Prista Charuworn 1 , Mani Subramanian<br />

1 , Scott Fung 11 , Magdy Elkhashab 12 , Huy N. Trinh 13 , Henry<br />

Lik-Yuen Chan 14 ; 1 Gilead Sciences, Foster City, CA; 2 Hôpital<br />

Beaujon, Université Paris-Diderot, Clichy, France; 3 Yonsei University<br />

College of Medicine, Seoul, Korea (the Republic of); 4 Kyungpook<br />

National University Hospital, Daegu, Korea (the Republic<br />

of); 5 Alice Ho Miu Ling Nethersole Hospital, Hong Kong, China;<br />

6 Drexel University College of Medicine, Philadelphia, PA; 7 Kaohsiung<br />

Medical University Chung-Ho Memorial Hospital, Kaohsiung,<br />

Taiwan; 8 Princess Margaret Hospital, Hong Kong, China; 9 Yong<br />

Loo Lin School of Medicine, National University of Singapore, Singapore,<br />

Singapore; 10 Liver Clinic, Surat, India; 11 Toronto General<br />

Hospital, Toronto, ON, Canada; 12 Toronto Liver Centre, Toronto,<br />

ON, Canada; 13 San Jose Gastroenterology, San Jose, CA; 14 The<br />

Chinese University of Hong Kong, Hong Kong, China<br />

Background: Differences in viral and patient characteristics<br />

between Asians and non-Asians may impact response to HBV<br />

therapy. We examined differences in between Asian and<br />

non-Asian patients enrolled in the TDF plus PEG combination<br />

study GS-US-174-0149. Methods: 740 CHB patients without<br />

advanced liver disease were randomized 1:1:1:1 to receive<br />

TDF+PEG x48 weeks (Arm A); TDF+PEG x16 weeks followed<br />

by TDF x32 weeks (Arm B); continuous TDF (Arm C); PEG x48<br />

weeks (Arm D). Baseline characteristics, Week 72 efficacy and<br />

overall safety were compared between Asians and non-Asians.<br />

Primary efficacy was HBsAg loss at Week 72. Results: Of 554<br />

Asian patients randomized, most were male and HBeAg-positive,<br />

and predominantly genotype (Gt)-B/C; while among<br />

186 non-Asian patients, most were male, HBeAg-negative and<br />

predominantly Gt- A/ D (Table 1). Overall, rates of HBsAg<br />

decline ≥ 1 log 10<br />

IU/mL from baseline to Week 48 were lower<br />

among non-Asians (11%) vs. Asians (26%) (p0.05 for<br />

Asian vs non-Asian comparison within each treatment group).<br />

Within each treatment arm, no significant differences were<br />

observed for rates of normal ALT, HBeAg loss/seroconversion<br />

between Asian and non-Asian patients at Week 48. Rates of<br />

treatment-emergent AEs, discontinuations and SAEs were similar<br />

between Asians and non-Asians. Conclusion: Although more<br />

Asians had >1log 10<br />

HBsAg decline from baseline, no statistically<br />

significant differences in efficacy were observed between<br />

Asians and non-Asians.<br />

Table 1<br />

*Among 497 Asian and 162 Non-Asians who reached Week<br />

48.<br />

N.S.=Not Significant<br />

Disclosures:<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Won Young Tak - Advisory Committees or Review Panels: Gilead Korea; Grant/<br />

Research Support: SAMIL Pharma; Speaking and Teaching: Bayer Korea<br />

Xiaoli Ma - Consulting: Gilead Sciemces, Inc<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Seng Gee Lim - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals, Gilead<br />

Pharmaceuticals, Roche Pharmaceuticals; Speaking and Teaching: Bristol-Myers<br />

Squibb, Novartis Pharmaceuticals, Merck Sharp and Dohme Pharmaceuticals,<br />

Gilead Pharmaceuticals<br />

Eduardo B. Martins - Employment: Gilead Sciences, Inc.; Stock Shareholder:<br />

Gilead Sciences, Inc.<br />

Leland J. Yee - Employment: Gilead Science<br />

Phillip Dinh - Employment: Gilead Sciences; Stock Shareholder: Gilead Sciences<br />

Lanjia Lin - Employment: Gilead; Stock Shareholder: Gilead<br />

Prista Charuworn - Stock Shareholder: Gilead Sciences<br />

Mani Subramanian - Employment: Gilead Sciences<br />

Scott Fung - Advisory Committees or Review Panels: Gilead Sciences, Merck,<br />

Vertex, Roche; Speaking and Teaching: BMS, Gilead, AbbVie, Janssen, Merck<br />

Magdy Elkhashab - Advisory Committees or Review Panels: GSK, Gilead, Merk,<br />

BMS, ABBVIE; Grant/Research Support: EISAI, GENENTECH, GSK, GILEAD,<br />

ABBVIE, Merk, BMS<br />

Huy N. Trinh - Advisory Committees or Review Panels: BMS, Gilead; Grant/<br />

Research Support: BMS, Gilead; Speaking and Teaching: BMS, Gilead; Stock<br />

Shareholder: Gilead<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

The following authors have nothing to disclose: Sang Hoon Ahn, Owen Tak Yin<br />

Tsang, Aric Josun Hui, Rajiv Mehta<br />

2026<br />

Characterization of Hepatitis B Surface Proteins in<br />

HBeAg-positive Patients with Chronic Hepatitis B (CHB)<br />

Treated With PegInterferon Alfa-2a (40KD)<br />

Franziska Rinker 1,2 , Corinna M. Bremer 3,4 , Steffen B. Wiegand 1 ,<br />

Birgit Bremer 1 , Michael P. Manns 1,2 , Heiner Wedemeyer 1,2 , Lei<br />

Yang 5 , Vedran Pavlovic 6 , Cynthia Wat 6 , Dieter Glebe 3,4 , Anke R.<br />

Kraft 1 , Markus Cornberg 1,2 ; 1 Gastroenterology, Hepatology and<br />

Endocrinology, Medical School Hannover, Hannover, Germany;<br />

2 German Center for Infection Research (DZIF), Hannover, Germany;<br />

3 Institute of Medical Virology, National Reference Center<br />

for Hepatitis B and D Viruses, Justus Liebig University Giessen,<br />

Giessen, Germany; 4 German Center for Infection Research (DZIF),<br />

Giessen, Germany; 5 Roche Products Ltd, Shanghai, China; 6 Roche<br />

Products Ltd, Welwyn Garden City, United Kingdom<br />

Background: HBsAg consists of three proteins called large<br />

(L-), middle (M-) and small (S-)HBs that differ in amino-terminal<br />

sequences and glycosylation status. Commercial HBsAg assays<br />

only quantify the total amount of HBsAg. We have validated<br />

an ELISA method for quantification of HBs proteins and demonstrated<br />

that SHBs represents the commonest subtype (accounting<br />

for 88% of total HBsAg, L- and M- HBs account for 8%<br />

and 3%) in untreated HBeAg-positive patients. The aim of current<br />

study was to explore HBs protein fractions kinetics during<br />

pegylated interferon alfa-2a (PegIFNα-2a, Pegasys) therapy<br />

and examine whether HBs fractions are more predictive of<br />

treatment response than quantitative HBsAg (qHBsAg) levels.<br />

Methods: Serum samples of 128 HBeAg-positive patients, from<br />

two phase III/IV PegIFNα-2a trials were retrospectively analysed.<br />

Patients received PegIFNα-2a ± lamivudine therapy for<br />

48 weeks. Clinical and laboratory patient data from completed<br />

trials was provided by Roche. HBs fractions were quantified


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1197A<br />

using a quantitative ELISA and well-defined monoclonal antibodies<br />

against the S-domain (Total_HBs), the preS1-domain<br />

(LHBs) and the N-glycosylated preS2-domain (MHBs) of the HBs<br />

protein. SHBs levels were derived by subtracting L- and M- HBs<br />

from Total_HBs. Data was log-transformed (log 10<br />

x+1) prior to<br />

analysis. Response was defined as HBeAg seroconversion 24<br />

weeks post-treatment. Results: During treatment, mean HBs fractions<br />

and qHBsAg levels were consistently lower in responders<br />

than non-responders. HBs fractions declined by approximately<br />

1 log by Week 24 (range 0.9-1.2), which was comparable<br />

to the 0.9 log decline observed in qHBsAg levels. Receiver<br />

operating characteristics (ROC) analysis demonstrated that<br />

HBV-DNA, qHBsAg and qHBeAg (but not ALT) levels can predict<br />

treatment response moderately well with area under the<br />

ROC curve (AUROC) scores of 0.70-0.86 (see Table). AUROC<br />

scores for HBs fractions (range 0.70-0.75) were comparable<br />

to qHBsAg scores (0.70-0.74) at each time-point. Conclusions:<br />

The results demonstrate that during PegIFNα-2a therapy, HBs<br />

fractions are lower in responders than non-responders, which<br />

is consistent with the pattern observed with qHBsAg levels. The<br />

ROC analysis demonstrates that individual HBs fraction levels<br />

are equivalent and not superior to qHBsAg levels, for predicting<br />

treatment response in HBeAg-positive patients.<br />

AUROC Scores at Baseline, Week 12 and 24<br />

2027<br />

Liver Gene Expression Profiles Correlate with Virus<br />

Infection and Response to Interferon Therapy in Chronic<br />

Hepatitis B Patients<br />

Chun-Jen Liu, Pei-Jer Chen, Ding-Shinn Chen, Jia-Horng Kao, Hui-<br />

Lin Wu; Graduate Institute of Clinical Medicine, National Taiwan<br />

University College of Medicine, Taipei City, Taiwan<br />

Background: The natural course of chronic hepatitis B (CHB)<br />

infection and treatment response are determined mainly by the<br />

genomic characteristics of the individual. We investigated liver<br />

gene expression profiles to reveal the molecular basis associated<br />

with viral infection and IFN-alpha treatment response<br />

in CHB patients. Methods: Nineteen IFN-alpha responders<br />

and 19 non-responders with CHB were included in this study.<br />

Expression profiles were compared between paired liver<br />

biopsy samples taken before and 6 months after successful<br />

IFN-alpha treatment or between pretreatment biopsy samples<br />

of IFN-alpha responders and non-responders. Results: A<br />

total of 132 differentially up-regulated and 39 differentially<br />

down-regulated genes were identified in the pretreatment livers<br />

of CHB patients. The up-regulated genes were mainly related<br />

to immune response and cell proliferation. IFN-alpha and B cell<br />

signatures were also enriched. Lower intrahepatic HBV pregenomic<br />

RNA levels and a total of 118 differentially up-regulated<br />

and 33 differentially down-regulated genes were identified in<br />

IFN-alpha responders. The up-regulated gene set in responders<br />

significantly overlapped with the up-regulated genes associated<br />

with the pretreatment livers of CHB patients. In contrast,<br />

no clear pattern was found in the down-regulated genes of<br />

both analyses. Results of the bioinformatic analyses were validated<br />

in independent cohorts and relevant animal models.<br />

Conclusions: CHB infection evokes significant gene expression<br />

associated with immune responses. The up-regulated genes are<br />

predictive of responsiveness to IFN-alpha therapy, as are lower<br />

intrahepatic levels of HBV pregenomic RNA and pre-activated<br />

host immune responses. Our findings provide novel insights<br />

into the immune mechanism and manipulation of CHB.<br />

Disclosures:<br />

Pei-Jer Chen - Advisory Committees or Review Panels: BMS, GSK, BMS, GSK,<br />

Medigene; Independent Contractor: J & J; Speaking and Teaching: Roche, Roche<br />

The following authors have nothing to disclose: Chun-Jen Liu, Ding-Shinn Chen,<br />

Jia-Horng Kao, Hui-Lin Wu<br />

*HBV-DNA analyses performed on PEgIFN monotherapy cohort<br />

Disclosures:<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

Heiner Wedemeyer - Advisory Committees or Review Panels: Transgene, MSD,<br />

Roche, Gilead, Abbott, BMS, Falk, Abbvie, Novartis, GSK, Roche Diagnostics;<br />

Grant/Research Support: MSD, Novartis, Gilead, Roche, Abbott, Roche Diagnostics;<br />

Speaking and Teaching: BMS, MSD, Novartis, ITF, Abbvie, Gilead<br />

Lei Yang - Employment: Roche (China) Holding Ltd.<br />

Vedran Pavlovic - Employment: Roche Products Ltd; Stock Shareholder: Roche<br />

Products Ltd<br />

Cynthia Wat - Employment: Roche Products Ltd<br />

Markus Cornberg - Advisory Committees or Review Panels: Merck (MSD Germamny),<br />

Roche, Gilead, Novartis, Abbvie, Janssen Cilag, BMS; Grant/Research<br />

Support: Merck (MSD Germamny), Roche; Speaking and Teaching: Merck (MSD<br />

Germamny), Roche, Gilead, BMS, Novartis, Falk, Abbvie<br />

The following authors have nothing to disclose: Franziska Rinker, Corinna M.<br />

Bremer, Steffen B. Wiegand, Birgit Bremer, Dieter Glebe, Anke R. Kraft<br />

2028<br />

A study to evaluate the dynamics changes of HBsAg<br />

quantity and its relation with HBeAg seroconversion following<br />

48 weeks pegylated-interferon-alpha treatment<br />

in patients with HBeAg positive chronic hepatitis B after<br />

long term nucleos(t)ide analogue maintenance therapy<br />

(Roll Over trial) : Interim analysis at 24 weeks<br />

Jeong Heo 1 , Hyun Young Woo 1 , Won Young Tak 2 , Soo Young<br />

Park 2 , Won Lim 1 , Youngmi Hong 1 , Young Oh Kweon 2 , Ki Tae<br />

Yoon 1 , Mong Cho 1 ; 1 Pusan National University Hospital, Pusan,<br />

Korea (the Republic of); 2 Kyungpook National University School of<br />

Medicine, Daegu, Korea (the Republic of)<br />

Background & Aims: Durable post-treatment response is uncommon<br />

in patients with chronic hepatitis B (CHB) on nucleos(t)<br />

ide analogue (NA) therapy. The aim of this study is to investigate<br />

pegylated interferon (PI) after long term NA therapy might<br />

potentiate the antiviral efficacy directly via its effect on broad<br />

antiviral activities and indirectly via activation of innate and<br />

adaptive immune responses leading to HBeAg seroconversion<br />

and eventually HBsAg loss and/or seroconversion. Methods:<br />

The patient with HBeAg-positive CHB who had been treated<br />

with any NA except telbivudine, who have an undetectable<br />

HBV DNA (


1198A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

However, none of these on treatment HBV DNA elevation was<br />

accompanied by elevated level of alanine aminotransferase<br />

(ALT). The level of baseline ALT showed association with HBeAg<br />

seroconversion (p=0.075). PI alfa-2a was well-tolerated. Conclusions:<br />

This interim analysis showed that, for patients who<br />

achieve virological suppression with oral NA, switching to a<br />

24 weeks PI alfa-2a treatment significantly decrease HBsAg<br />

titer and increases rates of HBeAg seroconversion.<br />

Disclosures:<br />

Jeong Heo - Advisory Committees or Review Panels: Abbvie; Grant/Research<br />

Support: Roche<br />

Won Young Tak - Advisory Committees or Review Panels: Gilead Korea; Grant/<br />

Research Support: SAMIL Pharma; Speaking and Teaching: Bayer Korea<br />

Ki Tae Yoon - Grant/Research Support: Handok; Speaking and Teaching: Gilead,<br />

BMS, MSD, Roche, GSK<br />

The following authors have nothing to disclose: Hyun Young Woo, Soo Young<br />

Park, Won Lim, Youngmi Hong, Young Oh Kweon, Mong Cho<br />

2029<br />

Reduction of HBsAg by peg-interferon given sequentially<br />

after long term nucleot(s)ide analogue therapy:<br />

nationwide multicenter study by Japanese Red Cross<br />

Liver Study Group.<br />

Masayuki Kurosaki, Namiki Izumi; Department of Gastroenterology<br />

and Hepatology, Musashino Red Cross Hospital, Tokyo, Japan<br />

Background and Aims: The ideal endpoint of antiviral therapy<br />

for chronic hepatitis B is loss of HBs antigen (HBsAg). However<br />

HBsAg loss is rarely achieved by nucleot(s)ide analogues<br />

(NUC). The aim of the present study is to elucidate the degree<br />

of HBsAg reduction and incidence of HBsAg loss by peg-interferon<br />

when given sequentially after long term NUC therapy.<br />

Methods: A total of 54 patients were sequentially treated by 48<br />

weeks of peg-interferon alpha-2a (180 microgram weekly) after<br />

long-term NUC therapy with an overlap of 4 weeks (sequential<br />

therapy group). Preceding NUC therapy was entecavir in 85%<br />

and lamivudine plus adefovir in 15% of patients. The average<br />

duration of preceding NUC therapy was 4.1 years. At the<br />

start of peg-interferon, HBV DNA was undetectable in 93%,<br />

and HBeAg was negative in 69% of patients. For comparison,<br />

68 patients treated by 48 weeks of peg-interferon alpha-2a<br />

(180 microgram weekly) without preceding NUC therapy<br />

was studied (monotherapy group). Serial changes in the titer<br />

of HBsAg were measured by quantitative assay. Results: In<br />

sequential therapy group, HBV DNA remained undetected in<br />

80% of patients after start of peg-interferon and withdrawal of<br />

NUC. In HBeAg positive patients, HBeAg became negative at<br />

the end of therapy in 50% in sequential therapy and 14% in<br />

monotherapy (p=0.04). At the start of peg-interferon, titer of<br />

HBsAg was 850 (1.9-36500) IU/mL in sequential therapy and<br />

1990 (6.0-80360) IU/mL in monotherapy which declined to<br />

166 (1.0 Log IU/mL was 34% vs. 19%, HBsAg reduction<br />

>2.0 Log IU/mL was 16% vs. 7%, and HBsAg loss was 6%<br />

vs. 0% for sequential and monotherapy, respectively. Within<br />

the same individual, average HBsAg reduction during sequential<br />

peg-interferon was significantly higher compared to those<br />

during preceding NUC therapy (0.64 vs. 0.16 Log/year,<br />

p=0.015). Among patients with HBsAg 60 IU/L and decrease in HBsAg level<br />

1.0 Log IU/mL during PEG-IFN alfa-2a therapy. An 8- to 9-year<br />

follow-up of the clinical course showed that the HBsAg levels<br />

decreased at 1.1 Log/mL per year (3.5 Log IU/mL at NA therapy<br />

initiation Ç 2.4 Log IU/mL at the end of the follow-up) in<br />

the 33 patients who had received continuous NA therapy and<br />

at 3.1 Log/mL per year (3.5 Log IU/mL at NA therapy initiation<br />

Ç 2.5 Log IU/mL at PEG-IFN alfa-2a therapy initiation Ç<br />

1.7 Log IU/mL at the end of the follow-up period) in those who<br />

underwent sequential therapy, showing a significant decrease<br />

(p < 0.0001). [Conclusions] Compared with patients who had<br />

undergone continuous NA therapy to decrease HBsAg levels,<br />

those who had received sequential therapy with Peg-IFNα<br />

-2a after a long-term NA had lower HBsAg levels, suggesting<br />

sequential therapy may accelerate the timing of HBsAg loss.<br />

Disclosures:<br />

The following authors have nothing to disclose: Miwa Kawanaka, Ken Nishino,<br />

Jun Nakamura, Tomohiro Tanikawa, Takahito Oka, Noriyo Urata, Mitsuhiko<br />

Suehiro, Hirofumi Kawamoto, Gotaro Yamada


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1199A<br />

2031<br />

Impact of pre-treatment platelet count on hepatocellular<br />

carcinoma development during entecavir treatment of<br />

chronic hepatitis B virus infection<br />

Akira Kawano 1 , Hirohisa Shigematsu 1 , Koichiro Miki 1 , Hironori<br />

Tanimoto 2 , Yasunori Ichiki 10 , Chie Morita 3 , Kimihiko Yanagita 4 ,<br />

Kazuhiro Takahashi 5 , Kazufumi Dohmen 6 , Masataka Seike 7 ,<br />

Hideyuki Nomura 2 , Hiromi Ishibashi 8 , Shinji Shimoda 9 ; 1 Department<br />

of Medicine, Kitakyushu Municipal Medical Center, Kitakyushu,<br />

Japan; 2 The Center for Liver Disease, Shin-Kokura Hospital,<br />

Kitakyushu, Japan; 3 Department of Internal Medicine, JR Kyushu<br />

Hospital, Kitakyushu, Japan; 4 Department of Internal Medicine,<br />

Saiseikai Karatsu Hospital, Saga, Japan; 5 Department of Medicine,<br />

Hamanomachi Hospital, Fukuoka, Japan; 6 Department of<br />

Internal Medicine, Chihaya Hospital, Fukuoka, Japan; 7 Department<br />

of Gastroenterology, Oita University, Oita, Japan; 8 Department<br />

of Hepato-Biliary-Pancreatic Medicine, Fukuoka Sanno Hospital,<br />

Fukuoka, Japan; 9 Department of Medicine and Biosystemic Science,<br />

Kyushu University, Fukuoka, Japan; 10 Department of Internal<br />

Medicine, Japan Community Health care Organization Kyushu<br />

Hospital, Kitakyushu, Japan<br />

Background and Aims: Entecavir (ETV) is one of the first-line<br />

nucleoside analogs for treating patients with chronic HBV infection.<br />

The risk of the hepatocellular carcinoma (HCC) development<br />

for ETV-treated patients remains unclear. The objective of<br />

this study was to evaluate the pre-treatment predictors of the<br />

development of HCC in the patients of chronic HBV infection<br />

treated with ETV, and its influence on clinical course. Methods:<br />

This retrospective multicenter study consisted of 151 Japanese<br />

patients with chronic HBV infection who received ETV treatment<br />

more than 12 months. The median age was 51.0 yo and 97<br />

patients (64.2%) were male. Genotype C was the most prevalent<br />

(86.0%). Median HBV DNA was 7.1 log copies (LC)/mL.<br />

Seventy two patients (47.7%) were positive for hepatitis B e<br />

antigen (HBeAg). Median aminotransferase (ALT) level, serum<br />

albumin level, platelet (PLT) count and serum alpha-fetoprotein<br />

(AFP) level were 111 IU/L, 41 g/L, 157 x10 9 /L and 6.6 ng/<br />

mL, respectively. Median treatment duration was 47.4 months.<br />

Results: Fourteen patients developed HCC during follow-up.<br />

The cumulative incidence of HCC at 3, 5, and 7-year was 7.7,<br />

11.4, and 13.3 %, respectively. Multivariate analysis extracted<br />

lower PLT counts (p=0.0055) and higher age (p=0.0133) as<br />

significant independent pre-treatment predictors of the development<br />

of HCC. Cumulative incidence of HCC was significantly<br />

higher in the low PLT group (


1200A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

antiHBs at the time of NUCs discontinuation. No HBsAg seroreversion<br />

occurred during a median of 12 months follow-up<br />

(range, 2-55 months). No patient developed hepatic complications<br />

or hepatocellular carcinoma. A new 24 months-follow-up<br />

analysis is currently on-going to evaluate previous outcomes.<br />

Those results will be presented during AASLD congress. Conclusions:<br />

These data suggest the time between the diagnosis<br />

of CHB and HBsAg loss is longer in HBeAg negative than<br />

in HBeAg positive patients. In addition, HBsAg loss occurred<br />

faster with TDF/ETV than with LAM/ADV. HBsAg loss was<br />

maintained in all patients after discontinuation of NUCs.<br />

Disclosures:<br />

Emilio Suarez - Advisory Committees or Review Panels: Gilead; Speaking and<br />

Teaching: Gilead, Bristol Myers Squibb<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

Martin Prieto - Advisory Committees or Review Panels: Bristol, Gilead<br />

Rafael Gómez Rodríguez - Consulting: Gilead<br />

Moisés Diago - Grant/Research Support: MSD, JANSSEN; Speaking and Teaching:<br />

BMS, ABBVIE, GILEAD<br />

Rosa María Morillas - Advisory Committees or Review Panels: BRISTOL, GILEAD,<br />

Abvvie; Speaking and Teaching: ROCHE, JANSSEN, MSD<br />

The following authors have nothing to disclose: Miguel A. Simon, Juan Manuel<br />

Pascasio, Manuel Rodriguez, Teresa Casanovas, Javier Crespo, Juan I. Arenas,<br />

Blanca Figueruela, Jose M. Zozaya, José Luís Calleja, Marta Casado, Esther<br />

Molina, Javier Fuentes<br />

2033<br />

Prolongation of interferon therapy increases maintained<br />

virological response rates and improves the natural<br />

course of chronic delta hepatitis<br />

Onur Keskin, Ali Tuzun, Fatih Karakaya, Cagdas Kalkan, Aysun<br />

Caliskan Kartal, Ersin Karatayli, Senem C. Karatayli, A.Mithat<br />

Bozdayi, Ramazan Idilman, Cihan Yurdaydin; Gastroenterology,<br />

Ankara University School of Medicine, Ankara, Turkey<br />

Introduction and Aim: Chronic delta hepatitis (CDH) is the<br />

most severe form of chronic viral hepatitis. Interferons are the<br />

only approved therapy in CDH. HDV RNA 6 months post-treatment<br />

is an unreliable surrogate of treatment efficacy. Hence<br />

maintained virologic response (MVR) is important. Aim of<br />

this study was to assess effect of prolonged interferon therapy<br />

on MVR and on natural course of CDH. Patients and<br />

Methods: We searched our CDH database for patients who<br />

received interferon treatment. 99 CDH patients (70M/29F;<br />

mean age:49.8±11.3 years, 21 cirrhotic/78 non-cirrhotic at<br />

baseline) who received at least 6 months of conventional or<br />

pegylated interferon were identified and included in the analysis.<br />

MVR was defined as durable negative HDV RNA after<br />

2 years of post-treatment follow-up. HDV RNA was assessed<br />

by in-house PCR. Median treatment duration was 24 months<br />

(6-126) and median number of treatment episodes was 2 (1-8).<br />

Duration of IFN therapies were classified as 6-12 (n:32),13-<br />

24 (n:26), 25-36 (n:11), 37-48 (n:15), 49-60 (n:7) and >60<br />

months (n:8), respectively. Median follow up time after treatment<br />

discontinuation was 118 (24-255) months. Overall, 35<br />

patients had a MVR and 64 did not have MVR . Development<br />

of hepatocellular cancer (HCC), hepatic decompansation, liver<br />

related mortality, liver transplantation and HBsAg loss were<br />

considered as clinical end-points. Results: Only 16 patients<br />

(16%) had MVR with 12 months of IFN therapy. With prolongation<br />

of IFN treatment, 19 new patients (6, 2, 4, 3 and 4<br />

patients after 24, 36, 48, 60 and > 60 months of IFN treatment,<br />

respectively) achieved MVR. Cumulative probability of<br />

MVR was 16%, 24%, 26%, 30%, 33% and 38% at end of<br />

6-12, 24, 36, 48, and > 60 months of interferon treatment,<br />

respectively. Development of HCC (p: 0.05), decompansation<br />

(p:0.013), liver-related mortality (p:0.03) and transplantation<br />

(p:0.012) were more common in non-responders and HBsAg<br />

clearance occurred more often in patients with MVR (p:0.002).<br />

By multivariate analysis, low platelet count and no response<br />

to therapy were independent factors for development of any<br />

clinical endpoint (p=0.014 and 0.021, respectively). Conclusion:<br />

Longer duration of IFN therapy increases MVR rates (from<br />

16.1% to 38%) in CDH. Low platelet count at baseline and<br />

no response to treatment were independent predictors for the<br />

development of a clinical endpoint. These results suggest that<br />

treatment with pegylated interferon should be started early in<br />

CDH and that prolonged periods of treatment may be needed.<br />

Further, the results underline the urgent need for new drug<br />

development in CDH.<br />

Disclosures:<br />

Cihan Yurdaydin - Advisory Committees or Review Panels: Janssen, Roche,<br />

Merck, Gilead, AbbVie; Speaking and Teaching: BMS<br />

The following authors have nothing to disclose: Onur Keskin, Ali Tuzun, Fatih<br />

Karakaya, Cagdas Kalkan, Aysun Caliskan Kartal, Ersin Karatayli, Senem C.<br />

Karatayli, A.Mithat Bozdayi, Ramazan Idilman<br />

2034<br />

TKM-HBV, a Novel RNA Interference Treatment for<br />

Chronic Hepatitis B, has a Complementary Mode of<br />

Action to Current Standard of Care Nucleos(t)ide Analogs<br />

Emily P. Thi, Ammen P. Dhillon, Amy C. Lee; Tekmira Pharmaceuticals,<br />

Burnaby, BC, Canada<br />

Current approved nucleos(t)ide analog (NA) drugs for hepatitis<br />

B virus (HBV) are highly effective at inhibiting viral replication<br />

but do not cure the chronic infection and are largely ineffective<br />

in preventing viral protein production. HBV proteins are understood<br />

to play a variety of roles in contributing to suppression<br />

of host immune responses and viral persistence. Quantitative<br />

changes in peripheral HBV surface antigen (HBsAg) in particular<br />

have been correlated with differences in clinical outcomes.<br />

TKM-HBV is a novel RNA interference HBV therapeutic<br />

intervention currently in Phase 1 clinical development. It is a<br />

mixture of three different oligonucleotide duplexes, encapsulated<br />

within a lipid nanoparticle delivery system, that cleaves<br />

all forms of HBV RNAs thus preventing the synthesis of viral<br />

proteins. Because TKM-HBV is understood to act on a separate<br />

viral target and in a different cellular compartment than that<br />

of NAs (in the cytoplasm versus the viral capsid), there is little<br />

basis to expect drug:drug interference in a co-treatment setting.<br />

TKM-HBV combination treatment with NA drugs (including the<br />

current frontline compounds, entecavir and tenofovir) was studied<br />

in two HBV infection models: primary human hepatocytes<br />

(PHH) and humanized mice. To mimic a potential clinical study<br />

setting, TKM-HBV was administered after onset of NA treatment.<br />

Reductions in HBV DNA and HBsAg were observed in<br />

PHH cultures co-treated with TKM-HBV and either entecavir,<br />

tenofovir or lamivudine. No antagonism of either drug activity<br />

was detected and most often additive effects were observed.<br />

A second PHH study examining performance against different<br />

viral genotypes also did not detect any drug:drug interference<br />

in the ability of entecavir to inhibit viral replication or TKM-HBV<br />

to reduce viral antigen load when applied against gt A, B, C<br />

or D. Interestingly, the magnitude of TKM-HBV activity was<br />

more consistent than that of entecavir across the four genotypes<br />

examined. The combination approach was also demonstrated<br />

to be effective in the uPA/SCID humanized mouse model of<br />

chronic HBV, resulting in simultaneous control of both HBV<br />

DNA titre and peripheral HBsAg (3 log and 1 log reductions,<br />

respectively). In summary, these results show that TKM-HBV<br />

and NA modes of action are complementary, and combination


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1201A<br />

therapy allows effective disease targeting at multiple critical<br />

nodes of the viral life cycle.<br />

Disclosures:<br />

Emily P. Thi - Employment: Tekmira Pharmaceuticals<br />

Amy C. Lee - Employment: Tekmira<br />

The following authors have nothing to disclose: Ammen P. Dhillon<br />

2035<br />

The New Cyclophilin Inhibitor STG-175 Efficiently Inhibits<br />

Mono- as well as Co-Infections of HIV-1, HCV and<br />

HBV<br />

Philippe Gallay 1 , Michael Bobardt 1 , Udayan Chatterji 1 , Zhengyu<br />

Long 2 , Shengli Zhang 2 , Zhuang Su 2 ; 1 The Scripps Research Institute,<br />

La Jolla, CA; 2 S & T Global, Inc., Woburn, MA<br />

BACKGROUND: Human immunodeficiency virus type-1 (HIV-<br />

1), hepatitis B virus (HBV) and hepatitis C virus (HCV) share<br />

a common route of transmission. HBV and HCV mono-infection<br />

represent the major causes of chronic liver disease globally.<br />

HIV-1 co-infection with HBV or HCV is associated with<br />

accelerated progression to severe liver disease, increased risk<br />

of hepatotoxicity from antiretroviral therapy and reduced survival.<br />

Co-infected patients are often refractory to most therapies<br />

and develop liver fibrosis, cirrhosis and liver cancer more<br />

often than mono-infected patients. Since a growing body of<br />

evidence suggests that HIV-1, HCV and HBV, all exploit the<br />

host protein cyclophilin A (CypA) to optimally infect and replicate<br />

in human cells, we tested a new cyclophilin inhibitor<br />

STG-175 for its capacity to inhibit mono- as well as co-infections<br />

of these three prime viral human threats. MATERIAL AND<br />

METHODS: For mono-infections: i) human PBMCs were infected<br />

with HIV-1 (JR-CSF) and viral replication was quantified by<br />

HIV-1 capsid/p24 ELISA; ii) hepatoma Huh7.5.1 cells were<br />

infected with HCV (JFH-1) and viral replication was quantified<br />

by HCV core ELISA; and iii) NTCP-positive Huh7 cells were<br />

infected with HBV AD38 and viral replication was quantified<br />

by HBV HBeAg ELISA. For dual and triple infections: target<br />

cell populations were mixed prior to virus exposure. For drug<br />

treatments, STG-175 was added to cells either i) together with<br />

viruses or ii) 3 days post-infection and then every 3 days for<br />

a period of 12 days. RESULTS: In this in vitro co-culture infection<br />

system, we found that STG-175 inhibits in a dose-dependent<br />

manner mono-infections, dual co-infections as well as the<br />

triple HIV-1/HCV/HBV co-infection. The degree of STG-175<br />

antiviral efficacy was HCV > HIV-1 > HBV. The addition of<br />

STG-175 together with virus totally blocked HCV and HIV-1<br />

infection and greatly attenuated HBV infection. When added<br />

3 days post-infection and then every 3 days, STG-175 totally<br />

eradicated the pre-established HCV infection, almost totally<br />

aborted the established HIV-1 infection and prevented the viral<br />

expansion of the pre-established HBV infection. Similar results<br />

were observed during the triple HIV-1/HCV/HBV co-infection.<br />

STG-175 was found to be constantly more efficacious than<br />

the well-characterized cyclophilin inhibitor alisporivir in both<br />

mono- and co-infections. CONCLUSIONS: By demonstrating a<br />

potent and broad spectrum of antiviral activity, the new cyclophilin<br />

inhibitor STG-175 represents an attractive drug partner<br />

for an IFN-free regimen for the treatment of HIV-1/HCV/HBV<br />

co-infections.<br />

Disclosures:<br />

The following authors have nothing to disclose: Philippe Gallay, Michael Bobardt,<br />

Udayan Chatterji, Zhengyu Long, Shengli Zhang, Zhuang Su<br />

2036<br />

Association of Interferon-gamma Inducible Protein 10<br />

Polymorphism with Treatment Response to Pegylated<br />

Interferon in HBeAg-Positive Chronic Hepatitis B<br />

Pisit Tangkijvanich 1 , Umaporn Limothai 1 , Natthaya Chuaypen 1 ,<br />

Apichaya Khlaiphuengsin 1 , Rujipat Wasitthankasem 2 , Yong Poovorawan<br />

3 ; 1 Biochemistry, Chulalongkorn University, Bangkok, Thailand;<br />

2 Chulalongkorn University, Bangkok, Thailand; 3 Pediatrics,<br />

Chulalongkorn University, Bangkok, Thailand<br />

Background: Interferon-gamma inducible protein 10 (IP-10)<br />

plays an important role in the natural history and treatment<br />

outcome of patients with chronic hepatitis B (CHB). The aim of<br />

this study was to investigate the association of single nucleotide<br />

polymorphism (SNP) G-201A in the promoter region of the IP-10<br />

gene and treatment response to pegylated interferon (PEG-IFN)<br />

in patients with HBeAg-positive CHB. The potential effect of<br />

SNP rs12979860 in the interferon lambda-3 (IFNL3) gene was<br />

also examined. Method: We retrospectively analyzed data of<br />

Thai patients with HBeAg-positive CHB treated with PEG-IFN<br />

for 48 weeks. Virological response (VR) was defined as HBeAg<br />

clearance and HBV DNA < 2,000 IU/mL at 24 weeks post<br />

treatment. The SNPs G-201A and rs12979860 were identified<br />

by PCR–RFLP method. HBV genotype was performed by direct<br />

sequencing. Baseline serum IP-10 levels were measured by a<br />

commercially available assay. Results: Among 107 patients<br />

enrolled (72 male, mean age 34.3 years), VR was achieved<br />

in 45 (42.1%) patients. HBsAg clearance and HBsAg decline<br />

(< 100 IU/mL) at 24 weeks post treatment were achieved in<br />

10 (9.3%) and 22 (20.6%) patients, respectively. The distribution<br />

of HBV genotypes B and C was 12.1% and 81.3%,<br />

respectively. The overall distribution of GG, GA and AA genotypes<br />

of G-201A was 76.6%, 19.6% and 3.7%, respectively.<br />

Patients with GG genotype, compared to those with non-GG<br />

genotype, achieved significantly higher rated of VR (48.8% vs.<br />

19.2%; P=0.011), HBsAg decline (25.6% vs. 4.0%, P=0.019),<br />

and a trend of HBsAg clearance (11.0% vs. 4%, P=0.294).<br />

Patients with GG-genotype had more pronounced decline of<br />

serum HBsAg during and after therapy (weeks 0, 4, 12, 24,<br />

48 and 72) and had higher baseline IP-10 levels than those<br />

with non-GG genotype (432.2±339.0 vs. 257.3±145.7 pg/<br />

mL, P=0.028). Patients achieving VR had significantly higher<br />

pretreatment IP-10 levels than non-responders (496.0±394.6<br />

vs. 328.0±232.2, P=0.013). The distribution of CC, CT and<br />

TT genotypes of rs12979860 was 89.7%, 10.3% and 0%,<br />

respectively. Patients with CC and CT genotypes achieved<br />

VR in 41.7% and 45.5%, respectively (P=0.810). In logistic<br />

regression analysis, SNP G-201A was an independent factor<br />

associated with VR (odds ratio 3.81, 95% confidence interval:<br />

1.31-11.12, P=0.014). Conclusions: These data demonstrated<br />

that IP-10 polymorphism, but not IFNL3 polymorphism, was<br />

independently associated with response to PEG-IFN therapy in<br />

Thai patients with HBeAg-positive CHB.<br />

Disclosures:<br />

The following authors have nothing to disclose: Pisit Tangkijvanich, Umaporn Limothai,<br />

Natthaya Chuaypen, Apichaya Khlaiphuengsin, Rujipat Wasitthankasem,<br />

Yong Poovorawan


1202A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2037<br />

Whole-Genome Sequencing of Patients Achieving Hepatitis<br />

B Virus Serum Antigen Loss Following TDF Treatment<br />

Sarah E. Kleinstein 1,2 , Harry L. Janssen 3,4 , Patrick Marcellin 5,6 ,<br />

Dongliang Ge 7 , Anuj Gaggar 7 , Mani Subramanian 7 , John G.<br />

McHutchison 7 , Maria Buti 8 , Alex J. Thompson 9 , David B. Goldstein<br />

1 ; 1 Center for Human Genome Variation, Duke University,<br />

New York, NY; 2 Molecular Genetics & Microbiology, Duke University,<br />

Durham, NC; 3 Toronto Centre for Liver Disease, University<br />

Health Network, Toronto, ON, Canada; 4 Erasmus Medical Center,<br />

Rotterdam, Netherlands; 5 Viral Hepatitis Research Centre, Clichy,<br />

France; 6 Service d’Hépatologie, Hôpital Beaujon, Clichy, France;<br />

7 Gilead Sciences Inc., Foster City, CA; 8 Hospital Valle Hebron and<br />

Ciberehd del Institut Carlos III, Barcelona, Spain; 9 St Vincent’s Hospital<br />

and the University of Melbourne, Melbourne, VIC, Australia<br />

Purpose: While treatment with tenofovir disoproxil fumerate<br />

(TDF) provides durable suppression of hepatitis B virus (HBV),<br />

serum antigen (HBsAg) loss, which is considered a “cure”, is<br />

infrequent; thus, we investigated the role of rare human genetic<br />

variants in predicting HBsAg loss following treatment with TDF<br />

among chronic hepatitis B (CHB) patients. Methods: 10 Caucasian<br />

patients with CHB (HBeAg+) achieved HBsAg loss following<br />

treatment in a Phase 3 study, consented, and had available<br />

DNA. Whole-genome sequencing (WGS) was performed on<br />

patient DNA using the Illumina HiSeq2000 (v3 chemistry).<br />

The genetic data from the 10 HBsAg loss cases were compared<br />

to 122 healthy Caucasian WGS controls, matched on<br />

sequencing platform and chemistry. All samples underwent<br />

standard quality control checks. Case-control gene-based and<br />

single-variant (Fisher’s exact test) associations were run across<br />

all genetic models. Gene-based analyses were limited to coding<br />

region, functional variants. A Bonferroni-corrected p-value<br />

for the number of genes or variants tested was set as the statistical<br />

significance threshold. Results: WGS analysis of the 10<br />

HBsAg loss patients did not identify any rare variants that conferred<br />

a significant beneficial response to TDF when analyzed<br />

at either the gene or single-variant level. However, non-significant<br />

suggestive associations were identified in 9 genes.<br />

Rare coding variants enriched among patients with HBsAg loss<br />

were identified in: PDCD1, IL2RA, SCTR, IFI16, ULK4, GPA33,<br />

CCHCR1, GMEB1, and GPNMB. PDCD1 has been associated<br />

with impaired immune response in CHB. SCTR, GPA33 and<br />

CCHCR1 have been linked to the natural history of CHB. Variants<br />

were identified under a dominant (gene-based) or allelic<br />

(single-variant) model, except for IFI16 (recessive model) and<br />

ULK4 (compound-heterozygous model). Conclusions: WGS of<br />

a small cohort of Caucasian CHB patients with HBsAg loss<br />

following TDF treatment identified enrichment of several interesting<br />

rare variants. Variants were identified in genes involved<br />

in the immune system and/or virus replication. It is currently<br />

unclear whether these identified genes and variants play a role<br />

in HBsAg loss or control of HBV replication. Validation <strong>studies</strong><br />

are planned in an independent cohort.<br />

Rare coding variants enriched among patients with HBsAg loss<br />

a Gene-based;<br />

b Different control variant<br />

Disclosures:<br />

Harry L. Janssen - Consulting: AbbVie, Bristol Myers Squibb, GSK, Gilead Sciences,<br />

Innogenetics, Merck, Medtronic, Novartis, Roche, Janssen, Medimmune,<br />

ISIS Pharmaceuticals, Tekmira; Grant/Research Support: AbbVie, Bristol Myers<br />

Squibb, Gilead Sciences, Innogenetics, Merck, Medtronic, Novartis, Roche,<br />

Janssen, Medimmune<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

Dongliang Ge - Employment: Gilead Sciences, Inc<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Maria Buti - Advisory Committees or Review Panels: Gilead, Janssen, MSD;<br />

Grant/Research Support: Gilead, Janssen; Speaking and Teaching: Gilead,<br />

Janssen, BMS<br />

Alex J. Thompson - Advisory Committees or Review Panels: Gilead, Abbvie,<br />

BMS, Merck, Spring Bank Pharmaceuticals, Arrowhead, Roche; Grant/Research<br />

Support: Gilead, Abbvie, BMS, Merck; Speaking and Teaching: Roche, Gilead,<br />

Abbvie, BMS<br />

David B. Goldstein - Advisory Committees or Review Panels: Sever Adverse<br />

Events Consortium; Grant/Research Support: Gilead, Astra Zeneca, Biogen,<br />

Johnson and Johnson, Labcorp, Merck, Fundazione Telethon, NIH, CURE, UCB,<br />

IL28B and ITPA finds<br />

The following authors have nothing to disclose: Sarah E. Kleinstein


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1203A<br />

2038<br />

A randomized trial evaluating the antiviral efficacy of<br />

switching from Lamivudine plus Adefovir to Tenofovir<br />

disoproxil fumarate monotherapy in Lamivudine-resistant<br />

chronic hepatitis B patients with undetectable hepatitis<br />

B virus DNA - The 48 week interim analysis<br />

Heon Ju Lee 2 , Jeong Min Kim 1 , Young Oh Kweon 3 , Soo Young<br />

Park 3 , Jeong Heo 4 , Hyun Young Woo 4 , Jaeseok Hwang 1 , Woo<br />

Jin Chung 1 , Chang Hyeong Lee 5 , Byung Seok Kim 5 , Jeong Ill Suh 6 ,<br />

Won Young Tak 3 , Byoung Kuk Jang 1 ; 1 Internal medicine, Keimyung<br />

University School of Medicine, Daegu, Korea (the Republic of);<br />

2 Internal medicine, Yeungnam University College of Medicine,<br />

Daegu, Korea (the Republic of); 3 Internal medicine, Kyungpook<br />

National University College of Medicine, Daegu, Korea (the<br />

Republic of); 4 Internal medicine, Pusan National University School<br />

of Medicine, Busan, Korea (the Republic of); 5 Internal medicine,<br />

Catholic University of Daegu College of Medicine, Daegu, Korea<br />

(the Republic of); 6 Internal medicine, Dongguk University College<br />

of Medicine, Gyeongju, Korea (the Republic of)<br />

Introduction: Tenofovir disoproxil fumarate (TDF) monotherapy<br />

is an effective treatment for patients who have lamivudine<br />

(LAM)-resistant chronic hepatitis B (CHB). However, the efficacy<br />

of switching to TDF monotherapy for LAM resistant CHB patient<br />

with undetectable HBV DNA while on LAM plus adefovir (ADV)<br />

combination therapy (stable switching) is not well established.<br />

Material and Methods: In this non-inferiority trial, LAM-resistant<br />

CHB patients who had undetectable serum HBV DNA (40 IU/mL at two consecutive timepoints,<br />

or persistent HBV DNA levels of 20-40 IU/mL at three<br />

consecutive timepoints. Results: A total of 169 CHB patients<br />

were enrolled in this study including 74(43.8%) with compensated<br />

cirrhosis. There were no significant differences between<br />

two groups in age, gender, biochemistry findings and duration<br />

of LAM and ADV combination therapy at baseline. Twelve<br />

(20.7%) patients in the LAM/ADV group and 18 (16.2%)<br />

patients in TDF group were HBeAg-positive. Nine patients (4 in<br />

the LAM/ADV group and 5 in the TDF group) discontinued the<br />

study. After a mean follow up period of 48 weeks, there were<br />

no subjects in either group who experienced viral reactivation.<br />

The number of patients with HBeAg loss at week 48 was<br />

2/12(16.7%) and 3/18(16.7%) in the LAM/ADV and TDF<br />

groups, respectively. Two patients achieved HBeAg seroconversion<br />

(1/12 [8.3%] in LAM/ADV group and 1/18 [5.6%]<br />

in TDF group). Transient virological rebound occurred in 10<br />

patients (5 patients in LAM/ADV group and 5 patients in TDF<br />

group) through 48 weeks; however, all patients subsequently<br />

achieved HBV DNA undetectablility at the next study visit. No<br />

patient experienced significant increases in serum creatinine<br />

levels above baseline. The mean changes in serum creatinine<br />

levels of both groups were minimal throughout the 48 weeks of<br />

treatment. Conclusions: Stable switching to TDF monotherapy<br />

for 48 weeks demonstrated a comparable virological response<br />

to maintaining LAM plus ADV combination therapy in LAM-resistant<br />

CHB patients with undetectable HBV DNA. (Trial registration<br />

number Clinicaltrial.gov ID NCT01732367)<br />

Disclosures:<br />

Jeong Heo - Grant/Research Support: Roche<br />

Won Young Tak - Advisory Committees or Review Panels: Gilead Korea; Grant/<br />

Research Support: SAMIL Pharma; Speaking and Teaching: Bayer Korea<br />

The following authors have nothing to disclose: Heon Ju Lee, Jeong Min Kim,<br />

Young Oh Kweon, Soo Young Park, Hyun Young Woo, Jaeseok Hwang, Woo<br />

Jin Chung, Chang Hyeong Lee, Byung Seok Kim, Jeong Ill Suh, Byoung Kuk Jang<br />

2039<br />

WITHDRAWN<br />

2040<br />

Association between Pre-S2 Deletions and Virologic<br />

Rebound during Peginterferon Therapy in Patients with<br />

Hepatitis B E Antigen-Positive Chronic Hepatitis B<br />

Cheng-Yuan Peng 1 , Chao-Jen Shih 2 , Pei-Shuan Lo 1 , Chia-Lin<br />

Huang 1 , Chia-Hsin Lin 1 , Hsueh-Chou Lai 1 , Wen-Pang Su 1 ; 1 Division<br />

of Hepatogastroenterology, Department of Internal Medicine,<br />

China Medical University Hospital, Taichung, Taiwan; 2 Agricultural<br />

Biotechnology Research Center, Academia Sinica, Taipei,<br />

Taiwan<br />

Background: To determine the incidence of on-treatment virologic<br />

rebound (VR) during peginterferon (Peg-IFN) therapy in<br />

patients with HBeAg-positive chronic hepatitis B (CHB) and its<br />

association with genomic mutations in hepatitis B virus (HBV).<br />

Methods: We analyzed the on-treatment HBV DNA kinetics<br />

in 103 consecutive HBeAg-positive CHB patients treated with<br />

Peg-IFN for 24 (n=58) or 48 weeks (n=45). On-treatment<br />

serum HBV DNA levels were measured every 3 months (Cobas<br />

Amplicor HBV Monitor Test, Roche Diagnostics, sensitivity:<br />

60 IU/mL). VR was defined as an on-treatment increase in<br />

HBV DNA level of >1 log IU/mL over prior nadir level with or<br />

without an initial decline. Full-length HBV genomic population<br />

sequencing was done at baseline, nadir, VR, and 24 weeks<br />

off therapy. Precore and basal core promoter mutations and<br />

the pre-S2 deletions were examined at baseline. Proportions<br />

of the pre-S2 deletion mutants among total viral populations<br />

were measured using quantitative PCR at baseline, nadir, VR,<br />

and 24 weeks off therapy. The replication capability of HBV<br />

mutants was determined after transient transfection with the<br />

pTriEx-HBV vector in Huh7 cell line. Results: Twelve (12%)<br />

patients displayed on-treatment VR. Univariate analysis identified<br />

HBV DNA


1204A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Cheng-Yuan Peng - Advisory Committees or Review Panels: AbbVie, BMS, Gilead,<br />

MSD, Roche<br />

The following authors have nothing to disclose: Chao-Jen Shih, Pei-Shuan Lo,<br />

Chia-Lin Huang, Chia-Hsin Lin, Hsueh-Chou Lai, Wen-Pang Su<br />

2041<br />

WITHDRAWN<br />

2042<br />

Is efficacy of tenofovir in nucleos(t)ide analogue-experienced<br />

chronic hepatitis B patients comparable to that in<br />

nucleos(t)ide analogue-naïve patients?<br />

Eun Ju Cho 1 , Jeong-Hoon Lee 1 , Jeong-Ju Yoo 1 , Yuri Cho 1 , Donghyeon<br />

Lee 1 , Yun Bin Lee 2 , Su Jong Yu 1 , Yoon Jun Kim 1 , Jung-Hwan<br />

Yoon 1 ; 1 Internal Medicine, Seoul National University Hospital,<br />

Seoul, Korea (the Republic of); 2 Internal Medicine, CHA Bundang<br />

Medical Center, Seoungnam, Korea (the Republic of)<br />

Background/Aim: Recent study reported that entecavir fails to<br />

show comparable efficacy in nucleos(t)ide analogue (NA)-experienced<br />

chronic hepatitis B (CHB) patients as compared<br />

to NA-naïve ones. We evaluated the efficacy of tenofovir in<br />

NA-experienced CHB patients in comparison with NA-naïve<br />

patients. Methods: We retrospectively included 221 consecutive<br />

patients who had serum hepatitis B virus (HBV) DNA<br />

level > 2,000 IU/mL at the initiation of tenofovir treatment<br />

and received tenofovir for > 3 months. Complete virologic<br />

response (CVR) was defined as undetectable serum HBV DNA<br />

by real-time PCR. CVR rates were calculated by Kaplan-Meier<br />

analysis, and a multivariate Cox proportional-hazard model<br />

was generated in order to find predictive factors independently<br />

associated with the time to a CVR. Results: Of the 221 patients<br />

(129 males; mean age, 48.2 years), 149 were NA-naïve and<br />

72 were NA-experienced patients. Ninety-eight patients had<br />

hepatitis B e antigen (HBeAg)-positive CHB. The median duration<br />

of tenofovir treatment was 47.4 (interquartile range, 36.2<br />

to 56.6) weeks. The CVR rates of the NA-naïve and NA-experienced<br />

groups were comparable in both HBeAg-negative<br />

(49.4% vs 58.1 at week 24, P=0.53; 96.4% vs 93.3% at<br />

week 48, P=0.60) and HBeAg-positive patients (22.2% vs<br />

16.7% at week 24, P=0.60; 75.6% vs 75.9% at week 48,<br />

P=1.00). According to the multivariate Cox regression model,<br />

only HBeAg positivity (hazard ratio [HR], 0.40; 95% confidence<br />

interval [CI], 0.29 to 0.56; P< 0.001) and baseline<br />

HBV DNA levels (HR, 0.77 for 1 log10 IU/mL increase; 95%<br />

CI, 0.69 to 0.86; P< 0.001) had a significant influence on the<br />

time to a CVR. The presence of resistance mutations to lamivudine/telbivudine/entecavir<br />

did not influence the response<br />

rates, however, patients with adefovir resistance showed significantly<br />

lower CVR rate (25%; P=0.03). Conclusion: Tenofovir<br />

has comparable efficacy in NA-naïve and NA-experienced<br />

patients without adefovir resistance.<br />

Disclosures:<br />

Yoon Jun Kim - Grant/Research Support: Bristol-Myers Squibb, Roche, JW Creagene,<br />

Bukwang Pharmaceuticals, Handok Pharmaceuticals, Hanmi Pharmaceuticals,<br />

Yuhan Pharmaceuticals; Speaking and Teaching: Bayer HealthCare<br />

Pharmaceuticals, Gilead Science, MSD Korea, Yuhan Pharmaceuticals, Samil<br />

Pharmaceuticals, CJ Pharmaceuticals, Bukwang Pharmaceuticals, Handok Pharmaceuticals<br />

The following authors have nothing to disclose: Eun Ju Cho, Jeong-Hoon Lee,<br />

Jeong-Ju Yoo, Yuri Cho, Donghyeon Lee, Yun Bin Lee, Su Jong Yu, Jung-Hwan<br />

Yoon<br />

2043<br />

Serologic and Virologic Changes Following ALT flares<br />

in adolescents With Chronic Hepatitis B during therapy<br />

with Tenofovir Disoproxil Fumarate or Placebo.<br />

Karen F. Murray 1 , Leszek Szenborn 2 , Jacek Wysocki 3 , Benedetta<br />

Massetto 4 , Anuj Gaggar 4 , Kyungpil Kim 4 , Mani Subramanian 4 ,<br />

Luminita Luminos 5 , Malgorzata Pawlowska 6 , Jacek Mizerski 7 ;<br />

1 Children’s Hospital, Seattle, WA; 2 Wroclaw Medical University,<br />

Wroclaw, Poland; 3 Poznan University of Medical Sciences,<br />

Poznan, Poland; 4 Gilead Sciences, Inc., Foster City, CA; 5 Prof. Dr.<br />

Matei Bals Institute for Infectious Diseases, Bucharest, Romania;<br />

6 Wojewodzki Szpital Obserwacyjno-Zakazny im. T. Browicza,<br />

Bydgoszcz, Poland; 7 John Paul II Hospital, Kraków, Poland<br />

Background and Aims: We previously described the safety and<br />

antiviral efficacy of TDF in adolescents aged 12 to 2x baseline and<br />

>5x upper limit of normal during the first 48 weeks. HBeAg<br />

loss, HBV DNA decline and HBsAg decline were evaluated 24<br />

weeks following a flare. Results: Twelve of the 52 (23%) evaluable<br />

PBO patients and 1 of the 52 (2%) TDF-treated patients<br />

had ALT flares during the first 48 weeks. PBO patients with<br />

ALT flare had an increased rate of 1-log 10<br />

HBV DNA decline<br />

and HBeAg loss compared to those without ALT flare (Table,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1205A<br />

Anuj Gaggar - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

The following authors have nothing to disclose: Leszek Szenborn, Kyungpil Kim,<br />

Luminita Luminos, Malgorzata Pawlowska, Jacek Mizerski<br />

2044<br />

The baseline combination of HBcrAg and HBsAg titers<br />

enhance treatment outcome and HBsAg loss predictions<br />

in HBeAg negative chronic hepatitis B patients treated<br />

with pegylated interferon alfa-2a or PegIFN plus Tenofovir-disoproxil-fumarate.<br />

Michelle Martinot-Peignoux 1,2 , Martine Lapalus 1 , Nathalie Boyer 2 ,<br />

Corinne Castelnau 2 , Nathalie Giuily 2 , Michele Pouteau 2 , Rami<br />

Moucari 1 , Tarik Asselah 2,1 , Patrick Marcellin 2,1 ; 1 Université Denis<br />

Diderot Paris 7, INSERM, UMR1149, Centre de Recherche sur<br />

l’Inflammation, Clichy, France; 2 Hôpital Beaujon AP-HP, Service<br />

d’Hépatologie, Clichy, France<br />

Background. It remains unclear whether adding a nucleot(s)ide<br />

analogue (NAs) enhances the efficacy of pegylated interferon<br />

alfa-2a (PegIFN) by accelerating HBsAg clearance. Baseline<br />

HBsAg level is predictor of SVR in patients receiving PegIFN.<br />

Hepatitis B core-related antigen (HBcrAg) is a new serological<br />

HBV marker that could to be used for NAs treatment cessation.<br />

It combines 3 viral proteins coded by the precore/core region<br />

of the covalently closed circular DNA: HBeAg, HBcAg and<br />

a core related protein (p22cr) by sharing a 149 amino-acid<br />

sequence. We aimed to evaluate if the baseline combination<br />

of HBcrAg and HBsAg levels might have a complimentary role<br />

predicting treatment response, in patients receiving PegIFN)<br />

therapy or PegIFN plus Tenofovir-disoproxil-fumarate (PegIF-<br />

N+TDF) Patients-Methods. 62 patients HBeAg negative CHB<br />

patients were treated with 48 weeks PegIFN or PegIFN+TDF.<br />

Patients were followed 3 years after treatment cessation. HBsAg<br />

and HBcrAg titers measured at baseline. HBsAg and HBV DNA<br />

were measured at baseline, week 24, end of therapy (EOT),<br />

24 weeks, 1, 2 and 3 years after treatment cessation. Results.<br />

30 patients received PegIFN and 32 patients received PegIF-<br />

N+TDF. The 2 groups of patients were similar with regards to<br />

age, sex ratio, ALT levels, HBV genotypes distribution, histologic<br />

lesions, mean HBsAg, HBcrAg and HBV DNA titers. An<br />

EOT response was observed in 90% and 100% of PegIFN and<br />

PegIFN+TDF patients, respectively. SVR was 10/30 (33%) and<br />

17/32 (53%) in PegIFN and PegIFN+TDF patients, respectively.<br />

At the end of 3 years post-treatment follow-up an HBsAg<br />

loss was observed in 7/30 (1 at EOT) and 6/32 (4 at EOT) in<br />

PegIFN and PegIFN+TDF patients, respectively. Baseline negative<br />

predictive values (NPVs) for SVR of HBsAg threshold 3.302<br />

log IU/ml and HBcrAg threshold 3.699 log U/ml and combination<br />

of both thresholds were: 78% or 63%, 74% or 61%<br />

and 80% or 75% in patients receiving PegIFN or PegIFN+TDF,<br />

respectively. Baseline NPVs for HBsAg loss of HBsAg threshold<br />

3.302 log IU/ml and HBcrAg threshold 3.699 log U/ml and<br />

combination of both thresholds were: 88% or 100%, 79% or<br />

89% and 87% or 100% in patients receiving PegIFN or PegIF-<br />

N+TDF, respectively. Conclusions. In HBe negative patients,<br />

our results indicate that the addition of TDF to PegIFN enhance<br />

the rate of viral suppression with 53% SVR and 12.5% EOT<br />

HBs loss, while SVR was observed in 33% and EOT HBs loss<br />

in 3.3% patients receiving with PegIFN monotherapy. Baseline<br />

combination of HBsAg and HBcrAg titers allows the identification,<br />

prior therapy, of patients with low probability of SVR and<br />

HBsAg loss indicating that these markers could be used for “à<br />

la carte therapy”.<br />

Disclosures:<br />

Nathalie Boyer - Board Membership: MSD, JANSSEN, Gilead, Abbvie; Speaking<br />

and Teaching: BMS<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

The following authors have nothing to disclose: Michelle Martinot-Peignoux,<br />

Martine Lapalus, Corinne Castelnau, Nathalie Giuily, Michele Pouteau, Rami<br />

Moucari<br />

2045<br />

Baseline HBsAg titer allows identification of patients<br />

that will benefit from pegylated interferon therapy and<br />

experience an HBsAg loss.<br />

Michelle Martinot-Peignoux 2,3 , Nathalie Boyer 3 , Corinne Castelnau<br />

3 , Nathalie Giuily 3 , Michele Pouteau 3 , Martine Lapalus 1,2 , Rami<br />

Moucari 1,2 , Tarik Asselah 3,1 , Patrick Marcellin 3,2 ; 1 Université Paris<br />

Diderot, INSERM U773/CRB3 et Service d’hépatologie, Clichy,<br />

France; 2 Université Denis Diderot Paris 7, INSERM, UMR1149,<br />

Centre de Recherche sur l’Inflammation, Clichy, France; 3 Hôpital<br />

Beaujon AP-HP, Service d’Hépatologie, Clichy, France<br />

Background. A 48 weeks treatment with pegylated interferon<br />

alfa-2a (PegIFN) is the only treatment allowing to obtain 30% of<br />

sustained virological response (SVR) and 10% HBs loss within<br />

3 years post-treatment follow-up. It is well demonstrated that<br />

early on treatment HBs decline is predictive of response. We<br />

aimed to evaluate the predictive value of baseline HBsAg titer<br />

to predict SVR and HBs loss, in patients treated with PegIFN or<br />

PegIFN plus Tenofovir-disoproxil-fumarate (PegIFN+TDF) Methods.<br />

62 HBeAg(-) CHB patients treated for 48 weeks. HBsAg<br />

and HBVDNA titers were measured at baseline, end of therapy<br />

(EOT), 24 weeks, 1 2 and 3 years after treatment cessation.<br />

HBV genotype distribution at baseline. Results. 30 patients<br />

received PegIFN and 32 patients received PegIFN+TDF. An<br />

EOT response was observed in 90% and 100% of PegIFN and<br />

PegIFN+TDF patients, respectively. An SVR was observed in<br />

27/ 62 (43%), 10/30 (33%) and 17/32 (53%) in the overall,<br />

PegIFN and PegIFN+TDF patients, respectively. At the end of<br />

3 years post-treatment follow-up an HBsAg loss was observed<br />

in 13/62 (21%) 7/30 (23%) and 6/32 (19%), in the overall,<br />

PegIFN and PegIFN+TDF patients, respectively. HBsAg and<br />

HBV DNA titers were significantly higher in SVR than in non-responders<br />

(p2000 IU/ml including<br />

27/35 (77%) non-responders patients and 36/49 (74%)<br />

patients with no-HBs loss within 3 years post-treatment follow-up.<br />

The positive predictive value of HBsAg titer ≤ 2000UI/<br />

ml for SVR and HBs loss were 67% and 46% or 55% and<br />

45% or 77% and 46% in the overall population or patients<br />

receiving PegIFN or PegIFN+TDF, respectively. The negative<br />

predictive value (NPVs) of baseline HBsAg titer >2000 IU/ml<br />

for SVR and HBs loss were 71% and 85% or 79% and 89% or<br />

63% and 100% in the overall population or patients receiving<br />

PegIFN or PegIFN+TDF, respectively. The NPVs for genotypes<br />

A, B, D, and E, were 50%, 80%, 74% and 82%, respectively.<br />

Conclusions. Our results show that at baseline an HBsAg threshold<br />

of 2000 IU/ml is highly predictive of treatment response<br />

and HBsAg loss in patients treated with 48 weeks of PegIFN<br />

patients or PegIFN+TDF. Among the patients with an HBsAg<br />

level ≤ 2000 IU/ml 59% demonstrated a SVR and 85% expe-


1206A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

rienced HBs loss within 3 years after treatment cessation. This<br />

observation demonstrates for the first time that baseline HBsAg<br />

titer might allows identifying patients that could or could not<br />

benefit from PegIFN therapy.<br />

Disclosures:<br />

Nathalie Boyer - Board Membership: MSD, JANSSEN, Gilead, Abbvie; Speaking<br />

and Teaching: BMS<br />

Tarik Asselah - Advisory Committees or Review Panels: AbbVie, Merck, Gilead,<br />

BMS, Roche, Janssen<br />

Patrick Marcellin - Consulting: Roche, Gilead, BMS, Vertex, Novartis, Janssen,<br />

MSD, Abbvie, Alios BioPharma, Idenix, Akron; Grant/Research Support: Roche,<br />

Gilead, BMS, Novartis, Janssen, MSD, Alios BioPharma; Speaking and Teaching:<br />

Roche, Gilead, BMS, Vertex, Novartis, Janssen, MSD, Boehringer, Pfizer,<br />

Abbvie<br />

The following authors have nothing to disclose: Michelle Martinot-Peignoux,<br />

Corinne Castelnau, Nathalie Giuily, Michele Pouteau, Martine Lapalus, Rami<br />

Moucari<br />

2046<br />

Intrahepatic IP-10 expression correlates with plasma<br />

IP-10 levels and is a response marker for HBeAg-positive,<br />

but not HBeAg-negative chronic hepatitis B patients<br />

treated with peginterferon and adefovir<br />

Sophie Willemse 2 , Louis Jansen 2 , Annikki de Niet 2 , Marjan J. Sinnige<br />

1 , Bart Takkenberg 2 , Joanne Verheij 3 ; 1 Experimental Immunology,<br />

Academic Medical Center, Amsterdam, Netherlands;<br />

2 Gastroenterology and Hepatology, Academic Medical Center,<br />

Amsterdam, Netherlands; 3 Pathology, Academic Medical Center,<br />

Amsterdam, Netherlands<br />

Background and aims Interferon-y–inducible protein-10 (IP-10)<br />

is an interferon-stimulated gene that is produced by different<br />

types of cells such as monocytes, neutrophils and hepatocytes.<br />

After binding to its receptor CXCR3, IP-10 functions as a chemotactic<br />

cytokine for T-lymphocytes, monocytes and NK-cells<br />

and induces adhesion of activated memory/effector T-cells.<br />

We aimed to establish if IP-10 expression in liver tissue and<br />

IP-10 levels in plasma of chronic hepatitis B (CHB) patients<br />

correlated with each other and further to investigate if IP-10<br />

levels could predict treatment outcome in CHB patients treated<br />

with peginterferon and adefovir. Methods A total of 86 CHB<br />

patients (41 HBeAg-positive and 45 HBeAg-negative) received<br />

combination therapy of peginterferon and adefovir for 48<br />

weeks. Combined Response (CR) (HBeAg-negativity, HBV-DNA<br />

≤2,000 IU/mL, and ALT normalization) and non-response (NR)<br />

were assessed at Week 72. Plasma IP-10 levels were measured<br />

at baseline and during treatment at Day 3 (D3) and<br />

Week 1 (W1). Pre-treatment liver biopsies were obtained in<br />

69 of 86 patients, of which 41 were stored in liquid nitrogen,<br />

enabling the analysis of intrahepatic IP-10 expression by<br />

RT-qPCR. Results CR was achieved in 14/41 HBeAg-positive<br />

and 17/45 HBeAg-negative patients. Mean baseline plasma<br />

IP-10 levels were significantly higher in HBeAg-positive patients<br />

with CR than NR (3.20 vs 3.00 log pg/mL p=0.023). This<br />

difference was not observed in HBeAg-negative patients. Baseline<br />

IP-10 levels correlated with ALT-levels in HBeAg-positive<br />

(r0.66 p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1207A<br />

Disclosures:<br />

Nowlan Selvapatt - Speaking and Teaching: Gilead, BMS<br />

Ashley S. Brown - Advisory Committees or Review Panels: MSD, Roche, Bristol-Myers-Squibb,<br />

Gilead, Janssen, Abbvie, Achillion; Speaking and Teaching:<br />

MSD, Roche, Bristol-Myers Squibb, Gilead, Janssen, Abbvie<br />

The following authors have nothing to disclose: Bilal Khan, Abbesega Ananthavarathan<br />

2048<br />

A Multicenter, Prospective, Observational, Non-interventional<br />

Cohort Study in Chinese Subjects with HBeAg<br />

Negative Chronic Hepatitis B Receiving Therapy with<br />

Peginterferon alfa: An Interim Analysis<br />

Xinyue Chen 1 , Qianguo Mao 2 , Hui Li 3 , Xu Li 4 , Yuqin Xu 5 , Yao<br />

Xie 6 , Yingxia Liu 7 , Weiping Zhu 8 , Xiaoguang Dou 9 , Fusheng<br />

Zhu 10 , Huanwei Zheng 11 , Shuqin Zhang 12 , Mingxiang Zhang 13 ,<br />

Yongfeng Yang 14 , Yaoren Hu 15 , Chuanwu Zhu 16 , Jiguang Ding 17 ,<br />

Youwen Tan 18 , Qing Xie 21 , Jifang Sheng 19 , Zhiliang Gao 20 , Yi<br />

Xie 21 , Hong Ren 22 , Jidong Jia 23 ; 1 Beijing YouAn Hospital, Capital<br />

Medical University, Beijing, China; 2 Xiamen Hospital of Traditional<br />

Chinese Medicine, Xiamen, China; 3 The Third People’s<br />

Hospital of KunMing City, Kunming, China; 4 The First Affiliated<br />

Hospital of Anhui Medical University, Hefei, China; 5 The 211<br />

Hospital of People’s Liberation Army, Haerbin, China; 6 Beijing<br />

Ditan Hospital, Capital Medical University, Beijing, China; 7 Shenzhen<br />

Third People’s Hospital, Shenzhen, China; 8 QingDao Infectious<br />

Diseases Hospital, Qingdao, China; 9 Shengjing Hospital of<br />

China Medical University, Shenyang, China; 10 General Hospital<br />

of Dagang Oilfield, Tianjin, China; 11 The Fifth Hospital of Shijiazhuang,<br />

Shijiazhuang, China; 12 Hepatology Hospital of Jilin<br />

Province, Changchun, China; 13 The Sixth People’s Hospital of<br />

Shenyang, Shenyang, China; 14 The Second Hospital of Nanjing,<br />

Nanjing, China; 15 Ningbo No. 2 Hospital, Ningbo, China; 16 The<br />

Fifth People’s Hospital of Suzhou, Suzhou, China; 17 Ruian People’s<br />

Hospital, Ruian, China; 18 The Third People’s hospital of Zhenjiang,<br />

Zhenjiang, China; 19 The First Affiliated Hospital of Zhejing<br />

University, Hangzhou, China; 20 The Third Affiliated Hospital, Sun<br />

Yat-sen University, Guangzhou, China; 21 Shanghai Roche Pharmaceuticals<br />

Ltds., Shanghai, China; 22 The Second Affiliated Hospital<br />

of Chongqing Medical University, Chongqing, China; 23 Beijing<br />

Friendship Hospital, Capital Medical University, Beijing, China<br />

Objectives: To evaluate the efficacy and safety of Peginterferon<br />

alfa (Peg-IFNα) in Chinese patients with HBeAg negative<br />

chronic hepatitis B (CHB) in routine clinical practice. Methods:<br />

Chinese adult HBeAg negative CHB patients, not co-infected<br />

with HIV, HAV, or HCV, not on concomitant telbivudine therapy,<br />

and treated with Peg-IFNα (dosing in accordance with<br />

approved SmPC in China and as per Chinese standard clinical<br />

practice) were included in a prospective, multicenter, observational,<br />

non-interventional cohort study. An interim analysis<br />

of efficacy and safety of Peg-IFNα was performed for the first<br />

50% patients at 48 weeks on-treatment (end of standard treatment).<br />

Results: The study included 503 patients (80.7% of male)<br />

with mean age of 37.5 years, 2.84 log IU/mL mean baseline<br />

HBsAg and 5.27 log IU/mL mean baseline HBV-DNA. Of<br />

these, 411 patients (81.7%) completed the 48-week treatment.<br />

Response rate of HBV-DNA suppression (HBV-DNA


1208A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

CLUSION: The baseline HBV DNA level and the duration of<br />

consolidation therapy in inactive carriers were indicators for<br />

SOVR after stopping preemptive AVT.<br />

Disclosures:<br />

The following authors have nothing to disclose: Seung Hyeon Jang, Hyo Jin Kim,<br />

jung hee kim, Dong Hyun Sinn, Geum-Youn Gwak, Yong-Han Paik, Moon Seok<br />

Choi, Joon Hyeok Lee, Kwang Cheol Koh, Seung Woon Paik, Byung Chul Yoo<br />

2050<br />

Prediction of sustained off-treatment response in chronic<br />

hepatitis B patients undergoing pegylated interferon<br />

treatment by a new scoring model<br />

Yin-Chen Wang 1 , Yi-Hsiang Huang 1,3 , Sien-Sing Yang 2 , Chien-<br />

Wei Su 1 , Kuei-Chuan Lee 1 , Han-Chieh Lin 1 ; 1 Taipei Veterans<br />

General Hospital, Division of Gastroenterology, Department of<br />

Medicine, Taipei City, Taiwan; 2 Cathay General Hospital Medical<br />

Center, Liver Center, Taipei, Taiwan; 3 National Yang-Ming University,<br />

Institute of Clinical Medicine, Taipei, Taiwan<br />

Background The pegylated interferon-α-2a (PEG-IFN-α2a) is<br />

one of the first line treatment options for chronic hepatitis B<br />

(CHB). However, the overall response rate is not satisfactory<br />

mainly due to the lack of a practical model to select patients<br />

who would have the best chance of response. The aim of this<br />

study was to develop a new simple scoring model to predict sustained<br />

off-treatment response of PEG-IFN-α2a in CHB patients.<br />

Methods We retrospectively reviewed consecutive 201 CHB<br />

patients who underwent PEG-IFN-α2a treatment from Taipei<br />

Veterans General Hospital and Cathay General Hospital. Of<br />

them, 118 were hepatitis B e antigen (HBeAg)-positive and 83<br />

were HBeAg-negative. Sustained off-treatment response was<br />

defined as HBeAg seroconversion in accompanied with HBV<br />

DNA less than 2,000 IU/ml at 48 weeks post treatment for<br />

HBeAg-positive patients, and HBV DNA less than 2,000 IU/ml<br />

at 48 weeks post treatment for HBeAg-positive patients. Multivariate<br />

Logistic regression was performed to determine factors<br />

associated with the response. Results For HBeAg-positive CHB<br />

patients, baseline hepatitis B s antigen (HBsAg) 5 ULN CK elevation and progressive<br />

muscular symptoms, it was planned to evaluate the course of<br />

treatment to stop or to go on. Results Total 178 patients (105<br />

male, 73 female) were included to the study. Mean age was<br />

39.9±11.4 year (17-72 years) and mean duration of therapy<br />

25 months (6-40 months). Mean serum CK level at the admission<br />

was 133±106 U/L (34-823 U/L). In 66 patients (37%)<br />

CK level was remained normal throughout the treatment. The<br />

highest level of CK was 2221 U/L during the treatment. Significant<br />

elevation of CK levels were observed after ninth month of<br />

therapy (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1209A<br />

2052<br />

Biologic Responses to GS-4774 in Virally-Suppressed<br />

Chronic HBV Patients<br />

Carlo Ferrari 1 , Edward J. Gane 2 , Steven-Huy B. Han 3 , W. Jeffrey<br />

Fessel 4 , Huy N. Trinh 5 , Tim Rodell 6 , Anuj Gaggar 7 , Benedetta<br />

Massetto 7 , Ondrej Podlaha 7 , Ann D. Johnson 7 , Leanne Peiser 7 ,<br />

Jacky Woo 7 , Mani Subramanian 7 , John G. McHutchison 7 , Stuart<br />

C. Gordon 8 , Calvin Q. Pan 9 , Hannah Lee 10 , Anna S. Lok 11 ;<br />

1 University of Parma, Parma, Italy; 2 Auckland General Hospital,<br />

Auckland, New Zealand; 3 Ronald Reagan UCLA Medical Center,<br />

Los Angeles, CA; 4 Kaiser Permanente, San Francisco, CA; 5 San<br />

Jose Gastroenterology, San Jose, CA; 6 Globeimmune, Louisville,<br />

CO; 7 Gilead Sciences, Inc., Foster City, CA; 8 Henry Ford Hospital,<br />

Detroit, MI; 9 Medical Procare, PLLC, Flushing, NY; 10 Tufts Medical<br />

Center, Boston, MA; 11 University of Michigan Health System, Ann<br />

Arbor, MI<br />

Background: Chronic HBV (CHB) infection is characterized by<br />

a dysfunctional HBV-specific T-cell response that permits the<br />

persistence of infected hepatocytes. GS-4774 is a heat-inactivated<br />

yeast-based therapeutic T-cell vaccine that expresses a<br />

recombinant protein comprised of the HBV X, surface, and core<br />

proteins designed to elicit an HBV-specific T-cell response. This<br />

study evaluated GS-4774 in virally-suppressed CHB patients.<br />

Methodology: 178 patients with CHB were randomized to<br />

continue oral antiviral treatment (OAV) alone or to continue<br />

OAV and receive GS-4774 (2, 10 or 40 YU; 1YU=10^7 yeast<br />

cells) every 4 weeks for 6 doses. At Week 48 ex vivo IFN-γ<br />

ELISpot was done using PBMCs and recombinant proteins and<br />

overlapping 15 mer peptides spanning the HBV core, S and X<br />

regions in GS-4774. ELISpot response was quantified as total<br />

spots/10^6 cells. Flow cytometry was performed at baseline,<br />

Weeks 12, 24, and 48 using whole blood. RNASeq from<br />

whole blood was performed on a subset of patients (N=30) at<br />

baseline and at Week 24. Results: ELISpot results were available<br />

for 108 patients. The strength of response, as determined<br />

by absolute IFN-γ spot counts to HBV protein and peptides, was<br />

generally higher among GS-4774 treated patients particularly<br />

among the HBeAg- patients, compared to controls. These differences<br />

were most consistent with HBcAg proteins and peptides<br />

and was significant among HBeAg- patients (p


1210A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

levels increased significantly during therapy (8.1 vs. 11.9 vs.<br />

15.5ng/ml, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1211A<br />

2056<br />

Long-term Renal Safety of Adefovir Dipivoxil Treatment<br />

: Focused on development of Fanconi Syndrome<br />

Sung Won Cho, Soo Yeon Cho, Choong Kyun Noh, Soon Sun<br />

Kim, Dukyong Yoon, Rae Woong Park, In-Whee Park; Ajou University<br />

School of Medicine, Suwon, Korea (the Republic of)<br />

Background Adefovir dipivoxil (ADV) is a commonly used<br />

antiviral agent for the treatment of hepatitis B virus or human<br />

immunodeficiency virus infection. ADV exhibits dose-dependent<br />

nephrotoxicity that can cause Fanconi syndrome. Fanconi<br />

syndrome is a generalized dysfunction of the proximal renal<br />

tubule resulting in impaired reabsorption and increased urinary<br />

loss of phosphate and other solutes such as uric acid, glucose,<br />

amino acids and bicarbonate. Although low-dose ADV therapy<br />

(10mg/day) has been reported to be safe, there are increasing<br />

number of reports stating that the long-term use of lowdose<br />

ADV causes Fanconi syndrome. Therefore, we assessed<br />

the incidence and clinical characteristics of Fanconi syndrome<br />

during long-term ADV therapy of hepatitis B virus or human<br />

immunodeficiency virus infection. Methods We constructed a<br />

retrospective cohort composed of a total of 619 patients who<br />

took ADV for more than 30 days in a single tertiary teaching<br />

hospital between Jun 1994 and Dec 2013. Diagnosis of Fanconi<br />

syndrome required de novo appearance of at least two<br />

of three features: hypophosphatemia (below 2.5 mg/dL or<br />

decrement at least 0.5 mg/dL from baseline), hypouricemia<br />

(below 3.7 mg/dL or decline at least 0.5 mg/dL from baseline),<br />

or serum creatinine elevation (above 1.2 mg/dL or by at<br />

least 0.5 mg/dL from baseline). Results During the mean 27.6<br />

± 21.6 months ADV-treated period, 44 patients (7.1%) exhibited<br />

hypophosphatemia. Fourty-eight patients (7.8%) showed<br />

hypouricemia and 17 patients (2.7%) presented serum creatinine<br />

elevation. Among these patients, 6 patients (0.96%)<br />

developed Fanconi syndrome. Median time to development<br />

of Fanconi syndrome was 17.5 (range 15.17–35.07) months.<br />

All patients were treated with ADV for chronic hepatitis B.<br />

Conclusion Although Fanconi syndrome developed very rarely<br />

during low dose of ADV treatment, it occurred certainly in some<br />

patients. Physicians should carefully monitor the renal function<br />

and the level of serum phosphate and uric acid during the ADV<br />

treatment in particular after 1 year.<br />

Disclosures:<br />

The following authors have nothing to disclose: Sung Won Cho, Soo Yeon Cho,<br />

Choong Kyun Noh, Soon Sun Kim, Dukyong Yoon, Rae Woong Park, In-Whee<br />

Park<br />

2057<br />

The relationship between patient’s adherence to antiviral<br />

therapy and virological response in chronic hepatitis<br />

B and hepatitis C: Results of a 48-week observational<br />

study<br />

Iftihar Koksal, Seval Sonmez; Infectious Diseases, Black Sea Technical<br />

University Faculty of Medicine, Trabzon, Turkey<br />

Background:Patient’s adherence to antiviral drugs affects<br />

the success of chronic hepatitis B(CHB) and hepatitisC(CHC)<br />

treatment.In HBV, ‘non-adherence’ primarily refers to patientmissed<br />

doses.In HCV, in contrast,‘non-adherence’ primarily<br />

refers to dose reductions made by the clinician and early<br />

treatment discontinuation due to side-effects. Aim: To investigate<br />

the relationship between antiviral medication adherence<br />

and virological response in CHB & CHC patients. Materials–<br />

methods: Ninety-seven treated CHB (entecavir or tenofovir)<br />

and CHC (peg-IFN+ribavirin) patients were followed up for<br />

48 weeks in this prospective,non-interventional observational<br />

study.Patients were evaluated for virological response at the<br />

beginning of treatment and at the 4 th (CHC patients only),<br />

12 th , 24 th and 48 th weeks.Adherence was evaluated by counting<br />

prescribed drugs at each follow-up and through patient<br />

adherence questionnaires. Results: Twenty-one patients were<br />

excluded from the study.Of the 76 evaluated patients, 60.5%<br />

(n=46) had CHB and 39.5% (n=30) had CHC; 42.1% of the<br />

patients were women and 57.9% were men.Mean age was<br />

45.18±13.2.Based on data obtained from drug counts, 83.9%<br />

of the patients had at least one delay in drug administration at<br />

each visit.The most common causes of delayed drug administrations<br />

were omission (72.8%),running out of drugs and efforts<br />

to avoid side-effects.In addition, 92.6% of chronic hepatitis<br />

patients (n=68) exhibited adherence,and 92.1% of adherent<br />

and 62.5% of non-adherent patients responded to treatment.<br />

Adherent and non-adherent patient treatment response levels<br />

were statistically significant (p


1212A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

equal to that of Pegasys 180μg. Results: 32 subjects entered<br />

the trial and 31 subjects completed study treatment. Groups<br />

were well generally matched (65% male, 12.5% HBeAg-negative,<br />

mean HBV DNA approximately 7.0 log10 IU/mL) at<br />

each group with HBV genotypes reflective of the population.<br />

rHSA/IFNα2a and Pegasys have shown antiviral activity and<br />

acceptable safety, tolerability. rHSA/IFNα2a was better tolerated<br />

than Pegasys such as neutrophil absolute value, white<br />

blood cell count, platelet count, hemoglobin and albumin<br />

decreased more at Pegasys group. Mean changes in serum<br />

HBV DNA were -1.32-2.13-1.10-2.48 log10 IU/mL for 600,<br />

750, and 900μg groups and Pegasys group after treatment.<br />

E antigen quantitation decreased more and normalization rate<br />

of ALT and AST were higher at rHSA/IFNα2a group. Half-life<br />

(t1/2) of rHSA/IFNα2a is 120 to 140 hours and duration of<br />

antiviral activity that indicate potential suitability for dosing<br />

intervals of 2 weeks or longer. Serum IFNα2a area under the<br />

curve and maximum concentration increased with dose escalating.<br />

IFNα2a Accumulation is not obvious at each group.<br />

Pharmacokinetics of the last dose except t1/2 and Kinetics of<br />

viral decline were similar between rHSA/IFNα2a 750μg and<br />

Pegasys. Conclusions: rHSA/IFNα2a was safe and well tolerated;<br />

declines in HBV DNA were similar to Pegasys evaluated.<br />

rHSA/IFNα2a 750μg has been selected for further hepatitis B<br />

clinical development.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hong Zhang, Min Wu, Qingmei<br />

Li, xiaojiao li, Yanhua Ding, Junqi Niu<br />

2059<br />

Association between ALT flares and HBeAg loss and<br />

HBsAg decline in Patients with Chronic Hepatitis B<br />

during treatment with Tenofovir Disoproxil Fumarate or<br />

Adefovir Dipivoxil<br />

Patrick Marcellin 1 , Edward J. Gane 2 , Zahary Krastev 3 , Selim<br />

Gurel 4 , Geoffrey M. Dusheiko 5 , Anuj Gaggar 6 , Benedetta Massetto<br />

6 , Kyungpil Kim 6 , John F. Flaherty 6 , Mani Subramanian 6 ,<br />

Harry L. Janssen 7 , Maria Buti 8 ; 1 Hôpital Beaujon, Clichy, France;<br />

2 Auckland City Hospital, Auckland, New Zealand; 3 University<br />

Hospital St. Ivan Rilsky, Sofia, Bulgaria; 4 Uludag Universitesi Tip<br />

Fakultesi, Bursa, Turkey; 5 Royal Free Hospital, London, United<br />

Kingdom; 6 Gilead Sciences, Inc., Foster City, CA; 7 University of<br />

Toronto, Toronto, ON, Canada; 8 Universitari Vall d’Hebron and<br />

Ciberehd, Barcelona, Spain<br />

Background and Aims: Increase in ALT levels (flares) are<br />

observed after the initiation of oral antiviral therapy and<br />

thought to be immune-mediated. We evaluated the association<br />

between ALT flares during the first 48 weeks of treatment with<br />

tenofovir disoproxil fumarate (TDF) or adefovir dipivoxil (ADV)<br />

and HBeAg loss and HBsAg decline. Methods: HBsAg levels<br />

were determined by Abbott Assay (linear range 0.05-250<br />

IU/mL; quantification test on undiluted samples, and at dilutions<br />

of 1:500 and 1:999) in 245 and 359 chronic infected<br />

HBeAg-positive and HBeAg-negative patients, respectively. ALT<br />

flare was defined as ALT >2x baseline and >5x upper limit<br />

of normal during the first 48 weeks of treatment. HBeAg loss<br />

and HBsAg decline were evaluated 24 weeks after a flare.<br />

Logistic regression analyses were performed to examine host<br />

and viral baseline factors (including viral load, HBV genotype,<br />

HBeAg status and HBsAg level) associated with flares. Significant<br />

factors in univariate analyses were examined in a multivariate<br />

model. Results: 23 HBeAg-positive patients (9%) and 5<br />

HBeAg-negative patients (1%) experienced ALT flares during<br />

the first 48 weeks of treatment. All ALT flares were considered<br />

not clinically significant except for one which was reported as<br />

a transient mild adverse event. Among HBeAg-positive patients,<br />

HBeAg loss and HBsAg decline were greater in patients who<br />

had ALT flares (Table, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1213A<br />

2060<br />

Antiviral treatment for chronic hepatitis B in Australia: a<br />

nationwide analysis of treatment uptake and prescribing<br />

trends according to individual antiviral agent<br />

Jennifer H. MacLachlan 1,2 , Nicole Allard 1,2 , Benjamin C. Cowie 1,2 ;<br />

1 WHO Collaborating Centre for Viral Hepatitis, Doherty Institute<br />

for Infection and Immunity, Melbourne, VIC, Australia; 2 Department<br />

of Medicine, Dentistry and Health Sciences, University of<br />

Melbourne, Melbourne, VIC, Australia<br />

Background: Antiviral treatment for chronic hepatitis B (CHB)<br />

is subsidized through Australia’s national health insurance<br />

scheme, Medicare, with all approved treatments funded for<br />

patients meeting specified virological and disease activity<br />

criteria. However, awareness and uptake of CHB therapy is<br />

low, and prescribing patterns and adherence to treatment<br />

guidelines have not been examined at a population level.<br />

Methods: Aggregate, population-level data regarding expenditure<br />

for CHB treatment through Medicare were obtained<br />

for all approved therapies (adefovir, entecavir, lamivudine,<br />

pegylated interferon, telbivudine, and tenofovir). Numbers of<br />

patients receiving therapy were derived from these data, and<br />

prescribing trends by individual agent were assessed for the<br />

period 2011-2013. Individualized data were also obtained<br />

for 2013 to examine prescribing patterns for combination<br />

antiviral therapy. Results: The number of patients treated for<br />

CHB in Australia increased from 9,000 in 2011 to 10,900<br />

in 2013, for an estimated proportion of all people living with<br />

CHB in 2013 of 5%. Patients receiving recommended first-line<br />

oral therapies represented an increasing majority, with those<br />

receiving entecavir increasing from 44.1% to 46.8% and tenofovir<br />

increasing from 26.3% to 34.6%. Usage of lamivudine<br />

and adefovir decreased, however together these drugs still<br />

represented 17.4% of prescribing expenditure in 2013. Very<br />

small numbers of patients received either telbivudine (36 months, consolidation<br />

duration>12 months and baseline HBV DNA level


1214A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

mechanism triggered by introduction of small interfering RNA<br />

(siRNA). Tekmira’s clinically-validated lipid nanoparticle (LNP)<br />

platform enables effective delivery of siRNAs in vivo and in<br />

vitro. HDV-targeted siRNA combined with LNP technology may<br />

thus provide a novel opportunity to alleviate HDV pathogenesis.<br />

There is limited data regarding which of the HDV RNA<br />

species are susceptible to gene silencing. siRNAs targeting<br />

positive or negative strand HDV RNAs were designed and<br />

delivered to the Huh7-D12 stable HDV-expressing cell line<br />

using LNP technology. HDV RNAs were quantified using a<br />

branched DNA assay and nuclear/cytoplasmic fractionation,<br />

and HDAg protein was evaluated by ELISA. Rapid inhibitory<br />

effects were detected 1 day after LNP treatment, with HDAg<br />

mRNA-targeted siRNA showing a marked dose-dependent<br />

inhibitory effect on HDV positive strand RNAs but not on the<br />

negative strand viral genome. The nuclear/cytoplasmic fractionated<br />

RNA assay demonstrated that the cytosolic HDV RNAs<br />

were more susceptible to siRNA than nuclear HDV RNAs. In<br />

contrast to HDAg mRNA-targeted siRNAs that demonstrated<br />

gene silencing at day 1, genome-targeted siRNAs did not<br />

demonstrate gene silencing until day 4, hinting at differences<br />

in kinetics between the two siRNA targeting strategies. HDV-targeted<br />

siRNA-LNP effects were durable, with greater than 1-log<br />

reductions of viral RNAs maintained at even 21 days after a<br />

single treatment. Similarly, a 1-log reduction in HDAg protein<br />

was achieved out to day 12 so far and more extended time<br />

points are to be investigated. In conclusion, a direct HDV-targeted<br />

siRNA-LNP approach can effectively suppress positive<br />

and negative strand HDV RNAs and HDAg protein in vitro, and<br />

provides a promising novel strategy to treat HDV infection. The<br />

efficacy of direct HDV targeting relative to indirect effects from<br />

HBV gene silencing are currently under investigation.<br />

Disclosures:<br />

Xin Ye - Employment: Tekmira Pharmaceuticals Corporation<br />

Nicholas M. Snead - Employment: Tekmira Pharmaceuticals<br />

Trisha R. Barnard - Employment: Tekmira Pharmaceuticals Corporation<br />

Emily P. Thi - Employment: Tekmira Pharmaceuticals<br />

Amy C. Lee - Employment: Tekmira<br />

The following authors have nothing to disclose: Ian MacLachlan<br />

2063<br />

Serum hepatitis B core-related antigen as a treatment<br />

predictor during pegylated interferon therapy in<br />

patients with HBeAg-positive chronic hepatitis B<br />

Natthaya Chuaypen 1 , Nawarat Posuwan 2 , Sunchai Payungporn 1 ,<br />

Yasuhito Tanaka 3 , Yong Poovorawan 4 , Pisit Tangkijvanich 1 ; 1 Biochemistry,<br />

Chulalongkorn University, Bangkok, Thailand; 2 Chulalongkorn<br />

University, Bangkok, Thailand; 3 Virology and Liver unit,<br />

Nagoya City University Graduate School of Medical Sciences,<br />

Nagoya, Japan; 4 Paediatrics, Chulalongkorn University, Bangkok,<br />

Thailand<br />

Background: The role of quantitative serum hepatitis B core-related<br />

antigen (HBcrAg) in patients with chronic hepatitis B<br />

(CHB) receiving pegylated interferon (PEG-IFN) has never been<br />

investigated. The aims of this study were to assess the correlation<br />

of HBcrAg with intrahepatic covalently closed circular<br />

DNA (cccDNA), and compare its usefulness with quantitative<br />

serum HBsAg in patients with HBeAg-positive CHB during<br />

PEG-IFN therapy. Method: We retrospectively analyzed data<br />

of Thai patients with HBeAg-positive CHB treated with PEG-<br />

IFN for 48 weeks. Virological response (VR) was defined as<br />

HBeAg clearance and HBV DNA < 2,000 IU/mL at 24 weeks<br />

post treatment. Paired liver biopsies at weeks 0 and 48 were<br />

analyzed for cccDNA by real-time PCR. HBsAg and HBcrAg<br />

levels were assessed at weeks 0, 4, 12, 24, 48, and 72 by<br />

automated chemiluminescent immunoassays. HBV genotype<br />

was performed by direct sequencing. Results: A total of 46<br />

patients (31 male, mean age 33.2 years) were enrolled.<br />

The distribution of HBV genotypes B and C was 10.9% and<br />

89.1%, respectively. VR was achieved in 15 (32.6%) patients.<br />

Responders had higher cccDNA decline compared with non-responders<br />

(1.7±0.8 vs 0.4±1.1 log 10<br />

copies/cell, P=0.001),<br />

although baseline cccDNA were comparable between groups.<br />

Baseline HBsAg and HBcrAg were significantly correlated<br />

with cccDNA (Pearson correlation, r=0.450, P=0.013 and<br />

r=0.631, P20,000 IU/mL as the cut-off level, the<br />

negative predictive value (NPV) of achieving VR at weeks 12<br />

and 24 were 72.7% and 100%, respectively. For HBcrAg, the<br />

best cut-off level was log 10<br />

8.0 IU/mL, which provided NPV at<br />

week 12 and 24 of 89.5% and 100%, respectively. Conclusion:<br />

These results demonstrated that the convenient quantitative<br />

HBcrAg represented a good serum marker of intrahepatic<br />

cccDNA in patients with HBeAg-positive CHB, which was comparable<br />

with that of quantitative HBsAg. Monitoring HBcrAg<br />

levels during PEG-IFN therapy may be useful to identify patients<br />

with a very low probability of treatment response.<br />

Disclosures:<br />

Yasuhito Tanaka - Grant/Research Support: Chugai Pharmaceutical CO., LTD.,<br />

MSD, Bristol-Myers Squibb; Speaking and Teaching: janssen pharma, Bristol-Myers<br />

Squibb<br />

The following authors have nothing to disclose: Natthaya Chuaypen, Nawarat<br />

Posuwan, Sunchai Payungporn, Yong Poovorawan, Pisit Tangkijvanich<br />

2064<br />

Encapsidation and secretion of HBV RNA can be inhibited<br />

by Core Inhibitors but not by Nucleoside Analogs<br />

Angela Lam, Suping Ren, Christine Espiritu, Mollie Kelly, Vincent<br />

Lau, Robert Vogel, George D. Hartman, Lalo Flores, Klaus Klumpp;<br />

Novira Therapeutics Inc., Doyleston, PA<br />

Background: Infectious HBV particles contain a partially double<br />

stranded DNA genome that is replicated in infected hepatocytes<br />

through reverse transcription of HBV pre-genomic RNA<br />

(pgRNA). HBV capsid, which encloses pgRNA and viral polymerase,<br />

is assembled from HBV core protein building blocks<br />

and is the site of HBV DNA synthesis. HBV core inhibitors<br />

are therefore able to potently inhibit HBV replication. In the<br />

current study, the effect of the HBV core inhibitor NVR-3891 on<br />

HBV RNA encapsidation and secretion was examined along<br />

with other core inhibitors and nucleoside analogs Lamivudine<br />

(LMV) or Tenofovir (TFV). Methods: Intracellular encapsidated<br />

HBV pgRNA was examined by Northern blot. HBV DNA and<br />

RNA levels were determined using Quantigene assays in HBV<br />

infected HepaRG cells and human serum samples. The nature<br />

and function of HBV RNA containing particles was evaluated<br />

using detergent and nuclease treatments, and analytical gradient<br />

centrifugation. Results: NVR-3891 induced HBV core<br />

protein mis-assembly in vitro and inhibited HBV replication<br />

with a mean EC 50<br />

of 1.9 mM in persistently infected differentiated<br />

HepaRG cells. Northern blot analysis of HBV infected<br />

cells showed that pgRNA encapsidation was inhibited by NVR-<br />

3891, while treatment of infected cells with nucleoside analogs<br />

led to an increase in encapsidated pgRNA. Analysis of<br />

the supernatants of infected cells and human sera from HBV<br />

infected patients showed high levels of HBV RNA containing


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1215A<br />

enveloped virus particles. Both HBV DNA and RNA containing<br />

particles fractionated similarly by density gradient centrifugation.<br />

The formation of these secreted RNA-containing viral particles<br />

could be blocked by treatment with core inhibitors, but<br />

not with nucleoside analogs. Conclusions: HBV RNA containing<br />

enveloped particles were observed in the supernatant of HBV<br />

infected cells and in the sera of HBV infected patients. Core<br />

inhibitors and nucleoside analogs could block HBV DNA production<br />

in HBV infected cells, but only core modulators could<br />

block the formation of HBV RNA containing particles. The more<br />

complete block of enveloped HBV particle formation (RNA and<br />

DNA particles) differentiates HBV core inhibitors from nucleoside<br />

analogs, which inhibit formation of only DNA-containing<br />

particles. Inhibitory effects on both HBV DNA and RNA particles<br />

could contribute to increased efficacy of HBV core inhibitor<br />

based treatment of chronic hepatitis B patients alone or in<br />

combination with nucleoside analogs.<br />

Disclosures:<br />

Angela Lam - Employment: Novira Therapeutics, Merck & Co<br />

Christine Espiritu - Employment: Novira Therapeutics<br />

Mollie Kelly - Independent Contractor: Novira Therapeutics<br />

George D. Hartman - Management Position: Novira Therapeutics<br />

Klaus Klumpp - Board Membership: Riboscience LLC; Employment: Novira Therapeutics<br />

Inc<br />

The following authors have nothing to disclose: Suping Ren, Vincent Lau, Robert<br />

Vogel, Lalo Flores<br />

2065<br />

Development and Recurrence of Hepatocellular Carcinoma<br />

in Patients with Chronic Hepatitis B Treated with a<br />

Nucleic Acid Analog<br />

Itaru Ozeki, Tomoaki Nakajima, Shuhei Hige, Yoshiyasu Karino,<br />

Joji Toyota; Department of Gastroenterology, Sapporo Kosei General<br />

Hospital, Sapporo, Japan<br />

Abstract Objectives: (1) To analyze the incidence of hepatocellular<br />

carcinoma (HCC) and factors contributing to carcinogenesis<br />

in patients treated with a nucleic acid analog (NA)<br />

and (2) to examine the background factors of HCC patients<br />

according to tumor recurrence. Materials and Methods: This<br />

study included 469 patients who received NA at our hospital<br />

for ≥1 year (305 who received entecavir for the first time<br />

and 164 who started lamivudine). Patients with a history of<br />

HCC or who received NA for


1216A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

as proportions of resistances mutations maintain a relatively<br />

low level. Genotypic resistance to ETV was detected in all the<br />

partial responders with long-term therapy. The threshold related<br />

to substisutions requires corroboration by further <strong>studies</strong> with<br />

large samples to determine whether adapt the therapy.In addition<br />

to the common substitutions, the unknown amino acid substitutions,<br />

such as rtL145M/S, rtF151Y/L,rtR153Q, rtI224V,<br />

rtN248H, rtS223A, rtS256C,are also likely to influence the<br />

outcome of treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xiaxia Zhang, Minran Li, Ying<br />

Cao, Yu Zhang, Renwen Zhang, Yang Xu, Fang Li, Xiaoyuan Xu<br />

2067<br />

A Broad Spectrum Antibiotic Therapy As Empirical<br />

Treatment In Healthcare-Associated Infections Improves<br />

Survival In Cirrhotic Patients: A Randomized Trial<br />

Cristina Lucidi, Vincenza Di Gregorio, Barbara Lattanzi, Valerio<br />

Giannelli, Michela Giusto, Oliviero Riggio, Mario Venditti, Manuela<br />

Merli; Sapienza University of Rome, Rome, Italy<br />

Background and aims: An early diagnosis and an appropriate<br />

treatment of infections in cirrhosis are crucial due to their high<br />

morbidity and mortality. Multidrug-resistant (MDR) infections<br />

are growing in healthcare setting. Frequently, Healthcare-associated<br />

(HCA) infections are treated as community-acquired<br />

with a detrimental effect on survival. We aimed to prospectively<br />

evaluate for the first time in a randomized trial the effectiveness<br />

of a broad spectrum antibiotic treatment cirrhotic<br />

patients with HCA infections. Methods: Consecutive cirrhotic<br />

patients with HCA infections hospitalized in the last 3 years<br />

in our ward were enrolled. After culture sampling, patients<br />

were promptly randomized to empirically receive a standard<br />

or a broad spectrum antibiotic treatment (NCT01820026). The<br />

primary endpoint was the in-hospital mortality with the aim of<br />

66% reduction. Efficacy, sides effects, development of second<br />

infections and length of hospitalization was also considered.<br />

Treatment failure was followed by modification of therapies<br />

as appropriate. Results: As originally designed, 96 patients<br />

were randomized but 94 completed the treatment and were<br />

analyzed (two were excluded for incorrect randomization due<br />

to neoplastic ascites). Patients belonging to standard (48) and<br />

broad spectrum (46) groups were similar for demographic,<br />

clinical, and microbiological characteristics. The prevalence of<br />

MDR pathogens was similar in the 60 microbiologically documented<br />

infections (40% in standard vs 46% in broad spectrum<br />

group, p=ns). In-hospital mortality showed a 76% reduction<br />

in the broad spectrum vs standard group (respectively, 6% vs<br />

25%; p=0.01); the shift to another antibiotic treatment was not<br />

possible in about 20% of death patients. The standard therapy<br />

failed more often than the broad spectrum one (51 vs 18%,<br />

p=0.001). The length of the hospitalization was longer in the<br />

Standard Group (18 ± 15 days) than in the Broad Spectrum<br />

Group (12.3 ± 7 days) (p= 0.03). Only one patient showed<br />

a not severe side effect related to antibiotic treatment. Ten<br />

patients (12% of the group 1 and 9% of the group 2) developed<br />

a second episode of infection during the hospitalization<br />

with a similar prevalence of MDR of (60% in standard vs<br />

50% in broad spectrum, p=ns). Conclusions: A broad spectrum<br />

antibiotic therapy as empirical treatment in HCA infections<br />

improves survival in cirrhotic patients. This treatment resulted<br />

significantly effective, safe and cost-saving.<br />

Disclosures:<br />

The following authors have nothing to disclose: Cristina Lucidi, Vincenza Di<br />

Gregorio, Barbara Lattanzi, Valerio Giannelli, Michela Giusto, Oliviero Riggio,<br />

Mario Venditti, Manuela Merli<br />

2068<br />

Probiotic supplementation improves liver function but<br />

fails to restore neutrophil phagocytosis in stable cirrhosis.<br />

A randomized, double-blind, placebo-controlled<br />

study<br />

Angela Horvath, Bettina Leber, Bianca Schmerboeck, Monika<br />

Tawdrous, Astrid Hartl, Sandra Lemesch, Peter Fickert, Rudolf E.<br />

Stauber, Philipp Stiegler, Franziska Durchschein, Elisabeth Krones,<br />

Philipp Douschan, Gernot Zollner, Walter Spindelboeck, Karl<br />

Oettl, Doris Payerl, Vanessa Stadlbauer; Medical University Graz,<br />

Graz, Austria<br />

A multispecies probiotic was administered to cirrhotic patients<br />

in order to strengthen gut barrier function, decrease endotoxin<br />

load and ultimately restore phagocytic dysfunction of neutrophils<br />

that is commonly found in cirrhosis. Therefore, stable<br />

cirrhotic patients received either a mixture of Bifidobacteria bifidum<br />

and lactis, Lactobacili acidophilus, brevis, casei and salivarius,<br />

and Lactococcus lactis sp. (n=44) or a placebo (n=36)<br />

for 6 months. We found that during the intervention period<br />

the percentage of patients with Child-Pugh grade A increased<br />

significantly from 63% to 70% after three months and 66%<br />

after six months, only in the probiotic group. Of 16 patients<br />

in the probiotic group that started the study with Child-Pugh<br />

score 7 or higher, 6 improved their score after 6 months of<br />

probiotics, 7 did not change but only 3 deteriorated. Median<br />

Child-Pugh score decreased from 6 to 5 after three months of<br />

probiotics and MELD score decreased non-significantly from<br />

12 to 11 during the intervention period but returned to baseline<br />

6 months after intervention had ended. Gut permeability,<br />

measured with Lactulose-Mannitol ratio, improved significantly<br />

in both groups. With probiotics it remained significantly lower<br />

than baseline 6 months after the intervention has ended. On<br />

the other hand sucrose recovery, calprotectin and zonulin in<br />

stool and serum diamine oxidase stayed unchanged in both<br />

groups. Endotoxin levels decreased non-significantly with probiotics<br />

but similar changes occurred in the control group. Cytokine<br />

levels as well as soluble CD14 and LPS binding protein<br />

stayed unchanged in both groups. Probiotic supplementation<br />

did not have a significant influence on neutrophil phagocytosis.<br />

At baseline both groups had an intact phagocytic capacity but<br />

a significantly increased amount of non-phagocytic neutrophils<br />

compared to controls. Over the course of the study the amount<br />

of non-phagocytic cells remained constant while phagocytic<br />

capacity decreased to 60% of the baseline value. During the<br />

first 6 months of the study 1 severe infection occurred in the<br />

placebo group, none in the probiotic group. After intervention<br />

had ended 5 severe infections were reported in the probiotic<br />

group and 4 in the placebo group. In conclusion, probiotic<br />

supplementation is a save method to improve liver function in<br />

stable cirrhosis. However, its influence on gut barrier function<br />

and bacterial translocation in cirrhotic patients seem to be minimal.<br />

Independent of the intervention, patients gradually developed<br />

significant impairment in neutrophil phagocytosis over<br />

the course of one year and with this loss of function a higher<br />

incidence of severe infections occurred.<br />

Disclosures:<br />

Angela Horvath - Grant/Research Support: Instutut Allergosan<br />

Peter Fickert - Consulting: Falk Foundation, Falk Foundation, Falk Foundation,<br />

Falk Foundation; Speaking and Teaching: Roche Austria, Gilead Austria, MSD<br />

Austria, MERCK Austria, Roche Austria, Gilead Austria, MSD Austria, MERCK<br />

Austria, Roche Austria, Gilead Austria, MSD Austria, MERCK Austria, Roche<br />

Austria, Gilead Austria, MSD Austria, MERCK Austria<br />

Rudolf E. Stauber - Advisory Committees or Review Panels: Gilead, Janssen-Cilag,<br />

AbbVie, BMS; Grant/Research Support: MSD; Speaking and Teaching: Merz<br />

The following authors have nothing to disclose: Bettina Leber, Bianca Schmerboeck,<br />

Monika Tawdrous, Astrid Hartl, Sandra Lemesch, Philipp Stiegler, Franziska<br />

Durchschein, Elisabeth Krones, Philipp Douschan, Gernot Zollner, Walter<br />

Spindelboeck, Karl Oettl, Doris Payerl, Vanessa Stadlbauer


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1217A<br />

2069<br />

Impact of genetic variation of the AVP1a receptor on<br />

the presence of circulatory failure in patients with acute<br />

decompensation of liver cirrhosis or acute-on-chronic<br />

liver failure<br />

Annarein J. Kerbert 1 , Jelte Schaapman 1 , Johan van der Reijden 1 ,<br />

Amorós Àlex 2 , Aiden McCormick 3 , Bart van Hoek 1 , Vicente<br />

Arroyo 4 , Pere Gines 4 , Rajiv Jalan 5 , Victor Vargas 6 , Rudolf E.<br />

Stauber 7 , Hein W. Verspaget 1 , Minneke Coenraad 1 ; 1 Gastroenterology-Hepatology,<br />

Leiden University Medical Center, Leiden,<br />

Netherlands; 2 Data Management Center, CLIF consortium, Barcelona,<br />

Spain; 3 Gastroenterology-Hepatology, St Vincent’s University<br />

Hospital, Dublin, Ireland; 4 Liver Unit, Hospital Clinic, University<br />

of Barcelona, Barcelona, Spain; 5 Gastroenterology-Hepatology,<br />

University College London, London, United Kingdom; 6 Liver Unit,<br />

Hospital Vall d’Hebron, CIBERehd, Barcelona, Spain; 7 Internal<br />

Medicine, Medical University of Graz, Graz, Austria<br />

Background & aims: Acute-on-chronic liver failure (ACLF)<br />

is defined as an acute decompensation of former stable<br />

chronic liver disease (AD) accompanied by the presence of<br />

organ failure and a high risk of short-term mortality. Systemic<br />

hemodynamic derangement and activation of endogenous<br />

vasoconstrictor systems are thought to contribute to the pathogenesis.<br />

Arginine vasopressin is a key-regulator in hemodynamic<br />

homeostasis and mediates splanchnic vasoconstriction<br />

through the arginine vasopressin 1a receptor (AVP1aR). Aim<br />

of the present study was to assess whether genetic variation of<br />

AVP1aR is associated with the presence of circulatory failure<br />

in patients with AD or ACLF. Methods: Eight single nucleotide<br />

polymorphisms (SNPs) of AVP1aR with possible clinical relevance<br />

were identified. From 824 cirrhotic patients admitted for<br />

AD, clinical, laboratory and survival data were retrieved from<br />

the CANONIC database. Presence of circulatory failure was<br />

defined as a mean arterial blood pressure (MAP) < 70 mmHg<br />

or the use of vasopressors. ACLF was defined according to<br />

the CLIF Consortium Organ Failure score. All patients were<br />

genotyped for all eight SNPs using polymerase chain reaction<br />

(PCR) followed by restriction fragment length polymorphism<br />

or PCR allele-specific amplification primers. Fisher’s exact test<br />

and linear regression analysis were used to test for association<br />

between allele frequencies and dichotomous and continuous<br />

variables respectively. Results are shown as mean ± SD. P<<br />

0.05 was considered statistically significant. Results: Patients<br />

with ACLF (n=184) had a significantly lower MAP as compared<br />

to patients without ACLF (80 ± 13 vs. 84 ± 12 mmHg,<br />

p< 0.001). The use of vasopressors was also significantly more<br />

frequent in patients with ACLF as compared to those without<br />

ACLF (19.0% vs. 2.9%, p< 0.001). Circulatory failure was<br />

present in 61 out of 824 patients, of whom 44 fulfilled the<br />

criteria of ACLF. A C>T mutation in SNP rs7308855 showed<br />

a significant association with the presence of circulatory failure<br />

in patients with AD (p= 0.025) and a clear trend towards the<br />

presence of circulatory failure in patients with ACLF (p= 0.085).<br />

A trend was also found for a T>A mutation in SNP rs7298346<br />

to be associated with the presence of circulatory failure in<br />

patients with AD (p= 0.062). In addition, this mutation showed<br />

a significant association with the presence of circulatory failure<br />

in the subgroup of patients with ACLF (p= 0.046). Conclusions:<br />

Single nucleotide polymorphisms in the AVP1a receptor are<br />

associated with the presence of circulatory failure in patients<br />

with acute decompensation of liver disease and ACLF.<br />

Disclosures:<br />

Bart van Hoek - Advisory Committees or Review Panels: Janssen-Cilag, Bristol<br />

Meyers Squib, Gilead, Merck, Abbvie<br />

Vicente Arroyo - Speaking and Teaching: GRIFOLS<br />

Pere Gines - Advisory Committees or Review Panels: Ferring, Ikaria; Grant/<br />

Research Support: Sequana Medical, Grifols<br />

Rajiv Jalan - Consulting: Ocera Therapeutics, Conatus; Grant/Research Support:<br />

Grifols, Gambro; Patent Held/Filed: Yaqrit, Cyberliver<br />

Rudolf E. Stauber - Advisory Committees or Review Panels: Gilead, BMS; Grant/<br />

Research Support: Abbvie, MSD; Speaking and Teaching: Merz<br />

The following authors have nothing to disclose: Annarein J. Kerbert, Jelte Schaapman,<br />

Johan van der Reijden, Amorós Àlex, Aiden McCormick, Victor Vargas,<br />

Hein W. Verspaget, Minneke Coenraad<br />

2070<br />

Prevalence, risk factors and outcome of infections by<br />

multidrug resistant bacteria (MDR) in a liver intensive<br />

care unit (ICU): a prospective study<br />

Amrish Sahney 1 , Cyriac A. Philips 1 , Rakhi Maiwall 1 , Ashok<br />

Choudhary 1 , Lalita G. Mitra 2 , Ankur Jindal 1 , Manoj Kumar 1 , Kapil<br />

D. Jamwal 1 , Vikram Bhatia 1 , Vikas Khillan 3 , Shiv K. Sarin 1 ; 1 Hepatology,<br />

ILBS, New Delhi, India; 2 Critical care, ILBS, New Delhi,<br />

India; 3 Microbiology, ILBS, New Delhi, India<br />

Background & Aim:There is limited data on prevalence, predictors<br />

and impact on extra hepatic organ failure and mortality<br />

due to MDR bacterial infections in critically ill cirrhotic<br />

patients. Methodology: We prospectively studied 522 cirrhoticpatients<br />

admittedto dedicated liver ICU during March<br />

to October 2014. Organ failure was defined as: Kidney failure:<br />

SCr ≥2mg% or need for dialysis, cerebral failure: Grade<br />

3 or 4 hepatic encephalopathy, Circulation failure: shock, &<br />

Lung failure: PaO2/FiO2


1218A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Amrish Sahney, Cyriac A. Philips,<br />

Rakhi Maiwall, Ashok Choudhary, Lalita G. Mitra, Ankur Jindal, Manoj Kumar,<br />

Kapil D. Jamwal, Vikram Bhatia, Vikas Khillan, Shiv K. Sarin<br />

2071<br />

The macrophage activation markers sCD206 and<br />

sCD163 predict mortality in patients with liver cirrhosis<br />

without or with acute-on-chronic liver failure (ACLF)<br />

Henning Gronbaek 1 , Sidsel Rødgaard-Hansen 2 , Niels Kristian<br />

Aagaard 1 , Elisabet García 3 , Hendrik V. Vilstrup 1 , Søren K.<br />

Moestrup 4 , Vicente Arroyo 3 , Holger J. Møller 2 ; 1 Department of<br />

Hepatology & Gastroenterology, Aarhus University Hospital, Aarhus,<br />

Denmark; 2 Department of Clinical Biochemistry, Aarhus University<br />

Hospital, Aarhus, Denmark; 3 Liver Unit, Hospital Clinic,<br />

University of Barcelona, Barcelona, Spain; 4 Department of Biomedicine,<br />

Aarhus University, Aarhus, Denmark<br />

Introduction: Activation of liver macrophages plays a key role<br />

in liver and systemic inflammation and may be involved in<br />

the development and prognosis of ACLF. We therefore measured<br />

the macrophage activation markers soluble sCD206 and<br />

sCD163 in plasma and related them to the short- (1-3 months)<br />

and long-term (6 months) mortality in the cirrhosis patients of the<br />

CANONIC study. Methods: 86 patients had no ascites and no<br />

ACLF, 580 had ascites but no ACLF, and 100, 66 and 19 had<br />

ACLF-grade I (ACLF-1), ACLF-II, and ACLF-III, respectively. The<br />

patients’ clinical course was registered and their MELD, CLIF-C<br />

Acute Decompensation (AD), and ACLF (s)scores computed at<br />

study entry. Results: We found a stepwise increase (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1219A<br />

2073<br />

The Epidemiology and Clinical Associations of Budd-<br />

Chiari Syndrome. A Review of 507 Admissions from the<br />

2012 National Inpatient Sample<br />

Jennifer Saad 1 , Shawn Shah 1 , Joseph Shatzel 1 , Harley Friedman 1 ,<br />

Neil Volk 1 , Rolland C. Dickson 2 ; 1 Internal Medicine, Dartmouth<br />

Hitchcock Medical Center, Lebanon, NH; 2 Gastroenterology, Dartmouth<br />

Hitchcock Medical Center, Lebanon, NH<br />

Background: Budd-Chiari syndrome (BCS) is a rare condition<br />

resulting from obstruction of the hepatic venous outflow tract<br />

and is strongly associated with hypercoagulable states. Large<br />

series on BCS are scarce, so epidemiological data has been<br />

derived primarily from reported case series. The aim of this<br />

study was to evaluate the epidemiology, common clinical associations,<br />

and mortality of patients admitted with BCS from a<br />

large national database. Methods: Using the 2012 National<br />

Inpatient Sample (NIS), admissions with an International Classification<br />

of Diseases, 9 th Revision, Clinical Modification code<br />

for BCS were extracted, and correlated with age, gender,<br />

length of stay, mortality and commonly associated diagnoses.<br />

Results: Of the 7,296,968 unweighted admissions in the 2012<br />

NIS, 507 were associated with BCS (prevalence 0.007%).<br />

BCS was more common in women (55.8%) with a mean age<br />

of 47 years (range 0-90). The average length of stay was 10<br />

days (range 0-125) with an in-hospital mortality of 6.5%. Of<br />

BCS admissions, 23.08% had cirrhosis of the liver, 16.77%<br />

had portal HTN, 8.68% had both cirrhosis of the liver and<br />

portal HTN, 16.3% had a hepatocellular malignancy (8.3%<br />

primary, 8.1% metastatic) and 8.9% had an acquired or inherited<br />

thrombophilia. Overall, 37.5% of patients with BCS had<br />

a new or previously diagnosed malignancy. Myeloproliferative<br />

or bone marrow disorders were associated with 8.1% of<br />

BCS admissions including polycythemia vera (3.8%), essential<br />

thrombocytosis (1.6%), paroxysmal nocturnal hemoglobinuria<br />

(1.4%), myelofibrosis (0.8%) and chronic myelogenous leukemia<br />

(0.6%). 28% of admissions had an acute thrombus at an<br />

additional anatomic site (9.9% superficial or deep veins, 8.9%<br />

inferior vena cava, 6.1% pulmonary vein and 3.2% portal<br />

vein). There was no significant difference in the age, mortality<br />

or length of stay between males and females with BCS. Males<br />

were more likely to have a primary hepatocellular malignancy<br />

(OR 3.97, 95% CI 1.95, 8.10, p


1220A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2075<br />

Circulating interleukin 6, 10 and 17 as prognostic<br />

markers in patients with liver cirrhosis<br />

Josiane Fischer 1 , Telma E. Silva 1 , Pedro E. Soares e Silva 1 , Bruno<br />

S. Colombo 1 , Letícia M. Wildner 2 , Maria Luiza Bazzo 2 , Tania<br />

S. Frode 2 , Silvana Vigil de Melo 2 , Julia S. Rosa 2 , Esther B. Dantas-Correa<br />

1 , Janaína L. Narciso-Schiavon 1 , Leonardo L. Schiavon 1 ;<br />

1 Internal Medicine, Division of Gastroenterology, Federal University<br />

of Santa Catarina, Florianópolis, Brazil; 2 Department of Clinical<br />

Analysis, Federal University of Santa Catarina, Florianopolis,<br />

Brazil<br />

Introduction: Activation of inflammatory system is present from<br />

the early stages of cirrhosis and it is associated with elevated<br />

levels of cytokines. However, data about the prognostic significance<br />

of circulating cytokines in liver cirrhosis is still lacking.<br />

We sought to investigate the prognostic significance of IL-6,<br />

IL-10 and IL-17 in patients with stable cirrhosis and in subjects<br />

admitted for acute decompensation (AD) of cirrhosis. Methods:<br />

This prospective study included two cohorts: (1) stable cirrhosis<br />

attended in the Outpatient Clinic (n = 118), and (2) subjects<br />

hospitalized for AD (n = 130). Thirty healthy subjects served as<br />

control group. The acute-on-chronic liver failure (ACLF) criteria<br />

were applied according to the EASL-CLIF Consortium. Results:<br />

IL-6 and IL-10 levels were higher in both groups of patients with<br />

cirrhosis as compared to control group and also in patients<br />

with AD in relation to stable cirrhosis (P < 0.05). In stable<br />

cirrhosis, during a median follow-up of 17 months, an event<br />

(hospitalization, death or liver transplantation) occurred in 26<br />

patients and was associated with higher IL-6 (3.56 pg/mL vs.<br />

2.13 pg/mL, P = 0.013) and IL-10 (0.54 pg/mL vs. 0.22 pg/<br />

mL, P = 0.021), but not IL-17 levels. In the hospitalized cohort,<br />

39 patients died after 90 days of follow-up. Logistic regression<br />

analysis showed that death in AD cohort was independently<br />

associated with ascites (OR 6.286, 95% CI 1.826 – 21.635;<br />

P = 0.004), MELD (OR 1.300, 95% CI 1.175 – 1.439; P<br />

< 0.001) and IL-6 (OR 1.002, 95% CI 1.000 – 1.004, P =<br />

0.029). The AUROC of IL-6 to predict 90-day mortality was<br />

0.779 ± 0.046 and the Kaplan–Meier survival probability<br />

was 90.0% for IL-6 < 21 pg/mL and 46.7% for IL-6 ≥ 21<br />

pg/mL (P < 0.001). Cytokine levels were evaluated for the<br />

prediction of bacterial infection. Regression analysis showed<br />

that bacterial infection diagnosed during the first 48 hours<br />

after admission was associated with IL-6, CRP and ascites. IL-6<br />

exhibited higher AUROC than CRP for predicting bacterial<br />

infection (0.831 ± 0.043 vs. 0.763 ± 0.048, respectively).<br />

Higher IL-6 levels were observed in ACLF patients even in the<br />

absence of bacterial infection whereas IL-10 was higher only<br />

in subjects with infection-related ACLF. Cytokines levels were<br />

reassessed at the third day of hospitalization in 74 subjects.<br />

No differences between admission and third-day levels were<br />

noted for IL-6. Lower IL-10 levels were observed at third day<br />

regardless of the presence of ACLF or death during follow-up.<br />

However, IL-17 levels dropped significantly only in those who<br />

died during follow-up. Conclusion: Circulating IL-6, IL-10 and<br />

IL-17 are of prognostic value in patients with cirrhosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Josiane Fischer, Telma E. Silva,<br />

Pedro E. Soares e Silva, Bruno S. Colombo, Letícia M. Wildner, Maria Luiza<br />

Bazzo, Tania S. Frode, Silvana Vigil de Melo, Julia S. Rosa, Esther B. Dantas-Correa,<br />

Janaína L. Narciso-Schiavon, Leonardo L. Schiavon<br />

2076<br />

Systemic inflammatory response syndrome (SIRS) in<br />

cirrhotic patients with and without bacterial infections:<br />

clinical characteristics and correlation with liver and<br />

kidney function and prognosis<br />

Caroline B. Souza, Livia B. Victor, Camila M. Alcântara, Tatiana<br />

Valdeolivas, Vanessa L. Zenatti, Daniela M. Mariz, Zulane D.<br />

Veiga, Emilia Ahmed, Flavia F. Fernandes, Gustavo Pereira; Hospital<br />

Geral de Bonsucesso, Rio de Janeiro, Brazil<br />

SIRS has been associated with morbidity and mortality in cirrhosis.<br />

Although commonly described in infections, it’s also<br />

observed in patients without it. Differences in clinical data, incidence<br />

of complications and survival between these two groups<br />

have not been studied so far. Objective: Analyze differences<br />

in clinical characteristics, diagnostic criteria, liver and kidney<br />

function, complications and survival in cirrhotics with SIRS<br />

associated or not with infections. Methods: 256 patients were<br />

evaluated. Diagnosis of SIRS was established within 48 hours<br />

of admission. Patients were classified as having sepsis (SIRS<br />

associated with infections) or SIRS without infection. Additionally<br />

we compared these groups with patients with bacterial<br />

infections without SIRS and a control group, composed of cirrhotic<br />

patients without neither SIRS nor infections. Results: Prevalence<br />

of SIRS was 35%. SIRS was diagnosed at admission<br />

in 51 and developed within 48h in 39 patients. The majority<br />

of patients fulfilled 2 diagnostic criteria (83%) of which leucocytes<br />

and temperature were the most common (69 and 62%).<br />

Sepsis was diagnosed in 55 patients and SIRS without infections<br />

in the remaining 35. Patients with sepsis more frequently<br />

had 3 or more diagnostic criteria for SIRS (22 vs. 9%, p=0.1)<br />

and higher leucocyte count (9.6 ±8.1 vs 6.2±4.6, p=0.006).<br />

Patients with sepsis had worst liver and kidney function, as<br />

evidenced by lower albumin and higher INR and creatinine<br />

values. Complications of cirrhosis were equally distributed in<br />

both groups, except for higher frequency of gastrointestinal<br />

bleeding in SIRS group (29 vs 6%, p=0.003). Acute-on-Chronic<br />

Liver Failure (ACLF) was more frequent in sepsis group (46 vs<br />

13%, p=0.001). The probability of survival at 28 and 90 days<br />

was 79 and 68%. Patients with Sepsis had lower survival than<br />

patients with SIRS both at 28 days (62 vs. 86%) and 90 days<br />

(53% vs. 74%), p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1221A<br />

2077<br />

The NDP52 Val248Ala variant increases the risk for<br />

spontaneous bacterial peritonitis in patients with alcoholic<br />

liver cirrhosis.<br />

Philipp Lutz 1,2 , Benjamin Krämer 1,2 , Dominik J. Kaczmarek 1,2 ,<br />

Marc P. Hübner 3,2 , Bettina Langhans 1,2 , Beate Appenrodt 4 , Frank<br />

Lammert 4 , Jacob Nattermann 1,2 , Achim Hoerauf 3,2 , Christian<br />

P. Strassburg 1,2 , Ulrich Spengler 1,2 , Hans Dieter Nischalke 1,2 ;<br />

1 Department of Internal Medicine I, Bonn, Germany; 2 German<br />

Center for Infection Research, Bonn, Germany; 3 Institute for Medical<br />

Microbiology, Immunology and Parasitology, University of<br />

Bonn, Bonn, Germany; 4 Department of Medicine II, Saarland University<br />

Medical Center, Homburg, Germany<br />

Aim: Spontaneous bacterial peritonitis (SBP) is frequently a<br />

fatal infection in patients with advanced liver cirrhosis. We<br />

investigated if NDP52 (nuclear dot protein 52 kDa), a negative<br />

regulator of toll-like receptor signalling and an autophagy<br />

adaptor protein, might be involved in the development of SBP.<br />

Methods: Two cohorts comprising 152 (derivation cohort) and<br />

198 patients (validation cohort) with liver cirrhosis and ascites<br />

as well as 168 healthy controls were genotyped for the NDP52<br />

Val248Ala variant and stratified for the occurrence of SBP.<br />

Results: Overall, 57 (38%) patients in the derivation cohort<br />

and 77 (39%) in the validation cohort had SBP. Cirrhosis was<br />

due to alcohol abuse in 57% of the derivation and 66% of the<br />

validation cohort. NDP52 genotype distribution was consistent<br />

with Hardy-Weinberg equilibrium. In patients with alcoholic<br />

cirrhosis, patients with SBP had an increased frequency of<br />

the NDP52 minor variant in the derivation (p=0.04) and in<br />

the validation cohort (p=0.01). In patients with liver cirrhosis<br />

of non-alcoholic etiology, the frequency of the NPD52 minor<br />

variant was comparable between patients with and without<br />

SBP (11.3% versus 12.2%). Multivariate analysis confirmed<br />

the NDP52 minor variant (odds ratio 4.7, p=0.002) and the<br />

TLR2 -16934 TT variant (odds ratio 2.5, p=0.008) as independent<br />

risk factors for SBP in alcoholic liver disease. Analysis<br />

of the interaction between the TLR2 and the NDP52 risk<br />

variant revealed that patients carrying both risk variants had<br />

a particularly high risk to acquire SBP. Conclusion: Presence<br />

of the NPD52 minor variant is a risk factor for SBP in patients<br />

with alcoholic cirrhosis, indicating that autophagy might be<br />

involved in the pathogenesis of SBP.<br />

Disclosures:<br />

Christian P. Strassburg - Advisory Committees or Review Panels: Novartis, Roche;<br />

Speaking and Teaching: Novartis, Merz, MSD, Falk Pharma, BMS, Abbvie<br />

The following authors have nothing to disclose: Philipp Lutz, Benjamin Krämer,<br />

Dominik J. Kaczmarek, Marc P. Hübner, Bettina Langhans, Beate Appenrodt,<br />

Frank Lammert, Jacob Nattermann, Achim Hoerauf, Ulrich Spengler, Hans Dieter<br />

Nischalke<br />

2078<br />

Assessment and relevance of coagulation disorders in<br />

patients with acute-on-chronic liver failure during Systemic<br />

Inflammatory Response (SIRS) and development of<br />

sepsis<br />

Madhumita Premkumar, Priyanka Saxena, Roshni Mirza, Sukriti<br />

Sukriti, Chhagan Bihari, Ashok Choudhary, Chandan K. Kedarisetty,<br />

Shiv K. Sarin; Hepatology, Institute of Liver and Biliary Sciences,<br />

Delhi, India<br />

Background: Coagulation system is rebalanced in cirrhosis.The<br />

status of baseline coagulation disturbances in ACLF and their<br />

alterations with development of sepsis are largely unknown.<br />

Aim: To study the dynamic changes in coagulation profile in<br />

ACLF, and their correlation with evolving SIRS and sepsis.<br />

Patients and Methods : Of the 243 consecutive patients with<br />

ACLF (APASL criteria), 114 with no evidence of sepsis were<br />

recruited (mean age 44.3±11.7 yr, males 90%). Predominant<br />

etiology for underlying liver disease was ethanol (63.1%).<br />

Patients were evaluated by tests like INR and thromboelastography<br />

(TEG), Sonoclot, and coagulation factor assays.<br />

These results were compared with 25 controls. At presentation<br />

40/114 (35.1%) had SIRS. Twenty eight (24.5%) developed<br />

sepsis by day 3 and 52/112 (56.1%) by day 7. Only 11 /74<br />

(14.8%) developed sepsis de novo without prior SIRS. SIRS at<br />

presentation posed a mortality risk of 48%. A deranged TEG<br />

at presentation was a predictor of bleeding (OR 2.1, p=0.05)<br />

and mortality (OR 3.1, p=0.05).Platelet count and functions (as<br />

predicted by sonoclot) were significantly lower in ACLF patients<br />

with sepsis at day 3 and 7. INR at baseline was associated<br />

with a deranged R and K component of TEG (p=0.05) and<br />

ACT of Sonoclot (p=0.02). Baseline ACT, CR and PF (sonoclot<br />

parameters), were associated with development of SIRS at day<br />

3 (p=0.013) and sepsis at day 7( p=0.012). Likewise R and<br />

K of TEG (p=0.064) predicted sepsis. Coagulation assays for<br />

protein C,and antithrombin III were lower in all ACLF patients,<br />

when compared with controls, but did not change with sepsis.<br />

Factor 8 levels were significantly increased in all ACLF<br />

subgroups. Conclusions: Coagulation abnormalities in ACLF<br />

are associated and correlate with an increased tendency to<br />

bleed, risk of sepsis and mortality. Coagulation parameters<br />

such as INR, TEG and Sonoclot are affected by presence of<br />

SIRS. Sonoclot parameters at baseline can serve as predictors<br />

of development of sepsis and poor outcome<br />

Table 1: Coagulation parameters of ACLF when compared with<br />

those with sepsis at day 3 and day 7.<br />

ACT :Activated coagulation time<br />

Disclosures:<br />

The following authors have nothing to disclose: Madhumita Premkumar, Priyanka<br />

Saxena, Roshni Mirza, Sukriti Sukriti, Chhagan Bihari, Ashok Choudhary, Chandan<br />

K. Kedarisetty, Shiv K. Sarin<br />

2079<br />

Characteristics and outcome of Severe Alcoholic Hepatitis-related<br />

Acute-on-Chronic Liver Failure Syndrome<br />

Alexia Cornillie, Eric Trépo, Delphine Degré, Jonas Schreiber,<br />

Antonia Lepida, Christophe Moreno, Thierry Gustot; Gastroenterology<br />

and Hepato-Pancreatology, Erasme Hospital, Université<br />

Libre de Bruxelles, Brussels, Belgium<br />

Background & Aims: Although active alcoholism is known to<br />

be a potential trigger for Acute-on-Chronic Liver Failure (ACLF),<br />

place of severe alcoholic hepatitis (AH) as precipitating event is<br />

currently unknown. We assessed characteristics, outcome and<br />

predictive factors for the development of ACLF in patients with<br />

severe biopsy-proven alcoholic hepatitis (AH). Methods: We<br />

prospectively followed 145 consecutive biopsy-proven severe<br />

AH (mDF≥32) episodes for 6 months. We retrospectively<br />

reviewed ACLF diagnosis, grades (based on CANONIC criteria)<br />

and prognostic scores, CLIF-C Organ Failure score (OFs)<br />

and CLIF-C ACLFs (Jalan et al. J Hepatol 2014;61:1038-47) at<br />

the time of liver biopsy (baseline) and at different time points


1222A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

during 3-month follow-up. Results: At baseline, ACLF was diagnosed<br />

among 52% of severe AH episodes (17% of ACLF grade<br />

1 (ACLF-1), 16% of ACLF-2 and 19% of ACLF-3). Compared<br />

to AH episodes without ACLF at baseline, AH episodes with<br />

ACLF had a higher 28-day and 90-day transplant-free mortality<br />

rate (54 vs. 10% and 68% vs. 18% respectively, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1223A<br />

ease diagnosis (p=0.00001, Odds Ratio 0.22 and p=0.008,<br />

OR 2.0 respectively) Conclusion: We found that sarcopenia<br />

has a higher prevalence among cirrhotic patients undergoing<br />

liver transplantation at our Institution than reported in the<br />

literature. Interestingly, obesity was found to be a significant<br />

predictor of pre-transplant sarcopenia in this cohort<br />

Disclosures:<br />

The following authors have nothing to disclose: Trushar Patel, Valery Vilchez,<br />

Rashmi T. Nair, Jennifer Watkins, Anna Christina Dela Cruz, Roberto Gedaly,<br />

Terrence Barrett<br />

2082<br />

Liver Cirrhosis Is Strongly And Independently Associated<br />

With Mortality In Patients With Bacterial Endocarditis:<br />

Results Of A Case Control Multicenter Study Of 202<br />

Cases<br />

Jean françois D. Cadranel 1 , Isabelle Ollivier-Hourmand 2 , Salah<br />

Zerkly 3 , Christophe Bureau 4 , Thierry Thevenot 5 , Patrice Cacoub 6 ,<br />

Armand Garioud 1 , Gilles Macaigne 7 , Laurent Alric 8 , Vincent Jouannaud<br />

4 , Hortensia Lison 1 , Carine Chagneau-Derrode 9 , Alexandre<br />

Pariente 10 , Agnes Pelaquier 11 , Marc Bourlière 13 , Xavier Causse 12 ,<br />

Manon Allaire 2 , Jean-Baptiste Nousbaum 14 , Jerôme Dumortier 15 ,<br />

Alexandre Louvet 16 , Isabelle Rosa 17 , Nathalie Ganne-Carrié 18 ,<br />

Jerôme Gournay 19 , Hélène Blasco-Perrin 4 , Teresa Antonini 20 , Laurent<br />

Spahr 21 , Jean-Pierre Bronowicki 22 , Christine Silvain 9 , Xavier<br />

Amiot 23 , Vincent Di Martino 5 , Bruno Lesgourgues 24 , Jean-Didier<br />

Grangé 23 , Jean-Marie Péron 4 , Didier Samuel 25 , Xavier Adhoute 13 ,<br />

Hervé Hagège 17 , Philippe Mathurin 16 , Jacques Denis 26 , Pierre<br />

Iaria 27 , Thong Dao 2 ; 1 Hepatogastoenterology department,<br />

GHPSO Creil, Creil, France; 2 Hepatogastroenterology Unit, CHU,<br />

Caen, France; 3 Epidemiology Unit, CHU, Amiens, France; 4 Hepatogastroenterology<br />

Unit, CHU, Toulouse, France; 5 Hepatogastroenterology<br />

Unit, CHU, Besançon, France; 6 Internal Medicine Unit,<br />

CHU Pitié-Salpêtrière, Paris, France; 7 Hepatogastroenterology<br />

Unit, CH, Lagny-sur-Marne, France; 8 Internal medicine Unit, CHU,<br />

Toulouse, France; 9 Hepatogastroenterology Unit, CHU, Poitiers,<br />

France; 10 Hepatogastroenterology Unit, CH, Pau, France; 11 Hepatogastroenterology<br />

Unit, CH, Montélimar, France; 12 Hepatogastroenterology<br />

Unit, CHR, Orléans, France; 13 Hepatogastroenterology<br />

Unit, CH Saint-Joseph, Marseille, France; 14 Hepatogastroenterology<br />

Unit, CHU, Brest, France; 15 Hepatogastroenterology Unit,<br />

CHU, Lyon, France; 16 Hepatogastroenterology Unit, CHU, Lille,<br />

France; 17 Hepatogastroenterology Unit, CHIC, Créteil, France;<br />

18 Hepatogastroenterology Unit, CHU Jean Verdier, Bondy, France;<br />

19 Hepatogastroenterology Unit, CHU, Nantes, France; 20 Centre<br />

Hépato-Biliaire, Villejuif, France; 21 Hepatogastroenterology Unit,<br />

Geneva Hospitals, Geneva, Switzerland; 22 Hepatogastroenterology<br />

Unit, CHU, Nancy, France; 23 Hepatogastroenterology Unit,<br />

CHU Tenon, Paris, France; 24 Hepatogastroenterology Unit, CH,<br />

Montfermeil, France; 25 Hepatogastroenterology Unit, Centre Hépato-Biliaire,<br />

Villeuif, France; 26 Hepatogastroenterology Unit, CH Sud<br />

Francilien, Corbeil-Essonnes, France; 27 Cardiology Unit, GHPSO,<br />

Creil, France<br />

Background and Aims: Bacterial endocarditis (BE) is severe in<br />

patients (pts) with liver cirrhosis (LC); data rely on uncontrolled<br />

<strong>studies</strong>. Aims of this study were to compare clinical presentation<br />

and mortality of BE in LC pts with those of controls (CT) without<br />

cirrhosis matched for sex, age and diabetes. Methods: Medical<br />

charts for all LC pts with BE seen in 23 liver units (2000-2013)<br />

were reviewed and compared with those of CT. Short-term mortality<br />

was analyzed by univariate and logistic regression analysis.<br />

Results: 101 BE in LC pts and 101 BE in CT were analyzed.<br />

LC pts: 63.2 [42-87] years, CT: 65 [46-93]; 145 M (72.1%).<br />

LC causes: alcohol: 78 (67.2%), viral: 17 (14.6%), NAFLD 14<br />

(12%). Child-Pugh scores: A: 8pts (8.8%), B: 39 pts(42.9%), C:<br />

44 pts(48.4%). In LC pts: total bilirubin was 67.8μmol/l±77.7,<br />

prothrombin time (PT) 52.7%±18.1, serum albumin 25.2g/<br />

l±5.9, serum creatinine (SC) 123.5μmol/l±103.6. Blood cultures<br />

were positive in 181 pts (91.4%). 39 LC pts (40.2%)<br />

and 52 CT (53.6%) had a pre-existing heart disease (NS).<br />

At diagnosis, 43 LC (47.7%) and 31 CT (34.0%) exhibited<br />

heart failure (NS). Severe sepsis or septic shock was noted<br />

in 33 LC pts (33.3%) and 23 CT (23.0%) (NS). The source<br />

of contamination was digestive, urinary, or cutaneous (DUC)<br />

in 41 LC pts and in 21 CT (OR 2.4 [1.1–5.1], p


1224A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

The following authors have nothing to disclose: Jean françois D. Cadranel, Salah<br />

Zerkly, Thierry Thevenot, Armand Garioud, Gilles Macaigne, Vincent Jouannaud,<br />

Hortensia Lison, Carine Chagneau-Derrode, Agnes Pelaquier, Manon Allaire,<br />

Jean-Baptiste Nousbaum, Jerôme Dumortier, Alexandre Louvet, Isabelle Rosa,<br />

Jerôme Gournay, Hélène Blasco-Perrin, Teresa Antonini, Laurent Spahr, Christine<br />

Silvain, Bruno Lesgourgues, Jean-Didier Grangé, Jean-Marie Péron, Xavier<br />

Adhoute, Hervé Hagège, Pierre Iaria<br />

2083<br />

Hispanic ethnicity is associated with severe sepsis<br />

among infected patients with end-stage liver disease<br />

Vinay Sundaram 1 , Jason J. Xu 1 , Christie Jeon 1 , David S. Goldberg<br />

2 ; 1 Cedars-Sinai Medical Center, Los Angeles, CA; 2 Medicine,<br />

University of Pennsylvania, Philadelphia, PA<br />

Background & Aims: The risk and prognosis for severe sepsis<br />

varies by race and ethnicity in the general population. We<br />

aimed to determine whether these disparities apply in patients<br />

with end-stage liver disease. Methods: We studied the Nationwide<br />

Inpatient Sample, 2009-2011. The cohort included<br />

patients with decompensated cirrhosis using validated ICD-9-<br />

code based algorithms, with at least one or more of the following<br />

infections: bacteremia, skin/soft tissue infection, urinary<br />

tract infection, lower respiratory infection, Clostridium difficile<br />

infection, and spontaneous bacterial peritonitis. We adjusted<br />

for patient co-morbidities using the Deyo modification of the<br />

Charlson index. We determined predictors of severe sepsis<br />

using logistic regression models. Results: Of 130,088 hospitalized<br />

patients with decompensated cirrhosis, 44,680 (34.4%)<br />

had ≥1 infection. Among infected patients, 8,077 (18.1%)<br />

had severe sepsis. Infected patients comprised 27,137 whites<br />

(71.1%), 3,704 black (9.7%) and 7,306 Hispanics (19.1%).<br />

White patients were older (mean 59.7y, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1225A<br />

2085<br />

Neutrophil to Lymphocyte Ratio (NLR) Predicts Liver<br />

Related Death in Low MELD Cirrhotic Patients Awaiting<br />

Liver Transplant<br />

Avash Kalra 1 , Joel P. Wedd 2 , Kiran Bambha 1 , Jane Gralla 1 , Hugo<br />

R. Rosen 1 , Scott W. Biggins 1 ; 1 University of Colorado Denver,<br />

Aurora, CO; 2 Division of Digestive Diseases, Emory, Atlanta, GA<br />

MELD score has reduced accuracy in cirrhotic patients with<br />

MELD ≤20. The D’Amico 5 Stages of Cirrhosis model is a<br />

cumulative clinical staging tool and aids in identifying low<br />

MELD patients at higher risk for liver related death (Liver Transpl<br />

2014; 20:1193-1201). High (≥4, J Hepatol 2013; 58:58-64)<br />

Neutrophil to Lymphocyte Ratio (NLR) is associated with inflammatory<br />

states and may predict cirrhotic decompensation and<br />

low MELD death in cirrhosis. AIM: To evaluate the prognostic<br />

utility of high NLR (≥4) for liver related death among low MELD<br />

patients listed for liver transplant (LT), controlling for Cirrhosis<br />

Stage. METHODS: A nested case-control study was performed<br />

using adults (>17 yo) with cirrhosis awaiting LT from 2/2002-<br />

5/2011; Cases= liver related death and MELD≤20 within 90<br />

days of death; and Controls= alive for ≥90 days after LT listing.<br />

Controls were randomly chosen from all listed patients<br />

and matched (up to 3:1) to Cases by listing year, lab MELD,<br />

age, gender, and disease etiology. NLR, Cirrhosis Stage, Na,<br />

albumin, and hepatic encephalopathy (HE) were assessed at<br />

date of lowest MELD within 90 days of death for Cases and<br />

within 90 days after listing for Controls. Cirrhosis Stages are<br />

1=no varices or ascites, 2=varices, 3=variceal bleed, 4=ascites,<br />

5=ascites and variceal bleed. Conditional logistic regression<br />

was used to evaluate risk of liver related death. RESULTS:<br />

There were 41 Cases and 66 Controls; MELD scores were similar.<br />

Clinical states contributing to death in Cases were: sepsis<br />

49%, spontaneous bacterial peritonitis 15%, variceal bleeding<br />

24%, and hepatorenal syndrome 22%. NLR 25 th , 50 th , 75 th<br />

percentile cut offs were 1.9, 3.1, and 6.8. NLR was ≥4 in<br />

4/41 (59%) Cases and 14/66 (21%) Controls. Median (IQR)<br />

NLR for Cases and Controls was 5.0 (3.0-10.3) and 2.35<br />

(1.6-3.9). In univariate analysis, NLR (continuous, ≥1.9, ≥4,<br />

≥6.8), increasing Cirrhosis Stage (**categorical,1-5), variceal<br />

bleeding, large ascites, HE, Na, Albumin were significantly<br />

(p 30 ng/ml) had significantly (P = 0.009) higher daily<br />

exposure to all thresholds of lux intensities, as compared to<br />

patients with vitamin D deficiency. A seasonal comparison confirmed<br />

significantly higher lux accrued in spring and summer (P<br />

< 0.0001). Median dietary vitamin D intake (1.48 mcg/day,<br />

0.26 - 12.12) was lower than typical intakes in the general<br />

population (2 - 4 mcg). We observed no correlations between<br />

light exposure and BMI or physical activity, neither between<br />

body fat mass and serum vitamin D levels. Light intensity was<br />

the only independent predictor of serum vitamin D level in multivariate<br />

regression analysis (P = 0.004). Serum vitamin D levels<br />

were highest in patients with > 30 min exposure at > 1000<br />

lux/day (P = 0.002). Conclusions: In CLD patients, serum vitamin<br />

D levels positively correlate with light exposure but not with<br />

dietary vitamin D intake. The contribution of vitamin D from diet<br />

is minimal and below the recently revised recommendations for<br />

vitamin D intake. In contrast, our findings highlight the need<br />

for combined interventions to treat vitamin D deficiency and<br />

may guide specific recommendations for sunlight exposure in<br />

patients with CLD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elisabeth Nick, Ralf Kaiser, Frank<br />

Lammert, Caroline S. Stokes


1226A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2087<br />

Polymorphisms in the interleukin (IL)-1 gene cluster are<br />

associated with multiple markers of systemic inflammation<br />

in patients with Acute-on-Chronic Liver Failure<br />

(ACLF)<br />

Jose Alcaraz-Quiles 1 , Esther Titos 4 , Cristina Lopez-Vicario 1 ,<br />

Bibiana Rius 1 , Aritz Lopategi 1 , Mireia Casulleras 1 , Veronica<br />

Garcia-Alonso 1 , Andrea De Gottardi 5 , Mauro Bernardi 3 , Ivo Graziadei<br />

6 , Henning Gronbaek 7 , Pere Gines 8,4 , Vicente Arroyo 8,2 ,<br />

Joan Clària 1,2 ; 1 Department of Biochemistry and Molecular Genetics,<br />

Hospital Clínic-University of Barcelona, Bareclona, Spain;<br />

2 University of Barcelona, Barcelona, Spain; 3 EASL-CLIF Consortium,<br />

Barcelona, Spain; 4 CIBERehd, Barcelona, Spain; 5 University<br />

of Berne, Berne, Switzerland; 6 Innsbruck Medical University,<br />

Innsbruck, Austria; 7 Aarhus University Hospital, Aarhus, Denmark;<br />

8 Liver Unit, Hospital Clinic, Barcelona, Spain<br />

Background and aim: Acute-on-Chronic liver failure (ACLF) is<br />

an increasingly recognized entity encompassing multiorgan<br />

failure in patients with cirrhosis. Recent findings suggest that<br />

immune dysregulation predisposes decompensated cirrhotic<br />

patients to a cycle of adverse events culminating in uncontrolled<br />

systemic inflammation and ACLF progression. Although<br />

the exact mechanisms leading to inflammation-driven ACLF<br />

remain to be elucidated, systemic inflammation is influenced by<br />

multiple genes regulating the innate immune response. In this<br />

study, we screened a set of six inflammation-related genes in<br />

a population-based study of patients with decompensated liver<br />

cirrhosis. Patients and Methods: Two hundred seventy-nine cirrhotic<br />

patients (178 with ACLF and 101 without ACLF) from the<br />

CANONIC study of the CLIF consortium were genotyped for<br />

SNP polymorphisms in genes coding for IL-1beta (rs1143623,<br />

G/C), IL-1 receptor antagonist (IL-1Ra) (rs4251961, T/C),<br />

SOCS3 (rs4969170, G/A), NOD2 (rs3135500, G/A),<br />

CMKLR1 (rs1878022, T/C) and IL-10 (rs1800871, G/A) by<br />

allellic discrimination TaqMan assays. Serum cytokine levels<br />

were measured by fluorescent bead-based (Luminex) immunoassays.<br />

Results: Among the six SNPs analyzed, we identified<br />

two polymorphisms belonging to the IL-1 gene cluster (IL-1beta,<br />

rs1143623 and IL-1Ra, rs425196) in strong association with<br />

ACLF. Homozygous C carriers in the rs1143623 SNP showed<br />

lower serum levels of IL-1β in parallel with reduced markers<br />

of systemic inflammation (i.e. IL-1alpha, IL-6 and TNF-alpha).<br />

Also, heterozygous TC carriers of the rs425196 SNP had<br />

lower serum levels of IL-1beta, IL-1alpha, IL-6, TNF-alpha and<br />

IL-8. Under different inheritance models, both genotypes were<br />

protective factors for presence of ACLF (rs1143623; OR: 0.34,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1227A<br />

2089<br />

Adding C-reactive Protein and Procalcitonin to the MELD<br />

Score Improves Mortality Prediction in Patients Admitted<br />

with Complications of Cirrhosis<br />

Sakkarin Chirapongsathorn 1,2 , Worawan Bunraksa 2 , Dollapas<br />

Punpanich 3 , Amnart Chaiprasert 4 , Patrick S. Kamath 1 ; 1 Gastroenterology<br />

and Hepatology, Mayo Clinic, Rochester, Rochester,<br />

MN; 2 Gastroenterology, Phramongkutklao Hospital and College<br />

of Medicine, Royal Thai Army, Bangkok, Thailand; 3 Biomedical<br />

Research and Development Center, Phramongkutklao Hospital and<br />

College of Medicine, Royal Thai Army, Bangkok, Thailand; 4 Nephology,<br />

Phramongkutklao Hospital and College of Medicine, Royal<br />

Thai Army, Bangkok, Thailand<br />

Background: MELD score has been shown to be the best predictor<br />

of mortality prediction in cirrhosis. Serum C-reactive protein<br />

(CRP) and procalcitonin (PCT) predict the risk of death in critical<br />

illness patients and also help to predict short-term mortality<br />

in cirrhosis. A model adding CRP and procalcitonin to the<br />

MELD score may increase accuracy for mortality prediction in<br />

patients admitted with complications of cirrhosis. Methods: A<br />

prospective cohort study was carried out in consecutive cirrhotic<br />

patients admitted with complications of cirrhosis during September<br />

2012 to December 2013 at Phramongkutklao Hospital,<br />

Thailand. All patients had venous CRP, PCT and laboratory<br />

values for MELD score calculation measured within 48 hours<br />

of admission. Cox regression analysis and ROC curve were<br />

used to predict mortality. The MELD-CRP score was externally<br />

validated in 818 eligible patients from Mayo Clinic, Rochester.<br />

Results: A cohort of 223 patients with cirrhosis was admitted<br />

during the study period. Seventy-one patients were eligible for<br />

analysis. MELD score was predictive of 90-day mortality (OR<br />

1.19 (95%CI; 1.09-1.32); AUROC of 0.81). Adding CRP or/<br />

and PCT to the MELD score improved the predictive of 90-day<br />

mortality: MELD-CRP OR 2.71 (95% CI; 1.66-4.99); MELD-PCT<br />

OR 2.72 (95% CI; 1.66-4.99); MELD-CRP-PCT OR 2.71 (95%<br />

CI; 1.67-4.92). AUROC for MELD, MELD-CRP, MELD-PCT, and<br />

MELD-CRP-PCT were 0.81, 0.83, 0.84, and 0.85 respectively.<br />

Adding CRP or/and PCT to the MELD score also improved<br />

30-day mortality prediction. Similar results of MELD-CRP score<br />

were obtained from the Mayo Clinic external validation cohort.<br />

MELD-CRP score improved the predictive of 30-day mortality<br />

(OR 1.49 (95% CI; 1.28-1.73); AUROC 0.71) and 90-day<br />

mortality (OR 1.43 (95% CI; 1.27-1.62); AUROC 0.69) compared<br />

to MELD score. Conclusion: The MELD-CRP, MELD-PCT<br />

and MELD-CRP-PCT scores are superior to the MELD score in<br />

predicting mortality in cirrhotic patients admitted with complications.<br />

The following authors have nothing to disclose: Sakkarin Chirapongsathorn,<br />

Worawan Bunraksa, Dollapas Punpanich, Amnart Chaiprasert<br />

2090<br />

Presepsin as a new biomarker for old expectations in<br />

the diagnosis and prognosis of bacterial infection in<br />

cirrhosis<br />

Maria Papp 1 , Tamas Tornai 1 , David Tornai 2 , Zsuzsanna Vitalis 1 ,<br />

Istvan Tornai 1 , Peter Antal-Szalmas 2 ; 1 Department of Gastroenterology,<br />

University of Debrecen, Faculty of Medicine, Debrecen,<br />

Hungary; 2 Department of Laboratory Medicine, University of<br />

Debrecen, Faculty of Medicine, Debrecen, Hungary<br />

Background&Aims: Bacterial infections are frequent complications<br />

in cirrhosis with significant mortality. Early diagnosis is<br />

essential but still a diagnostic challenge from both the clinical<br />

and the laboratory part. The aim of this study is to evaluate<br />

and compare the diagnostic and prognostic value of presepsin<br />

plasma levels with CRP and PCT in bacterial infections of<br />

patients with cirrhosis. Methods: A total of 216 patients with<br />

cirrhosis (54.4% males, age: 57.6±10.3 years and median<br />

MELD score: 13 [95% CI: 10-17]) were consecutively enrolled.<br />

At admission enrollment presence of bacterial infection were<br />

assessed on the basis of conventional criteria, liver-oriented<br />

scores were calculated and plasma presepsin, CRP and PCT<br />

levels were measured. A short-term follow-up study was conducted<br />

to assess the development of organ system failure(s) and<br />

28-day mortality associated to bacterial infections. Results: Bacterial<br />

infection was found in 75 (34.7%) patients. Plasma presepsin<br />

levels were significantly higher in patients with infection as<br />

compared to those without (1002 pg/mL [575-2149] vs. 477<br />

[332-680] pg/mL, p0.5 pg/ml (OR:<br />

9.10, p=0.006) or CRP >40 mg/l (OR: 4.03, p=0.039) but<br />

not presepsin level were independet risk factor for 28-day mortality.<br />

Conclusions: Presepsin is a valuable new biomarker for<br />

defining severity of infections in cirrhosis proving same efficacy<br />

as PCT. However, for the prediction of short-term mortality,<br />

liver-oriented scores and admission level of conventional APP<br />

proteins, particularly PCT are the appropriate tools.<br />

Disclosures:<br />

Istvan Tornai - Advisory Committees or Review Panels: Roche, AbbVie, BMS,<br />

Janssen-Cilag<br />

The following authors have nothing to disclose: Maria Papp, Tamas Tornai,<br />

David Tornai, Zsuzsanna Vitalis, Peter Antal-Szalmas<br />

Disclosures:<br />

Patrick S. Kamath - Advisory Committees or Review Panels: Sequana Medical


1228A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2091<br />

The relationship of fecal bacterial isolates to systemic<br />

infections in cirrhotic patients<br />

Mary K. Rude 1 , Jimmy Ma 1 , Suki Chandrasekaran 2 , Phillip Tarr 2 ,<br />

Jaquelyn Fleckenstein 1 ; 1 Gastroenterology, Washington University,<br />

Saint Louis, MO; 2 Pediatrics, Washington University, St. Louis, MO<br />

Background: Infections in patients with cirrhosis increase mortality<br />

4-fold, with the most frequent infections being spontaneous<br />

bacterial peritonitis and urinary infections. Bacterial<br />

translocation from the gut in cirrhosis is postulated to play an<br />

etiologic role in developing infection. However, no strategies<br />

exist to proactively risk-stratify which patients with cirrhosis will<br />

develop systemic infection. This study sought to determine if the<br />

pathogens can be isolated from stool prior to the onset or during<br />

systemic infections. Methods: Twenty-six adults with cirrhosis,<br />

hospitalized for a variety of indications, were prospectively<br />

enrolled and stool samples were obtained approximately every<br />

48 hours. Patients received routine medical care, and if there<br />

was clinical evidence of an infection, appropriate cultures were<br />

obtained. Fecal samples collected prior to the infectious event<br />

were used to recover the pathogens that subsequently invaded<br />

their bloodstreams and/or urine. We then sought to recover the<br />

same organisms in stool samples in enrolled patients who did<br />

not have evidence of a systemic infection. Culture-based techniques<br />

were used to recover stool bacteria that provisionally<br />

matched the bloodstream and/or urine organisms. MALDI-TOF<br />

was then used to confirm identity between fecal and systemic<br />

isolates. Results: Of the 26 subjects enrolled, 10 patients developed<br />

extraintestinal infections. Of these 10 patients, 3 were<br />

infected with Klebsiella pneumoniae, and 2 were infected with<br />

Staphylococcus aureus. K. pneumoniae was cultured from the<br />

stool and confirmed with MALDI-TOF in all 3 patients (6 stool<br />

samples) who subsequently developed K. pneumoniae infection.<br />

K. pneumoniae was recovered from only 7 of the 23<br />

patients (9 of 50 stool samples) who did not develop infection<br />

with K. pneumoniae (Fisher’s exact test, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1229A<br />

of NC, and 2.4% of CHF patients. Sepsis rose from 2.2% to<br />

5.7% (159% increase) in cirrhotic patients and by 2010 was<br />

greater than NC (5.3%) or CHF (3.1%) patients. 35.8% of<br />

cirrhotic patients with sepsis died compared to 17.9% of NC<br />

and 16.4 % of CHF patients (p


1230A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

cirrhosis (M/F: 12/5, age: median 59.2 years (49-65), etiology<br />

of liver disease: HBV±HDV: 23.5%, HCV: 17.7%, alcohol:<br />

41.2%, other: 17.6%, Child-Pugh score: median 7.5 (5-12),<br />

MELD: median 10.5 (7-21), ascites (%): 3 (17.7) attending<br />

the outpatient clinics were included. None had hepatocellular<br />

carcinoma. Indications for PPIs administration were esophagitis<br />

(35%), peptic ulcer (59%) and others (6%). Bacterial DNA<br />

in serum and the levels of endotoxin, LBP, TGF-β, IL-1β, IL-6,<br />

IL-8, IL-12, IL-10, TNF-α and NOx were assessed before (time<br />

1) and at the end of treatment with PPIs (time 2). Bacterial<br />

DNA was amplified by PCR using universal 16SrRNA primers,<br />

endotoxin, LBP and TGF-β were measured by ELISA, the other<br />

cytokines by a Cytometric Bead Array method, and NO levels<br />

determined with a nitric oxide quantitation kit. Wilcoxon<br />

Signed Rank Test was used to evaluate significant differences<br />

in the parameters assayed before and after PPIs administration.<br />

Results: None of the patients developed an infection during the<br />

study period. Spearman’s correlation analysis demonstrated<br />

a correlation between baseline LBP and MELD score (r=0.63,<br />

p=0.02). Bacterial DNA was not detected before or after the<br />

treatment. Time 1: Median endotoxin: 3.19 (1.03-4.33), LBP:<br />

4.08 (3.23-12.99) μg/ml, NOx: 9.04 (2.44-15.72) μΜ,<br />

TGF-β: 4.85 (2.43-16.64) ng/ml, TNF-α: 3.05 (0-8.26), IL-1:<br />

4.7 (0-11.95), IL-6: 4.37 (0-16.09), IL-12: 7.69 (0-25.88),<br />

IL-10: 6.56 (0-27.87), IL-8: 103.4 (0-281.4) pg/ml. Time 2:<br />

Median endotoxin: 2.76 (1.49-86.55), LBP: 4.65 (0-17.53),<br />

NOx: 7.36 (0.56-23.08), TGF-β: 3.49 (1.54-15.94), TNFα:<br />

5.4 (0-60.98), IL-1: 8.57 (0-184.8), IL-6: 10.29 (0-31.5),<br />

IL-12: 15.96 (0-159), IL-10: 5.16 (0-34.8), IL-8: 64.45 (18.99-<br />

2892.4). No significant differences were observed between<br />

the concentrations of all measured indices between times 1 and<br />

2 (p>0.05). Subgroup analysis according to Child-Pugh stage<br />

yielded similar results. Conclusions: Short-term PPIs administration<br />

had no effect on serum bacterial DNA, bacterial products<br />

or cytokine concentrations in patients with liver cirrhosis.<br />

Disclosures:<br />

Christos K. Triantos - Speaking and Teaching: Bristol, Gilead<br />

The following authors have nothing to disclose: Maria Kalafateli, Panagiota Spadidea,<br />

Dimitrios Goukos, Efstratios Koutroumpakis, Stelios F. Assimakopoulos,<br />

Konstatinos Thomopoulos, Charalambos Gogos, Chrisoula Labropoulou-Karatza<br />

C, Athanasia Mouzaki, George Daikos, Vasiliki Nikolopoulou<br />

2096<br />

A prognostic model evaluating neutrophil-to-leucocyte<br />

ratio, organ failure and MELD scores as predictors of<br />

long-term survival in acute-on-chronic liver failure<br />

Danai Agiasotelli, Alexandra Alexopoulou, Larisa E. Vasilieva,<br />

Georgia Kalpakou, Spyros P. Dourakis; 2nd Dept of Medicine,<br />

Medical School University of Athens, Athens, Greece<br />

Purpose/Background: Acute-on chronic liver failure (ACLF) is<br />

defined as an acute deterioration of liver disease with high<br />

mortality in patients with cirrhosis. The time needed to reverse<br />

this condition and the factors affecting mortality after the early<br />

28-day-period have been evaluated. Methods: One-hundredninety-seven<br />

consecutive patients with cirrhosis were included.<br />

Patients were prospectively followed-up for 180 days. Results:<br />

ACLF was diagnosed in the 54.8% of patients. Infection was<br />

the most common precipitating event in patients with ACLF.<br />

Patients with ACLF had the worst prognosis compared to those<br />

without (log-rank P30 days & ≤75 days) and<br />

to 1.26 [(95% CI 0.53-3.00), P=0.609] during the third period<br />

of observation (>75 and


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1231A<br />

2098<br />

A novel mouse model for the study of pediatric Non-Alcoholic<br />

Fatty Liver Disease (NAFLD)<br />

Veronica Marin 1 , Natalia Rosso 1 , Matteo Dal Ben 1 , Alan Raseni 2 ,<br />

Cristina Degrassi 3 , Claudio Tiribelli 1,4 , Silvia Gazzin 1 ; 1 Fondazione<br />

Italiana Fegato, Trieste, Italy; 2 S.C.Laboratorio Analisi<br />

Cliniche, IRCCS Burlo Garofalo, Trieste, Italy; 3 Medical Research<br />

Institute, Trieste, Italy; 4 Department of Medical Science, University<br />

of Trieste, Trieste, Italy<br />

The increasing prevalence of pediatric NAFLD is considered<br />

a booming problem with worrying future outcome. Consistent<br />

pediatric models to mimic the clinical condition, as well as<br />

<strong>studies</strong> considering eventual gender differences, are still lacking.<br />

The aim of this study is to develop and characterize a<br />

Juvenile NAFLD model. Males (M) and females (F) C57BL/6<br />

mice, immediately after weaning were randomly assigned to<br />

control (CTRL) or high-fat high-carbohydrate diet (HFHCD).<br />

Animals had ad-libitum food for 16wks access. Body-weight,<br />

glycaemia, insulinemia, triglycerides, total cholesterol, HDL-C,<br />

LDL-C, ALT and liver histology were screened every 4wks and<br />

compared to CTRL. Soon after the 1 st week, HFHCD induced<br />

a significant bodyweight gain in both genders. Males, after<br />

4wks presented also hyperplasia of epididymal fat-pads and<br />

after week 12 th , a significant hepatomegaly. Males showed<br />

earlier alteration of glycemia, insulinemia, lipid profile and<br />

ALT (Table1). Interestingly, comparable body/blood alterations<br />

were observed in females only at the 16 th week. Liver histology<br />

showed in both genders a mixed macro-microvesicular steatosis<br />

increasing steadily after the 8 th week. Inflammatory cells foci<br />

were observed in males from the beginning and increased over<br />

the time. On the contrary, inflammation was absent in females.<br />

Surprisingly, both genders developed progressive fibrosis starting<br />

from the 8 th week and rising steadily over the time. This<br />

juvenile NAFLD model progresses faster than those previously<br />

reported in adults. A clear gender difference was found in the<br />

onset of the liver injury. Even if the final outcome was comparable<br />

between genders (fibrosis grade 2), males presented early<br />

signs of liver injury, whereas females did not.<br />

was January 2004 to March 2014. The total number of abnormal<br />

results for both caeruloplasmin and A1AT, defined as a<br />

value below the lower limit of normal, were assessed and subsequent<br />

diagnosis evaluated from medical records. The value<br />

for caeruloplasmin varied during the study, however the majority<br />

of samples were < 0.17 g/L. For A1AT deficiency the limit<br />

was < 0.9 g/L. Results During the study period caeruloplasmin<br />

was ordered on 2444 patients. The mean age of patients was<br />

46.4 years, 56.4% were male and 76 tests (3.1%) were abnormal.<br />

Plasma copper was performed on 41 patients (53.9%),<br />

however 36 (87.8%) were simultaneous with caeruloplasmin.<br />

Only 13 patients (17.1%) underwent 24-hour urinary copper<br />

measurement with 9 (69.2%) elevated. 21 patients (27.6%)<br />

had a liver biopsy, and 3 (14.3%) had hepatic copper quantification.<br />

Only 1 patient had WD confirmed. Other diagnoses<br />

included viral hepatitis (32.3%), alcohol (26.3%), drug induced<br />

liver injury (10.5%) and NAFLD (7.9%). The positive predictive<br />

value of a low caeruloplasmin level for WD was 1.4%. A1AT<br />

levels were requested for 3247 patients. The mean age of<br />

patients was 47 years, 55.9% were male and 88 tests (2.7%)<br />

were abnormal. Genotyping was performed on 55 patients<br />

(62.5%) with MZ (41.8%) most prevalent followed by MM<br />

(38.2%). Only 1 patient had ZZ genotype and was diagnosed<br />

with A1AT deficiency. Five patients had SZ genotype however<br />

only three had this noted in medical records. Of the 21 patients<br />

with MM genotype, 6 have been referred for further genotyping<br />

of rare genetic variants. 1 has shown MM procida<br />

. Other<br />

diagnoses included viral hepatitis (44.3%), NAFLD (18.2%),<br />

alcohol (11.4%) and drug induced liver injury (6.8%). The<br />

positive predictive value of a low A1AT level to detect ZZ genotype<br />

was 1.8%. Conclusion Caeruloplasmin and A1AT have<br />

very low detection rates when used as screening tests within the<br />

hepatology department of a tertiary hospital. The use of these<br />

investigations should be limited to patients in whom initial liver<br />

screen is negative and clinical suspicion remains.<br />

Disclosures:<br />

Leon A. Adams - Patent Held/Filed: Quest diagnostics<br />

The following authors have nothing to disclose: Tim Mitchell, John Beilby, Gary P.<br />

Jeffrey, George Garas, John Joseph, Ric Rossi, Gerry C. MacQuillan<br />

Disclosures:<br />

The following authors have nothing to disclose: Veronica Marin, Natalia Rosso,<br />

Matteo Dal Ben, Alan Raseni, Cristina Degrassi, Claudio Tiribelli, Silvia Gazzin<br />

2099<br />

The real world experience of screening investigations<br />

for Wilson’s disease and α1-antitrypsin deficiency<br />

within a tertiary liver unit – is it worthwhile?<br />

Tim Mitchell 1 , John Beilby 2 , Gary P. Jeffrey 1 , Leon A. Adams 1 ,<br />

George Garas 1 , John Joseph 2 , Ric Rossi 2 , Gerry C. MacQuillan 1 ;<br />

1 Sir Charles Gairdner Hospital, Nedlands, WA, Australia; 2 Path-<br />

West, Nedlands, WA, Australia<br />

Introduction Wilson’s disease (WD) and α1-antitrypsin (A1AT)<br />

deficiency are rare but important causes of liver disease. The<br />

phenotypic presentation is varied and diagnosis can be difficult.<br />

Despite limitations as screening tests, caeruloplasmin<br />

and A1AT levels have been performed routinely at this center<br />

among patients presenting with liver disease. We reviewed the<br />

yield of these investigations to guide future ordering practices.<br />

Methods All serum caeruloplasmin and A1AT levels requested<br />

by the hepatology department at a tertiary hospital from the<br />

state reference laboratory were reviewed. The study period<br />

2100<br />

Relapse of Porphyria Cutanea Tarda After Achieving<br />

Remission With Phlebotomy or Low Dose Hydroxychloroquine<br />

Ashwani K. Singal 1 , Eric Gou 2 , Maira Rizwan 1 , Marisol Albuerne 2 ,<br />

Csilla Kormos Hallberg 3 , Karl E. Anderson 3,2 ; 1 Gastroenterology<br />

and Hepatology, UAB, Birmingham, AL; 2 Internal Medicine,<br />

UTMB, Galveston, TX; 3 Preventive Medicine, UTMB, Galveston, TX<br />

Background: Porphyria cutanea tarda (PCT) is due to inhibition<br />

of hepatic uroporphyrinogen decarboxylase (UROD) and is<br />

effectively treated by phlebotomy (to serum ferritin of


1232A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(38%) patients relapsed biochemically (30% of 27 after phlebotomy<br />

and 50% of 18 after HCQ; Log Rank P=0.14). Relapse<br />

rates at 1, 3, and 5 years after phlebotomy or HCQ were<br />

8% vs. 24%, 30% vs. 30%, and 44% vs. 50% respectively.<br />

Relapse rates after these treatments did not differ by initial<br />

age, sex, race, alcohol use, smoking, HFE (hemochromatosis)<br />

genotypes C282Y/C282Y or C282Y/H63D, estrogen use,<br />

serum ferritin, or time to remission. Compared to non-relapsers,<br />

relapsing patients had higher pretreatment plasma porphyrin<br />

concentrations (14±20 vs.6±5 mcg/dL; P=0.046) and a higher<br />

proportion with HCV (94% vs. 71%, P=0.055). All remained<br />

abstinent from alcohol during treatment. But resumption of alcohol<br />

use tended to be more common among relapsers compared<br />

to non-relapsers (67 vs. 43%, P=0.1). Pooled data on relapse<br />

rate from 14 published <strong>studies</strong> on 1073 PCT patients showed<br />

relapse rates of 13.5% per year with phlebotomy and 17.9%<br />

per year with low dose hydroxychloroquine or chloroquine.<br />

The end-point of low dose HCQ or chloroquine treatment in all<br />

of these previous <strong>studies</strong> was normalization of urinary porphyrins<br />

with variable cut-off to define normal levels. Conclusions:<br />

PCT relapses may be more frequent with more severe disease<br />

and after achieving remission with low-dose HCQ than with<br />

phlebotomy. Longer follow-up and a larger sample size are<br />

planned with the Porphyrias Consortium study. The optimal<br />

length and end-point of treatment with low dose hydroxychloroquine<br />

need to be better defined.<br />

Disclosures:<br />

Karl E. Anderson - Consulting: Mitsubishi Tanabe Pharma America, Inc.<br />

The following authors have nothing to disclose: Ashwani K. Singal, Eric Gou,<br />

Maira Rizwan, Marisol Albuerne, Csilla Kormos Hallberg<br />

2101<br />

Hepatocellular Carcinoma in Patients with Wilson’s Disease:<br />

Results of a Single Centre Registry of 262 Patients<br />

Harshad Devarbhavi, Mallikarjun Patil; Department of Gastroenterology,<br />

St. John’s Medical College Hospital, Bangalore, India<br />

Introduction: Although hepatocellular carcinoma (HCC) is a<br />

well recognized complication of cirrhosis of various etiologies,<br />

HCC is believed to be extremely rare in Wilson disease.<br />

Around 30 cases have been reported worldwide illustrating<br />

the rarity of this complication. We undertook this study<br />

to determine the frequency and characteristics of HCC in a<br />

cohort of 262 patients over 19 years Patients and methods:<br />

We reviewed the case records of patients with Wilson disease<br />

from the Wilson disease registry maintained from 1996 to<br />

2014. Diagnosis of Wilson disease was based on a score ><br />

4 developed by the International group at Leipzig (J Hepatol<br />

2012;56:671-85). Patients were treated with D-penicillamine<br />

with or without Zinc Sulfate. Patients were regularly followed<br />

up at 3-6 monthly intervals. Patients with HCC were initially<br />

assessed by ultrasonography of abdomen followed by a contrast<br />

enhance CT scan. Results: We identified 3 cases of HCC<br />

among 262 patients with Wilson disease. All 3 had cirrhosis<br />

with portal hypertension at the time of diagnosis as evidenced<br />

by esophageal varices and ascites. The characteristics of these<br />

patients are depicted in Table. All three had multicentric HCC<br />

and presented with resistant complications of portal hypertension<br />

such as variceal bleeding, resistant ascites and one with<br />

hemoperitoneum. Given the advanced nature of HCC at presentation<br />

only supportive treatment could be given. Conclusions:<br />

The incidence of HCC in a large cohort of 262 patients<br />

is 1.1%. Although low, patients with WD are also at risk for<br />

HCC. HCC appears to be aggressive in our cohort of Wilson<br />

disease. Patients with Wilson disease need to undergo regular<br />

screening for early detection of HCC similar to other etiologies.<br />

Those who are insufficiently chelated appear to be more at risk<br />

for HCC.<br />

Characteristics of three patients with HCC<br />

Disclosures:<br />

The following authors have nothing to disclose: Harshad Devarbhavi, Mallikarjun<br />

Patil<br />

2102<br />

Congenital glycosylation defects mimicking Wilsons disease<br />

in patients with unexplained elevated aminotransferases<br />

and low ceruloplasmin<br />

Jos Jansen 1,2 , Joost Drenth 1 , Dirk Lefeber 2 , Monique van Scherpenzeel<br />

2 ; 1 Gastroenterology & Hepatology, Radboud University<br />

Medical Center Nijmegen, Utrecht, Netherlands; 2 Translational<br />

Metabolic Laboraty, glycosylation disorders, Radboud University<br />

Medical Center Nijmegen, Nijmegen, Netherlands<br />

Introduction Congenital disorders of glycosylation (CDGs) are<br />

a heterogenous group of autosomal recessive metabolic disorders<br />

with abnormal glycosylation as a hallmark. GI tract and<br />

liver symptoms are frequent but often secondary. Here we show<br />

that mutations in two uncharacterized genes are the cause<br />

for a primarily Wilson disease-like liver phenotype in twelve<br />

patients with unexplained glycosylation defects. Methods &<br />

Results These twelve patients (gene A: 5 pt., gene B: 8 pt.)<br />

showed an overlapping phenotype with elevated aminotransferases,<br />

high bone-derived alkaline phosphatase, hypercholesterolemia<br />

and elevated LDL-C. Further hepatic analysis showed<br />

low serum ceruloplasmin levels and liver biopsy indicated fibrosis,<br />

cirrhosis, steatosis and mild hepatic copper accumulation<br />

for some patients. Despite similarities, differences between the<br />

two gene defects were noticed. Gene A patients had a milder<br />

phenotype with slightly elevated aminotransferases and hypercholesterolemia.<br />

In contrast, gene B patients developed hepatosplenomegaly,<br />

resembling a glycogen or lysosomal storage<br />

disease, and end stage liver disease was seen for four patients.<br />

Additionally, for gene B, neurological symptoms were seen.<br />

Disease gene identification was done by functional analysis<br />

of exome sequencing. Based on yeast glycosylation models<br />

and BLAST searches two uncharacterized human proteins were<br />

identified. Subsequently analysis of raw exome sequencing<br />

data indicated incomplete coverage of gene A in a family<br />

where previous linkage and standard exome sequencing filtering<br />

was unsuccessful. Additional screening revealed six families<br />

with missense mutations, all segregating via an autosomal<br />

recessive inheritance. We analyzed intact serum transferrin<br />

via mass spectrometry for abnormal glycosylation and found<br />

a specific Golgi homeostasis defect profile. Additionally, we<br />

incubated patient fibroblasts with metabolic labeled sialic acid.<br />

Upon incorporation, the sialic acid gives a fluorescent signal.<br />

We show that fluorescence is reduced by app. 50% in tested<br />

patients. We transiently transfected HeLa cells with V5-tagged<br />

gene constructs and showed via immunofluorescence that<br />

both proteins are located in the ER-to-Golgi region. Together


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1233A<br />

with lipid vacuolization and dilated ER and Golgi on electron<br />

microscopy, these data point towards an effect on Golgi<br />

homeostasis for both defects. How this influences copper and<br />

lipid metabolism is still unknown. Conclusion We identified two<br />

new genes associated with unexplained elevated aminotransferases,<br />

low ceruloplasmin and hypercholesterolemia in twelve<br />

patients with abnormal glycosylation and suggest screening for<br />

glycosylation defects in these patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jos Jansen, Joost Drenth, Dirk<br />

Lefeber, Monique van Scherpenzeel<br />

2103<br />

Functional analysis and drug response to zinc and<br />

D-penicillamine in stable ATP7B mutant hepatic cell lines<br />

Gursimran Chandhok, Vanessa Sauer, Christoph Niemietz, Andree<br />

Zibert, Hartmut H. Schmidt; Klinik für Transplantationsmedizin, Universitätsklinikum<br />

Münster, Münster, Germany<br />

Objective: To assess in vitro functional analysis and drug<br />

response to zinc (Zn) and D-penicillamine (DPA) treatment of<br />

most common disease-causing ATP7B mutations observed in<br />

Wilson disease (WD) patients. Background: WD is a rare autosomal<br />

recessive disorder caused due to mutations in the copper<br />

transporter ATP7B. More than 600 disease-causing mutations<br />

are described. If not treated, WD is fatal. Commonly used<br />

drugs to reverse Cu toxicity in WD include direct chelation by<br />

DPA or metallothionein induction using Zn salts. The impact of<br />

these drugs in rescuing hepatocytes carrying individual mutations<br />

is unknown. Methods: Frequent homozygous mutations in<br />

Europe/US, Asia, and India were selected for functional analysis.<br />

An ATP7B knockout hepatic cell line previously established<br />

by our group (PLoS ONE, 9: e98809, 2014) was used to generate<br />

stable mutant cell lines. ATP7B gene and protein expression<br />

were determined for all mutant cell lines. Cu induced cell<br />

survival, apoptosis, and intracellular trafficking were studied.<br />

Cell viability against Cu toxicity was determined on treatment<br />

with Zn, DPA, or both. Results: ATP7B mRNA was similar<br />

whereas protein expression varied among the different mutant<br />

cell lines. Co-localization <strong>studies</strong> with late endosome lysosome<br />

marker lamp2 suggested impaired trafficking in presence or<br />

absence of Cu. All mutant cell lines showed different grades of<br />

ATP7B activity against Cu induced toxicity and apoptosis with<br />

p.H1069Q displaying moderate ATP7B activity. On treatment<br />

with Zn, DPA, or both most cell lines showed a characteristic<br />

response to treatment. DPA was more effective than Zn in rescuing<br />

cells against Cu toxicity. Only combination with Zn and<br />

DPA treatment was able to rescue the mutant cells similar to<br />

wild-type. Conclusions: The study provided insights into genotype-phenotype<br />

correlation in WD. Different ATP7B mutations<br />

showed varying sensitivity to Cu with high/moderate and low<br />

ATP7B activity. Individual ATP7B mutations showed a genotype-specific<br />

response to treatment. Combination treatment with<br />

Zn and DPA showed maximal survival of hepatic cells. Our<br />

data suggest that prognosis of WD patients may benefit from<br />

knowledge of mutation-specific in vitro functional analysis and<br />

use of Zn/DPA combination therapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Gursimran Chandhok, Vanessa<br />

Sauer, Christoph Niemietz, Andree Zibert, Hartmut H. Schmidt<br />

2104<br />

Cardiac MRI for Assessment of Iron Overload in Cirrhotic<br />

Patients Evaluated for Liver Transplantation<br />

R Todd Frederick 1 , Christopher R. Kagay 2 , Peter Hui 3 , Richard<br />

E. Shaw 3 , Robert W. Osorio 1 ; 1 Department of Transplantation,<br />

California Pacific Medical Center, San Francisco, CA; 2 Radiology,<br />

California Pacific Medical Center, San Francisco, CA; 3 Cardiology,<br />

California Pacific Medical Center, San Francisco, CA<br />

Background and Aims: Pathologic iron overload of the liver<br />

is often associated with cardiac iron overload and can have<br />

deleterious effects on cardiac function and compromise postliver<br />

transplant (LT) outcomes. We reviewed the experience<br />

at our center using cardiac MRI to estimate cardiac iron content<br />

in the evaluation of cirrhotic patients for LT. Methods: All<br />

patients undergoing evaluation for LT with moderate to severe<br />

hepatic iron overload (>2+ iron stain and/or >4000mcg/gm<br />

by liver biopsy or liver MRI) required cardiac MRI with calculation<br />

of T2* signal decay (GE StarMap). A minimum of<br />

3 regions of interest were selected and the mean T2* was<br />

determined (msec). Endomyocardial biopsy (EMBx) was performed<br />

for patients with suspected iron overload meeting all<br />

other LT criteria. EMBx was graded as normal (no iron), “borderline”<br />

(trace or 1+), “moderate” (2-3+), or “severe” (>3-4+).<br />

Comparison of T2* with iron content on EMBx was assessed<br />

using ANOVA. Results: 28 patients with hepatic iron overload<br />

underwent cardiac MRI between 6/2012 - 4/2015. Mean<br />

age was 54 (26-66), 71% male, 68% Caucasian. Etiology of<br />

liver disease included alcohol (9), HCV (5), HCV/alcohol (4),<br />

hereditary hemochromatosis (4), cryptogenic/biliary (3), sickle<br />

cell (1), HBV (1), and NASH (1). Mean transferrin saturation<br />

was 86.1% and mean ferritin 1600 ng/mL. The mean T2*<br />

was 37.2 msec (range 11.75-88.5, normal > 40). 15 patients<br />

underwent EMBx. 7 were considered “positive” for iron (3<br />

severe and 4 moderate), 4 were borderline, and 4 were negative.<br />

All 7 patients with “positive” EMBx were declined for LT<br />

at our center per protocol; 5 subsequently died (one following<br />

LT at another center), while 2 are awaiting LT at other centers.<br />

Of the 4 “borderline” patients, 1 underwent LT with complications,<br />

1 was declined for CAD (died), 1 died of sepsis, and 1<br />

remains listed. Of the 4 with negative cardiac iron, 1 received<br />

LT, 1 died of HCC, and 2 remain listed. The mean values of<br />

T2* were significantly different when compared across EMBx<br />

iron content groups (p=0.001). Using a T2* cutoff of 30msec<br />

provided 100% PPV and NPV for EMBx for “positive” vs. negative<br />

or “borderline” iron. Conclusions: Cardiac MRI estimation<br />

of myocardial iron utilizing T2* accurately identifies endomyocardial<br />

iron loading as identified by EMBx and may be useful<br />

in stratifying risk prior to liver transplantation.<br />

Comparison of T2* with EMBx iron (p= 0.001)<br />

Disclosures:<br />

R Todd Frederick - Advisory Committees or Review Panels: Gilead, Bristol Myers<br />

Squibb, Salix, Vital Therapies, Hyperion, AbbVie<br />

The following authors have nothing to disclose: Christopher R. Kagay, Peter Hui,<br />

Richard E. Shaw, Robert W. Osorio


1234A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2105<br />

Genetic variants associated with liver disease in α-1-antitrypsin<br />

deficiency<br />

Richard J. Thompson 1 , Jeid Turquet 1 , Rosamond Nuamah 2 , Patrick<br />

J. McKiernan 3 , Robert Stockley 4 , Gerome Breen 5 , Nedim Hadzic 6 ;<br />

1 Institute of Liver Studies, King’s College London, London, United<br />

Kingdom; 2 BRC Genomics Facility, Guy’s and St Thomas’ NHS<br />

Foundation Trust, London, United Kingdom; 3 Liver Unit, Birmingham<br />

Children’s Hospital, Birmingham, United Kingdom; 4 Lung<br />

Immuno Biochemical Research Laboratory, University of Birmingham,<br />

Birmingham, United Kingdom; 5 Institute of Psychiatry, King’s<br />

College London, London, United Kingdom; 6 Paediatric Liver, GI<br />

and Nutrition Centre, King’s College Hospital, London, United<br />

Kingdom<br />

Background Alpha-1-antitrysin is a major plasma serine protease<br />

inhibitor. It is encoded by SERPINA1. Several deficient<br />

alleles have been described. The Z allele (p.E342K) leads to<br />

protein polymerisation and intracellular retention. In homozygous<br />

form this allele is seen in patients with lung and liver<br />

disease; α-1-antitrypsin deficiency (A1ATD). The respiratory disease<br />

is highly penetrant, whereas neonatal cholestasis is seen<br />

in only 10-15% of Z homozygotes. The cause of the reduced<br />

penetrance seen in this liver phenotype is unexplained. Aim<br />

To identify genetic loci associated with neonatal presentation<br />

of A1ATD associated liver disease. Methods All patients were<br />

homozygous for p.E342K, the Z allele. One group consisted<br />

of 384 patients followed in an adult respiratory clinic. The<br />

second group consisted of 132 patients presenting with neonatal<br />

cholestasis. All were genotyped using Illumina Infinium<br />

HumanCoreExome arrays; interrogating >240,000 tagSNPs.<br />

Data was analysed using several software tools including<br />

GenomeStudio, PLINK and PRSice. Thus allowed single locus<br />

and polygenic associations to be identified. Results Differences<br />

in SNP allele frequencies between the 2 groups of patients<br />

were studied. Three separate loci with a lod score of >6 were<br />

identified. They are on chromosomes 1, 12 and 14. The latter<br />

is 10 Mb from the SERPINA1 gene. A large number of other<br />

loci have been subject to further analysis, looking for polygenic<br />

and pathway effects. Only 1 of the three loci lies within a gene.<br />

This in turn has not been linked to liver disease. Conclusions<br />

Although AIATD-associated liver disease is thought of as being<br />

genetic, the major locus (SERPINA1) shows markedly reduced<br />

penetrance. The fact that all Z alleles are inherited from a<br />

relatively recent ancestor means that all affected individuals<br />

are much more closely related than in nearly all similar <strong>studies</strong>.<br />

This in turn means that between phenotypically diverse<br />

groups, genetic associations can be identified using relatively<br />

small sample sizes. This proved to be the case and allowed the<br />

identification of 3 quite separate loci conferring susceptibility<br />

to liver disease. The mechanisms by which this risk is conferred<br />

remain to be identified, however this is the first step in the identification<br />

of the cause of A1ATD liver disease.<br />

Disclosures:<br />

Richard J. Thompson - Grant/Research Support: Shire; Speaking and Teaching:<br />

Shire<br />

Patrick J. McKiernan - Consulting: Alnylam Pharmaceuticals<br />

Nedim Hadzic - Consulting: Alnylam; Speaking and Teaching: Synageva<br />

The following authors have nothing to disclose: Jeid Turquet, Rosamond Nuamah,<br />

Robert Stockley, Gerome Breen<br />

2106<br />

Clinical features, biochemical measures and liver histology<br />

in Hereditary Hemochromatosis Argentinian<br />

patients with and without HFE gene mutations.<br />

Florencia Yamasato, Marcelo Castro, Alejandra Avagnina, Alejandra<br />

Vellicce, Alejandra Pernazza, Jorge Rey, Esteban González<br />

Ballerga, Juan A. Sorda, Jorge Daruich; Gastroenterologia, Hospital<br />

de Clínicas José de San Martín. Universidad de Buenos Aires,<br />

Buenos Aires, Argentina<br />

Introduction: Inclusion of HFE gene mutations, present in<br />

90-96% of Hereditary Hemochromatosis (HH) Anglo-Saxon<br />

patients (Pt), has been a great advance in diagnosis algorithm.<br />

However, the percentage of these mutations among HH Pt<br />

could be different in our population. Objectives: To describe<br />

in HH Pt, the differences in clinical and biochemical features,<br />

and the presence of cirrhosis among subjects with HFE gene<br />

mutations (G1) and those without them (G2). Materials and<br />

Methods: Cross sectional study. 277 Pt with HH diagnosis<br />

between 1989 and 2014 were included. Only Pt with studied<br />

HFE gene mutations were included. HH diagnosis was<br />

performed by internationally accepted criteria and compatible<br />

liver histology. HOMA, Diabetes Mellitus (DM), alcohol<br />

intake (>30g/d), articular and cardiac features, liver histology<br />

(cirrhosis) and non-alcoholic fatty liver disease (NAFLD) were<br />

investigated and compared between both groups. Statistical<br />

analysis: Descriptive statistics are displayed as Median values<br />

with 95% Confidence Intervals (CI). Mann Whitney method<br />

and logistic regressions were performed. P Values


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1235A<br />

2107<br />

Mutational analysis of ATP7B in Chinese Wilson disease<br />

patients<br />

Shan S. Peng, Lingxia QI, yuanyuan Cui, Wenli Kong, Linyuan<br />

Ma, Yu Pan, Junqi Niu, Rui Hua; Hepatology, The First Hospital,<br />

Jilin University, Changchun, China<br />

Wilson Disease (WD) is an inborn error of copper metabolism<br />

inherited in an autosomal recessive manner caused by<br />

the mutations in the P-type ATPase gene (ATP7B). In this study,<br />

we screen and detect the mutations of the ATP7B gene in unrelated<br />

Chinese WD patients.A total of 62 individuals from ten<br />

provinces of china with WD were recruited. Of them, 41 were<br />

males and 21 were females, and their onset ages were from 5<br />

to 61 years with a median age of 26 years. The full length of<br />

ATP7B gene of the patients was sequenced and aligned to the<br />

referred ATP7B gene sequence. The results suggested that all<br />

the patents carried with mutations and 48 different mutations<br />

were identified, of which 35 were missense, one was nonsense,<br />

two were splicing, and 10 were frameshift (five insertion<br />

and five deletion). Among these mutations, c.2333G>T<br />

(32.3%), c.2975C>T (11.5%), and c.3443T>C (10.4%) were<br />

the most prevalent mutants and c.2310C>G always linked with<br />

c.2333G>T. The eighth, 11 th , and 18 th exons carried more<br />

mutations (6/48, 5/48, and 5/48, respectively). After comparing<br />

with the mutations reported previously, 19 out of the 48<br />

mutations were novel made up of all the splicing and frameshift<br />

mutations and 10 of 35 missense mutations. Two popular algorithms<br />

were used to predict the effects of the novel mutations<br />

on ATP7B function and most of the mutations were unfavorable<br />

(18/21). Our study will broaden our knowledge about ATP7B<br />

mutations in WD patients in China and further analysis need to<br />

examine the relationship between these mutations and clinical<br />

characteristics of WD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Shan S. Peng, Lingxia QI,<br />

yuanyuan Cui, Wenli Kong, Linyuan Ma, Yu Pan, Junqi Niu, Rui Hua<br />

2108<br />

UGT1A regulation by a member of the proto-oncogenic<br />

miR-106b-25 cluster: hsa-miR-25<br />

Sandra Kalthoff, Anja Winkler, Christian P. Strassburg; Department<br />

of Internal Medicine, University Hospital Bonn, Bonn, Germany<br />

Introduction: UDP-glucuronosyltransferases (UGTs) catalyze<br />

the detoxification of xenobiotics including a variety of carcinogens.<br />

Therefore, UGTs play an essential role in defence<br />

mechanisms protecting the organism against the development<br />

of cancer. Recent research has shown that deregulation of<br />

miRNAs can contribute to cancer development. In hepatocellular<br />

carcinoma (HCC) tissues, hsa-miR-25, for example, is<br />

overexpressed and associated with a poor prognosis in HCC<br />

patients. Aim of this study was to examine the influence of hsamiR-25<br />

on UGT1A-expression. Methods: For luciferase assays,<br />

reporter gene constructs containing different UGT1A promoters<br />

(UGT1A1, UGT1A3, UGT1A4, UGT1A7, UGT1A9),<br />

combined with luciferase sequence and the UGT1A 3’untranslated<br />

region (UTR) sequence were generated. Every promoter<br />

was combined with either the UGT1A 3’UTR-wild type(WT)<br />

sequence or the UGT1A 3’UTR-SNP sequence (containing the<br />

3 SNPs rs10929303, rs1042640 and rs8330). The constructs<br />

were transfected together with 5 nM hsa-miR-25 into HepG2<br />

cells. For RNA quantification HepG2 cells were treated with 5<br />

nM hsa-miR-25 and isolated RNA was analyzed by TaqMan<br />

PCR. Results: Transfection of hsa-miR-25 led to a significant<br />

decrease in luciferase expression of the UGT1A1-, UGT1A3-,<br />

UGT1A4- UGT1A7 and UGT1A9 reporter gene constructs (0.4-<br />

0.5 fold). Interestingly, transfection of the construct containing<br />

the UGT1A7 promoter combined with the UGT1A 3’UTR-SNP<br />

sequence, led to a further reduction of luciferase expression<br />

compared to the UGT1A 3 ‘UTR-WT sequence (SNP: 0.21 fold<br />

compared to UGT1A 3’UTR-WT 0.42 fold). RNA analysis of<br />

HepG2 cells revealed that mRNA expression of UGT1A1 (0.47<br />

fold), UGT1A3 (0.49 fold), UGT1A4 (0.68 fold), UGT1A6<br />

(0.6 fold), UGT1A7 (0.38 fold) and UGT1A9 (0.36 fold) were<br />

significantly reduced by treatment with 5 nM hsa-miR-25. Conclusions:<br />

This is the first demonstration of UGT1A regulation<br />

by hsa-miR-25. All UGT1A genes were significantly downregulated<br />

by treatment with 5 nM hsa-miR-25 on mRNA level as<br />

well in reporter gene assays. The presence of 3 SNPs in the<br />

UGT1A 3’untranslated region combined with the UGT1A7<br />

promoter led to a further reduction of luciferase expression<br />

indicating a suppressive effect of theses SNPs on the activity<br />

of the major carcinogen detoxifying UGT1A7 enzyme in the<br />

presence of overexpressed hsa-miR-25. The elucidated significant<br />

reduction of UGT1A expression and the accompanying<br />

affected detoxification capacity mediated by hsa-miR-25 may<br />

represent a possible explanation for the potential association<br />

between hsa-miR-25 and HCC development.<br />

Disclosures:<br />

Christian P. Strassburg - Advisory Committees or Review Panels: Novartis, Roche;<br />

Speaking and Teaching: Novartis, Merz, MSD, Falk Pharma, BMS, Abbvie<br />

The following authors have nothing to disclose: Sandra Kalthoff, Anja Winkler<br />

2109<br />

Wilson disease: gestational methyl group supplementation<br />

affects hepatic oxidative phosphorylation and<br />

movement disorder pathways in fetal mouse liver<br />

Valentina Medici, Noreene Shibata, Charles H. Halsted, Janine M.<br />

LaSalle; University of California Davis, Sacramento, CA<br />

Background & Hypothesis. Wilson Disease (WD) is a condition<br />

with highly variable phenotypic presentation, including<br />

hepatic and neurological involvement. Previously we showed<br />

that methionine metabolism, central in the regulation of methylation<br />

reactions, is altered in the tx-j mouse model of WD and<br />

therefore may influence the expression of genes involved in<br />

liver damage. Since maternal diet is known to affect fetal gene<br />

transcript levels through epigenetic mechanisms, we hypothesize<br />

the phenotype of WD is influenced by maternal methylation<br />

status. Methods. The tx-j mouse model of WD and wildtype<br />

C3H female mice were started on choline-supplemented (choline<br />

36 mmol/Kg of diet) and control diets (choline 8 mmol/<br />

Kg of diet) 2 weeks before mating and through pregnancy<br />

to embryonic day 17. Livers of 6-11 embryos/dam were collected<br />

for RNA-sequencing. Pathway analysis was performed<br />

by DAVID software. An additional group of offspring was<br />

euthanized at 24 weeks to confirm persistence of changes in<br />

gene expression in adult livers. Results. Comparing tx-j and<br />

wildtype fetal livers, pathway analysis revealed differences<br />

in transcript levels of nuclear encoded genes related to oxidative<br />

phosphorylation, which overlapped with genes related to<br />

neurological movement disorders, including Huntington’s and<br />

Parkinson’s disease, both conditions listed in the differential<br />

diagnosis of WD. Maternal choline supplementation was associated<br />

with maintenance of gene transcript levels at control<br />

group levels in fetal livers. Transcript levels of neurexin-1, central<br />

to cell adhesion in the nervous system, was up-regulated<br />

67-fold in fetal livers of tx-j mice compared to wildtype mice,<br />

but were maintained at control levels in offspring of choline<br />

supplemented dams. In 24-week old tx-j offspring mouse livers,<br />

we found persistent changes in transcript levels of nuclear


1236A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

encoded genes related to oxidative phosphorylation, including<br />

succinate dehydrogenase A (Sdha), a major catalytic subunit<br />

of the succinate-ubiquinone oxidoreductase, a complex of the<br />

mitochondrial respiratory chain. Of note, maternal choline<br />

maintained control levels of gene transcript levels of Sdha.<br />

Conclusions. Fetal and 24 week old livers of a mouse model<br />

of WD are characterized by changes in transcript levels of<br />

genes related to the oxidative phosphorylation pathway which<br />

is also involved in neurological diseases and these findings<br />

were prevented by maternal choline supplementation. Since<br />

these findings were present in both fetal and 24 week old tx-j<br />

livers, they may influence WD phenotype.<br />

Disclosures:<br />

The following authors have nothing to disclose: Valentina Medici, Noreene Shibata,<br />

Charles H. Halsted, Janine M. LaSalle<br />

2110<br />

Association between Psoas Muscle FDG Uptake and<br />

Metabolic Syndrome by 18F-Fluorodeoxyglucose Positron<br />

Emission Tomography: May psoas muscle FDG<br />

uptake use as a biomarker for impaired metabolic<br />

health?<br />

Seung Min Lee 1 , Dae Won Jun 2 , Yong Kyun Cho 3 , Eun Chul Jang 4 ,<br />

Joo Hee Kwak 2 ; 1 Department of Food and Nutrition, Sungshin<br />

Women’s University, Seoul, Korea (the Republic of); 2 Department<br />

of Internal Medicine, Hanyang University College of Medicine,<br />

Seoul, Korea (the Republic of); 3 Department of Internal Medicine,<br />

Kangbuk Samsung Hospital, Sungkyunkwan University, School of<br />

Medicine, Seoul, Korea (the Republic of); 4 Department of Occupational<br />

and Environmental Medicine, Soonchunhyang University<br />

Cheonan Hospital, Cheonan, Korea (the Republic of)<br />

Purpose: To identify metabolic activity of several organs<br />

responsible for the glucose metabolism on 18 F-FDG PET/CT<br />

according to several metabolic phenotypes and to evaluate<br />

whether it could be used to predict the metabolic health impairment.<br />

Methods: We retrospectively analyzed the data from<br />

157 subjects who underwent 18 F-FDG PET/CT for health medical<br />

examination. Liver function test, total cholesterol, triglyceride,<br />

and fasting blood glucose (FBG) level as well as presence<br />

of metabolic syndrome (MS) were evaluated in these subjects.<br />

Fixed or flexible spherical volume of interest (VOIs) were used<br />

to evaluate maximal and/or mean standardized uptake value<br />

(SUV) to assess the metabolism of liver, pancreas, mesenteric<br />

visceral fat, psoas muscle, and abdominal subcutaneous fat on<br />

18 F-FDG PET/CT. Effect of SUVs in each organ on the clinical<br />

metabolic parameters as well as on the presence of MS were<br />

analyzed for statistical significance. Results: 52 subjects (33.1<br />

%) were obese with a body mass index (BMI) > 25 and forty<br />

subjects (40/157, 25%) had MS. SUVmax of psoas muscle,<br />

mesenteric visceral fat, and abdominal subcutaneous fat as<br />

well as SUVmean of liver were positively correlated with BMI,<br />

weight and FBG, after adjusted by FBG. SUVmax of psoas<br />

muscle, visceral fat, and pancreas were significantly higher in<br />

the subjects with obesity and in the subject with MS (p for trend<br />

< 0.05). Among them, SUVmax of psoas muscle was found<br />

to be only risk factor for the presence of MS independent to<br />

weight and FBG. With the cutoff value of 1.34, SUVmax of<br />

psoas muscle accurately predicted MS, central obesity, and<br />

hypertension (AUROC 0.779, 0.810, and 0.646, respectively,<br />

p < 0.05). Conclusion: SUVmax of psoas muscle on the 18 F-<br />

FDG PET/CT is well associated with various clinical metabolic<br />

parameters and it could be used to predict the metabolic health<br />

impairment. Moreover, it could be valuable tool to sort metabolically<br />

healthy or abnormal especially in the obese subjects.<br />

Disclosures:<br />

The following authors have nothing to disclose: Seung Min Lee, Dae Won Jun,<br />

Yong Kyun Cho, Eun Chul Jang, Joo Hee Kwak<br />

2111<br />

Pitfalls in Erythrocyte Protoporphyrin Measurement for<br />

Diagnosis and Monitoring of Protoporphyrias<br />

Eric Gou 1 , Manisha Balwani 3 , D Montgomery Bissell 4 , Joseph<br />

R. Bloomer 5 , Herbert L. Bonkovsky 2 , Robert J. Desnick 3 , John D.<br />

Phillips 6 , Ashwani Singal 5 , Bruce M. Wang 4 , Hetanshi Naik 3 ,<br />

Karl E. Anderson 1 ; 1 Preventive Medicine and Community Health,<br />

University of Texas Medical Branch, Galveston, TX; 2 Wake Forest<br />

University, Winston-Salem, NC; 3 Icahn School of Medicine at Mt.<br />

Sinai, New York, NY; 4 University of California San Francisdo, San<br />

Francisco, CA; 5 Universtiy of Alabama Birmingham, Birmingham,<br />

AL; 6 University of Utah, Salt Lake City, UT<br />

Erythropoietic protoporphyria (EPP) is due to the inherited<br />

deficiency of ferrochelatase, the enzyme that catalyzes the<br />

chelation of iron with protoporphyrin IX to complete heme biosynthesis.<br />

Increased erythrocyte and plasma protoporphyrin<br />

levels in EPP cause painful photosensitivity and impair quality<br />

of life. EPP is the third most common porphyria, the most common<br />

in children and its diagnosis is often delayed. Diagnosis<br />

of EPP requires: (1) a marked increase in total erythrocyte protoporphyrin<br />

(300-5,000 ug/dL erythrocytes) and (2) a predominance<br />

(85-100%) of metal-free protoporphyrin. X-linked<br />

protoporphyria (XLP) is less common and is due to gain of function<br />

mutations of the erythroid form of d-aminolevulinic acid<br />

synthase, the first enzyme in the heme biosynthetic pathway.<br />

XLP causes a similar increase in total erythrocyte protoporphyrin<br />

but with a somewhat lower fraction of metal-free protoporphyrin<br />

(50-85% of the total). The Porphyrias Consortium (PC)<br />

has enrolled more than 180 patients with EPP and XLP. Among<br />

the few patients with original laboratory reports of erythrocyte<br />

protoporphyrin levels, 15 appeared to have much higher levels<br />

(4.7 to 46.7 fold) when enrolled in <strong>studies</strong> at a PC Center. The<br />

diagnosis had been missed in 2 of these patients based on<br />

the earlier spurious results. All of the prior reports were from<br />

either Quest or LabCorp, which use only hematofluorometry for<br />

assessing erythrocyte protoporphyrin. The hematofluorometer<br />

was developed to screen for lead poisoning, and measures<br />

zinc protoporphyrin by front surface fluorescence using a drop<br />

of blood. The instrument displays two calculations based on<br />

assumed hematocrits each specified by OSHA and CDC, and<br />

variously reports results as “free” and “zinc” protoporphyrin<br />

(with different reference ranges), implying separate measurements<br />

of metal-free and zinc protoporphyrin. These reports are<br />

misleading and impair diagnosis and monitoring of patients<br />

with protoporphyria. This is concerning because reproducible<br />

methods are needed for diagnosis and longitudinal monitoring<br />

of erythrocyte protoporphyrin, as rising levels may predict<br />

severe hepatic complications requiring liver transplantation.


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1237A<br />

We suggest that laboratories offering erythrocyte protoporphyrin<br />

measurement should prioritize testing for EPP and XLP,<br />

which are essential for diagnosis and monitoring of these conditions,<br />

and less important for diagnosis of lead poisoning, iron<br />

deficiency and other erythrocyte disorders. Terms and abbreviations<br />

used in reporting results should be accurately defined.<br />

Disclosures:<br />

Joseph R. Bloomer - Advisory Committees or Review Panels: Recordati Rare<br />

Diseases; Consulting: Mitsubishi Tanabe Pharma; Grant/Research Support: Clinuvel,<br />

Inc., American Porphyria Foundation, NIH 5U54 DK083909, Alnylam<br />

Pharmaceuticals<br />

Herbert L. Bonkovsky - Advisory Committees or Review Panels: Recordati Rare<br />

Chemicals, Clinuvel, Inc.; Consulting: Alnylam, Inc, Clinuvel, Inc., Clinuvel, Inc.;<br />

Grant/Research Support: Gilead Sciences<br />

Robert J. Desnick - Advisory Committees or Review Panels: Recordati Rare Diseases;<br />

Consulting: Alnylam Pharmaceuticals; Grant/Research Support: Alnylam<br />

Pharmaceuticals; Patent Held/Filed: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Karl E. Anderson - Consulting: Mitsubishi Tanabe Pharma America, Inc.<br />

The following authors have nothing to disclose: Eric Gou, Manisha Balwani, D<br />

Montgomery Bissell, John D. Phillips, Ashwani Singal, Bruce M. Wang, Hetanshi<br />

Naik<br />

2112<br />

Variation in erythrocyte and plasma protoporphyrin levels<br />

over time in protoporphyrias<br />

Eric Gou 1 , Cindy Weng 2 , John D. Phillips 2 , Tom Greene 2 , Manisha<br />

Balwani 4 , D Montgomery Bissell 6 , Joseph R. Bloomer 5 , Herbert L.<br />

Bonkovsky 3 , Robert J. Desnick 4 , Hetanshi Naik 4 , VM Sadagoparamanujam<br />

1 , Karl E. Anderson 1 ; 1 Preventive Medicine and Community<br />

Health, University of Texas Medical Branch, Galveston, TX;<br />

2 University of Utah, Salt Lake City, UT; 3 Wake Forest University,<br />

Winston-Salem, NC; 4 Icahn School of Medicine at Mt. Sinai, New<br />

York, NY; 5 University of Alabama, Birmingham, AL; 6 University of<br />

California, San Francisco, CA<br />

Background: Erythropoietic protoporphyria (EPP) and X-linked<br />

protoporphyria (XLP) cause painful, non-blistering cutaneous<br />

photosensitivity. EPP, the third most common porphyria and the<br />

most common in children, is due to decreased activity of ferrochelatase<br />

(FECH), which catalyzes the insertion of iron into protoporphyrin<br />

IX to complete heme synthesis. XLP is less common<br />

and is due to gain of function mutations of the erythroid form<br />

of d-aminolevulinic acid synthase, the first enzyme in the heme<br />

biosynthetic pathway. Hepatotoxic effects of excess protoporphyrin<br />

cause protoporphyric hepatopathy in less than 5% of<br />

patients. Hepatopathy impairs protoporphyrin excretion, which<br />

leads to further increases in circulating protoporphyrin levels<br />

and accelerated liver damage, often leading to liver transplantation.<br />

Normal variability in protoporphyrin levels over time has<br />

not been characterized in EPP and XLP, and whether threshold<br />

levels lead to hepatopathy is not known. Purpose: To compare<br />

porphyrin levels in EPP and XLP, and characterize expected<br />

variability of erythrocyte and plasma protoporphyrin levels in<br />

the absence of hepatopathy. Methods: Coefficients of variation<br />

(CV) were determined for levels of erythrocyte and plasma<br />

protoporphyrin repeated at least 4 times in 86 subjects with<br />

documented EPP or XLP. Patients with protoporphyric hepatopathy<br />

or elevated transaminases were excluded. Results: Total<br />

erythrocyte protoporphyrin was higher, on average, in XLP<br />

than in EPP, with considerable overlap. Total erythrocyte protoporphyrin<br />

levels were higher, on average, in males with XLP<br />

than females. Differences in erythrocyte protoporphyrin levels<br />

were lower between related than unrelated subjects with EPP or<br />

XLP. The ratio of metal-free to zinc protoporphyrin was higher<br />

in EPP, and was very stable over time. The CV for repeated<br />

values of total erythrocyte protoporphyrin averaged 19.59,<br />

and was not significantly different in EPP and XLP (19.06 and<br />

23.06 respectively). Plasma porphyrin levels were much lower,<br />

correlated only roughly with erythrocyte levels, and were much<br />

more variable over time. To date, none in this cohort has developed<br />

hepatopathy or sustained upward trends in protoporphyrin<br />

levels. Conclusion: Variation in erythrocyte protoporphyrin<br />

levels of up to ~25% may occur in EPP and XLP in the absence<br />

of liver disease. Studies in more patients over longer periods<br />

of time are in progress to see if greater increases or threshold<br />

levels of erythrocyte and plasma protoporphyrin precede the<br />

development of overt protoporphyric hepatopathy.<br />

Disclosures:<br />

Joseph R. Bloomer - Advisory Committees or Review Panels: Recordati Rare<br />

Diseases; Consulting: Mitsubishi Tanabe Pharma; Grant/Research Support: Clinuvel,<br />

Inc., American Porphyria Foundation, NIH 5U54 DK083909, Alnylam<br />

Pharmaceuticals<br />

Herbert L. Bonkovsky - Advisory Committees or Review Panels: Recordati Rare<br />

Chemicals, Clinuvel, Inc.; Consulting: Alnylam, Inc, Clinuvel, Inc., Clinuvel, Inc.;<br />

Grant/Research Support: Gilead Sciences<br />

Robert J. Desnick - Advisory Committees or Review Panels: Recordati Rare Diseases;<br />

Consulting: Alnylam Pharmaceuticals; Grant/Research Support: Alnylam<br />

Pharmaceuticals; Patent Held/Filed: Alnylam Pharmaceuticals; Stock Shareholder:<br />

Alnylam Pharmaceuticals<br />

Karl E. Anderson - Consulting: Mitsubishi Tanabe Pharma America, Inc.<br />

The following authors have nothing to disclose: Eric Gou, Cindy Weng, John D.<br />

Phillips, Tom Greene, Manisha Balwani, D Montgomery Bissell, Hetanshi Naik,<br />

VM Sadagoparamanujam<br />

2113<br />

A Nationwide, Population-based Epidemiology and<br />

Disease Burden of Wilson ’s disease in South Korea in<br />

2009-2013<br />

Joo Yeong Baeg 1 , Eun Sun Jang 1 , Mo-ran Ki 3 , Bo Hyun Kim 3 ,<br />

Kyung-ah Kim 4 , Hwa Young Choi 2 , Sook-Hyang Jeong 1 ; 1 Internal<br />

Medicine, Seoul National University College of Medicine,<br />

Seongnam-Si,, Korea (the Republic of); 2 Department of Cancer<br />

Control and Policy, Graduate School of Cancer Science and Policy,<br />

National Cancer Center, Goyang, Korea (the Republic of);<br />

3 Center for Liver Cancer, National Cancer Center,, Goyang,<br />

Korea (the Republic of); 4 Departments of Internal Medicine, Inje<br />

University Ilsan Paik Hospital, Goyang, Korea (the Republic of)<br />

Background/Aims: Wilson’s disease (WD) is an inherited autosomal<br />

recessive disorder of copper metabolism in which copper<br />

accumulates in the organ, particularly in the liver and the central<br />

nervous system. As a rare disease, population-based epidemiology<br />

of WD is largely unknown. This study aimed to investigate<br />

the nationwide, population-based prevalence, comorbidity,<br />

treatment regimen, and direct medical cost of WD in South<br />

Korea from 2009 to 2013. Methods: Using 2 big data source<br />

from Health Insurance Review and Assessment Service (HIRA)<br />

claims database and Rare Intractable Diseases registration program<br />

database from 2009 to 2013 in Korea, we identified<br />

all patients with WD registered by physician as the International<br />

Classification of Diseases (ICD-10) code for WD (E83.0),<br />

and with the registration record (V174) for the rare disease in<br />

Korea. The five-year data included 1,509 patients linked with<br />

information including age, gender, comorbidity ICD-10 codes,<br />

prescribed medications, and direct medical costs. Results: The<br />

overall crude prevalence of WD was 2.38/100,000 people<br />

and it showed increasing tendency during the 5 years (2.27,<br />

2.28, 2.35, 2.45, and 2.53/100,000 at 2009-2013, respectively).<br />

Most (72%) patients were diagnosed before 30 years<br />

old, and 1.4% had additional disease code for neurologic/<br />

psychologic disease. Liver cirrhosis and liver cancer were<br />

detected in 33.2% and 9.9%, respectively. Liver transplantation<br />

was performed in 4.6% of patients. D-penicillamine, trientine,<br />

and zinc were prescribed for 58.3%, 29.7%, and 13.9%<br />

of patients, respectively. Mean annual total medical cost per


1238A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

person for WD was 1,643.15 USD. Conclusion: The prevalence<br />

of WD was slightly lower in Korean than other countries,<br />

showing an increase from 2009. Since about one third of WD<br />

patients had advanced liver disease such as cirrhosis and cancer,<br />

more efforts for early diagnosis and treatment for WD are<br />

warranted, especially during pediatric age.<br />

Disclosures:<br />

The following authors have nothing to disclose: Joo Yeong Baeg, Eun Sun Jang,<br />

Mo-ran Ki, Bo Hyun Kim, Kyung-ah Kim, Hwa Young Choi, Sook-Hyang Jeong<br />

2114<br />

Relapse of porphyria cutanea tarda after achieving<br />

remission: A Meta-analysis<br />

Maira Rizwan 1 , Karl Anderson 2 , Ashwani K. Singal 1 ; 1 Gastroenterology<br />

and Hepatology, University of Alabama, Birmingham, AL;<br />

2 university of texas, Galveston, TX<br />

Background: Porphyria cutanea tarda (PCT) is due to inhibition<br />

of hepatic uroporphyrinogen decarboxylase (UROD).<br />

Phlebotomy and 4-aminoquinolines (hydroxychloroquine and<br />

chloroquine, preferably low-dose) are effective treatment<br />

options. Data comparing relapse after documenting biochemical<br />

remission of PCT with these treatments are lacking. Methods:<br />

Pubmed and Embase databases were searched for <strong>studies</strong><br />

reporting data on PCT relapse with minimum follow up for<br />

1 year after achieving remission with either phlebotomy or<br />

4-aminoquinolines. End point of treatment with phlebotomy<br />

was ferritin


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1239A<br />

gene, encoding alpha-1-antitrypsin (AAT) protein. AAT is a<br />

serine proteinase inhibitor that is synthesized in the liver and<br />

secreted into the bloodstream to protect the lungs from neutrophil<br />

elastase. Mutations in SERPINA1, most commonly the Z-allele<br />

(Z-AAT), cause the protein to misfold resulting in retention<br />

of aggregated AAT protein in hepatocytes and a corresponding<br />

decrease in circulating levels of AAT leading to Alpha-1-Antitrypsin<br />

Deficiency (A1ATD). Aggregation of Z-AAT protein in<br />

hepatocytes is associated with liver injury, fibrosis, cirrhosis,<br />

and hepatocellular carcinoma, while decreased circulating levels<br />

of AAT often presents as emphysema or chronic obstructive<br />

pulmonary disease. Using a lipid nanoparticle (LNP) system we<br />

developed for hepatic delivery of DsiRNAs, we demonstrate<br />

a significant reduction of human Z-AAT protein in the bloodstream<br />

and livers of the PiZ transgenic mouse model of hepatic<br />

A1ATD disease. Additionally, we provide evidence that targeting<br />

SERPINA1 with DsiRNA results in a reduction of A1ATD-related<br />

liver disease markers in PiZ transgenic mice. To further<br />

extend DsiRNA delivery approaches, we have also developed<br />

DsiRNA conjugates that can effectively deliver DsiRNAs to the<br />

liver without the need for LNP delivery systems. DsiRNA conjugates<br />

were prepared through the addition of N-acetyl galactosamine<br />

(GalNAc) sugars to the DsiRNA molecule to mediate<br />

delivery to hepatocytes via the highly expressed asialoglycoprotein<br />

receptor. SERPINA1-GalNAc conjugates delivered to<br />

PiZ transgenic mice effectively reduced Z-AAT protein levels.<br />

These results indicate Dicerna’s DsiRNA platform, delivered by<br />

either LNP-mediated or GalNAc-conjugated delivery systems,<br />

may be an effective therapeutic approach for hepatic disease.<br />

Disclosures:<br />

Rohan Diwanji - Employment: Dicerna Pharmceuticals; Stock Shareholder:<br />

Dicerna Pharmceuticals<br />

Utsav H. Saxena - Employment: Dicerna Pharmaceuticals<br />

Hank Dudek - Employment: Dicerna<br />

Nicole Avitahl-Curtis - Employment: Dicerna Pharmaceuticals, Inc.<br />

Chaitali Dutta - Employment: Dicerna Pharmaceuticals<br />

Benjamin Holmes - Employment: Dicerna Pharmaceuticals; Stock Shareholder:<br />

Dicerna Pharmaceuticals<br />

Marc Abrams - Employment: Dicerna Pharmaceuticals<br />

Kevin Craig - Employment: Dicerna Pharmaceuticals, Inc<br />

Luciano Apponi - Employment: Dicerna Pharmaceuticals<br />

Martin Koser - Employment: Dicerna Pharmaceuticals<br />

Chengjung Lai - Employment: Dicerna Pharmaceuticals<br />

Bob D. Brown - Employment: Dicerna Pharmaceuticals<br />

The following authors have nothing to disclose: Natalie W. Pursell, Wei Zhou,<br />

Weimin Wang, Dongyu Chen, Purva Pandya, Xue Shui, Boyoung Kim, Rachel<br />

Storr, Anee Shah, Edmond Chipumuro, Girish R. Chopda, Jr-Shiuan Yang, Marita<br />

S. Larsson Cohen, Naim Nazef, Nandini Venkat, Shanthi Ganesh, Venkat<br />

Krishnamurthy, Wendy Cyr, Zakir Siddiquee<br />

2117<br />

GNPAT Variant D519G in patients referred for HFE<br />

Hemochormatosis Testing<br />

Alexander Levstik 2 , Alan Stuart 2 , Paul Adams 1 ; 1 Gastroenterology,<br />

University Hospital, London, ON, Canada; 2 Laboratory Medicine,<br />

Western University, London, ON, Canada<br />

Background: Previous <strong>studies</strong> in high and low expressors has<br />

demonstrated that a variant in the GNPAT gene (D519G,<br />

Rs11558492, chromosome 1, exon 11) has been associated<br />

with severe iron overload in C282Y homozygotes for hemochromatosis<br />

(McLaren C, et al, Hepatology 2015). The GNPAT<br />

gene is associated with peroxisomal diseases and may be<br />

involved with transferrin receptor recycling. Gene silencing of<br />

GNPAT reduces intracellular hepcidin. In this study, a GNPAT<br />

variant was assessed prospectively in patients over a range of<br />

serum ferritin levels referred for HFE testing. Methods: Consecutive<br />

patients sent for HFE testing were studied for the GNPAT<br />

variant using a TaqMan kit assay (Life Technologies, Burlington,<br />

ON). The assay was validated by sequencing. Serum ferritin<br />

was compared in C282Y homozygotes with and without<br />

the GNPAT variant. The frequency of the GNPAT variant in<br />

referred patients was compared to a control population of voluntary<br />

blood donors without HFE mutations. Results: There were<br />

473 patients that had GNPAT analysis. The allele frequency for<br />

the GNPAT variant in C282Y homozygotes (n=81) was 0.302<br />

and in wild type control patients (n =392) was 0.213 (p = .08)<br />

Forty-nine percent (40) of the C282Y homozygotes were heterozygous<br />

(n=31) or homozygous (n = 9) for the GNPAT variant.<br />

The mean ferritin ± standard deviation did not significantly<br />

differ between C282Y homozygotes with (1450 μg/L ±1529)<br />

and without the GNPAT variant (1464 μg/L ± 1169). Conclusions:<br />

C282Y homozygotes referred for HFE testing commonly<br />

have a GNPAT variant. This GNPAT variant does not appear<br />

be a co-modifying gene affecting expression of HFE related<br />

hemochromatosis in this population. Further functional <strong>studies</strong><br />

are in progress to determine the role of the GNPAT variant in<br />

cellular iron metabolism.<br />

GNPAT variant in patients<br />

Disclosures:<br />

The following authors have nothing to disclose: Alexander Levstik, Alan Stuart,<br />

Paul Adams<br />

2118<br />

Vitamin D deficiency promotes hepatocellular cancer<br />

through activation of WNT/TLR signaling with disruption<br />

of TGF-β/Smad3<br />

Ji-Hyun Shin 1 , Jian Chen 1 , Nina M. Muñoz 1 , Lior Katz 2 , Andrea<br />

C. Cortes 1 , Vivek Shukla 3 , Sangbae Kim 1 , H. Franklin Herlong 1 ,<br />

Keigo Machida 4 , Hidekazu Tsukamoto 4 , Kirti Shetty 5 , Asif Rashid 1 ,<br />

Wilma S Jogunoori 6 , Aiwu R. He 7 , Lynt B. Johnson 7 , Ju-Seog Lee 1 ,<br />

Jon White 6 , Lopa Mishra 1 ; 1 The University of Texas MD Anderson<br />

Cancer Center, Houston, TX; 2 Sheba Medical Center, Tel<br />

Hashomer, Israel; 3 NCI, Bethesda, MD; 4 University of Southern<br />

California School of Medicine, Los Angeles, CA; 5 Johns Hopkins<br />

University School of Medicine, Baltimore, MD; 6 Veterans Affairs<br />

Medical Center, Washington, DC; 7 Georgetown University, Washington,<br />

DC<br />

Disruption of the TGF-β pathway has been associated with<br />

liver tumorigenesis, conditions associated with low Vitamin D<br />

(VD) levels. We examined a potential chemo-preventive role<br />

of VD in hepatocellular cancer (HCC) in the context of TGF-β<br />

inactivation. Methods: (1) We analyzed transcriptome profiles<br />

of 488 HCC and screened for somatic mutations (n = 202) and<br />

expression of the TGF-β pathway and VD pathway molecules<br />

(n = 147) in HCC from The Cancer Genome Atlas (TCGA). (2)<br />

Expression of TGF-β and VD pathway molecules was also evaluated<br />

by immunohistochemistry in liver samples from patients<br />

with hepatitis C virus (HCV)-associated cirrhosis who received<br />

VD supplements. (3) We examined the effects of VD on HCC<br />

development through gene and protein expression profiles in<br />

diethylnitrosamine -treated WT, and in mutant mouse models in<br />

the TGF-β pathway. (4) We further validated the effects of VD<br />

and performed further functional analyses examining crosstalk<br />

between TGF-β, VD and key members of cell proliferation and<br />

inflammation that were altered with deprivation of VD. Results:<br />

(1) We observed somatic mutations and a strong correlation<br />

between VD-related genes and the TGF-β pathway in the TCGA<br />

genomic analysis. (2) We observed low levels of TGF-β path-


1240A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

way member expression profiles with activation of β-catenin<br />

in tissues from HCV cirrhotic patients not on VD supplementation,<br />

while those supplemented showed restoration of TGF-β<br />

pathway member expression profiles with lower expression<br />

of β-catenin. (3) Strikingly, we observed a threefold higher<br />

incidence of HCC in Smad3 mice deprived of VD compared<br />

to mutant and WT mice fed a high-VD diet. (4) RPPA analyses<br />

revealed an activation of β-catenin, Stat5A, and Bcl2-XL, with<br />

reduced levels of the tumor suppressor PDCD4 in Smad3 +/-<br />

mice deprived of VD diet. (5) VD treatment suppressed HCC<br />

cell proliferation and significantly reduced expression levels<br />

of TLR7 and TLR-mediated inflammation associated genes in<br />

Smad3 -/- MEFs. (6) Functional analyses revealed that loss of<br />

Smad3 and shRNA for VD receptor, induced nuclear localization<br />

and activation of β-catenin. (5) Luciferase and ChIP assays<br />

showed that TGF-β negatively regulates TLR7 transcriptional<br />

activity through Smad3 binding at Smad Binding Elements at<br />

the TLR7 promoter. Conclusions: We conclude that VD suppresses<br />

proliferation of hepatocellular cancer cells by restoring<br />

TGF-β signaling. VD deficiency promotes tumor growth in the<br />

context of Smad3 disruption potentially through regulation of<br />

TLR7 expression and β-catenin nuclear localization. VD could<br />

therefore be a strong candidate for hepatocellular cancer prevention<br />

in the context of aberrant Smad3 signaling.<br />

Disclosures:<br />

Hidekazu Tsukamoto - Consulting: Suntory Ltd.; Grant/Research Support: The<br />

Toray Co.<br />

The following authors have nothing to disclose: Ji-Hyun Shin, Jian Chen, Nina M.<br />

Muñoz, Lior Katz, Andrea C. Cortes, Vivek Shukla, Sangbae Kim, H. Franklin<br />

Herlong, Keigo Machida, Kirti Shetty, Asif Rashid, Wilma S Jogunoori, Aiwu R.<br />

He, Lynt B. Johnson, Ju-Seog Lee, Jon White, Lopa Mishra<br />

2119<br />

Chemokine CCL3/CCR5-recruited, Cytotoxic CD4 + T<br />

Cells Eradicate Hepatoma Cells in a Model of Anticancer<br />

Chemotherapy with High-Dose Cyclophosphamide<br />

Tatsushi Naito 1,2 , Tomohisa Baba 2 , Kazuyoshi Takeda 3 , Soichiro<br />

Sasaki 2 , Naofumi Mukaida 2 , Yasunari Nakamoto 1 ; 1 Second<br />

Department of Internal Medicine, University of Fukui, Fukui, Japan;<br />

2 Division of Molecular Bioregulation, Cancer Research Institute,<br />

Kanazawa University, Kanazawa, Japan; 3 Department of immunology,<br />

Juntendo University, School of Medicine, Tokyo, Japan<br />

BACKGROUND: The mortality rate of hepatocellular carcinoma<br />

(HCC) is high because of frequent recurrence due to<br />

the high carcinogenic potentials of background chronic liver<br />

inflammation and cirrhosis. Thus, novel therapeutic strategy<br />

is required including immune therapy. Today, accumulating<br />

evidences indicate the potential role of CD4 + T cells with direct<br />

cyototoxicity (CD4 + CTLs) in antitumor immunity. Correlation<br />

between decrease of CD4 + CTLs and high mortality is also<br />

reported in patients with HCC. However, the importance of<br />

CD4 + CTLs in patients treated with anticancer chemotherapy<br />

remains to be elucidated. Hence, we examined the role of<br />

CD4 + CTLs, in chemotherapy-induced eradication of hepatoma<br />

by means of an animal HCC model. METHODS and RESULTS:<br />

We inoculated a murine hepatoma cell line, BNL 1ME A.7R.1<br />

(BNL), subcutaneously into Balb/c mice and gave single intraperitoneal<br />

injection of 150 mg/kg cyclophosphamide (CTX)<br />

after tumor formation. CTX treatment eradicated tumors in wildtype<br />

(WT) mice, whereas none of the tumors disappeared in<br />

T cell lacking nude mice. Next, we depleted CD4 + or CD8 +<br />

cells by antibody to clarify which subset was involved in the<br />

efficacy of CTX. We found CD4 + but not CD8 + cell depletion<br />

abrogated CTX-induced tumor eradiation. Consistently, CTX<br />

treatment increased intratumoral CD4 + CTLs, which expressed<br />

cytolytic granule molecule, CD107a and granzyme B observed<br />

by flowcytometry and immunofluorescence analysis. To delineate<br />

the underlying mechanism of increase in CD4 + CTLs, congenic<br />

CD45.1 splenocytes were transferred to tumor-bearing<br />

CD45.2 mice one day after CTX treatment and tumors were<br />

analyzed with time by flowcytometry. Transferred CD4 + cells<br />

were recruited into tumors on day 3, and the recruited cells<br />

expressed CD107a without antigen presentation at draining<br />

lymph nodes and proliferation in tumor tissues. Moreover, CTX<br />

administration enhanced mRNA expression of CC chemokine<br />

CCL3 in tumor tissues, and CTX-mediated tumor regression<br />

was attenuated in mice deficient in CCR5 gene, the receptor<br />

for this chemokine. Consistently, CTX-induced accumulation of<br />

intratumoral CD107a-expressing CD4 + T cells was less in mice<br />

receiving CCR5-deficient mouse-derived splenocytes than those<br />

receiving WT mouse-derived splenocytes. CONCLUSIONS: CC<br />

chemokine CCL3/CCR5 plays an important role in intratumoral<br />

CD4 + CTL migration in the process of CTX-induced tumor eradication.<br />

The results suggest that the CC chemokine-dependent<br />

mechanism may be a plausible strategy to enhance the antitumor<br />

cytotoxicity in the treatment of HCC with chemotherapy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tatsushi Naito, Tomohisa Baba,<br />

Kazuyoshi Takeda, Soichiro Sasaki, Naofumi Mukaida, Yasunari Nakamoto<br />

2120<br />

IFN-α stimulates IFN-γ expression in type I NKT cells<br />

and enhances the inhibition of HCV replication in<br />

human hepatocyte chimeric mice<br />

Eisuke Miyaki 1 , Michio Imamura 1 , Nobuhiko Hiraga 1 , Takuro<br />

Uchida 1 , Hiromi Kan 1 , Masataka Tsuge 1 , Hiromi Abe 1 , C. Nelson<br />

Hayes 1 , Hiroshi Aikata 1 , Chise Tateno 2 , Kazuaki Chayama 1 ;<br />

1 Department of Gastroenterology and Metabolism, Applied Life<br />

Sciences, Institute of Biomedical & Health Sciences, Hiroshima<br />

University, Hiroshima, Japan; 2 PhoenixBio Co., Ltd., Higashihiroshima,<br />

Japan<br />

Background & Aims: Interferon (IFN) inhibits hepatitis C virus<br />

(HCV) replication through up-regulation of IFN-stimulated gene<br />

(ISG) expression. However, the effect of IFN on immune cells is<br />

not well known. In this study, we analyzed the immune cell-mediated<br />

antiviral effects of IFN-α using HCV-infected mice. Methods:<br />

Genotype 1b HCV-infected urokinase-type plasminogen<br />

activator (uPA)-severe combined immunodeficiency (SCID) mice<br />

with transplanted human hepatocytes were injected intra-peritoneally<br />

with 2 x 10 7 human peripheral blood mononuclear<br />

cells (PBMCs) obtained from a healthy volunteer. The mice then<br />

received daily intra-peritoneal injections with 1000 IU/g of<br />

IFN-α for seven days, during which mouse serum HCV RNA,<br />

human albumin and cytokine levels were analyzed. Flow cytometric<br />

analysis of liver infiltrating human immune cells and<br />

histopathological examination were performed. Results: Seven<br />

days of IFN-α treatment without human PBMC injection in mice<br />

reduced serum HCV RNA levels by 1.23 ± 0.15 log whereas<br />

in combination with PBMCs, HCV RNA levels were reduced<br />

by 2.5 ± 0.45 log (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1241A<br />

these mice, the anti-HCV effect and development of human<br />

mononuclear cell chimerism by human PBMCs and IFN-α treatment<br />

was completely negated, suggesting that IFN-γ produced<br />

by NKT cells is essential not only for its anti-viral effect, but also<br />

for the development of human lymphocyte chimerism in human<br />

PBMCs and IFN-α treated mice. Conclusion: IFN-α stimulates<br />

IFN-γ expression in type I NKT cells and enhances the inhibition<br />

of HCV replication. Type I NKT cells might represent a new<br />

therapeutic target for chronic hepatitis C patients.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

The following authors have nothing to disclose: Eisuke Miyaki, Michio Imamura,<br />

Nobuhiko Hiraga, Takuro Uchida, Hiromi Kan, Masataka Tsuge, Hiromi Abe, C.<br />

Nelson Hayes, Hiroshi Aikata, Chise Tateno<br />

2121<br />

Small specific-sized Hyaluronic Acid normalizes<br />

TLR4-mediated signaling in Kupffer cells after chronic<br />

ethanol exposure associated with M2 polarization<br />

Paramananda Saikia 1 , Katherine A. Pollard 1 , Megan R. McMullen<br />

1 , Laura E. Nagy 1,2 ; 1 Center For Liver Disease Research, Pathobiology,<br />

Lerner Research Institute, Cleveland Clinic, Cleveland,<br />

OH; 2 Molecular Medicine, Case Western Reserve University,<br />

Cleveland, OH<br />

Hyaluronan (HA) is the major glycosaminoglycan in the extracellular<br />

matrix and ubiquitously present in many tissues. While<br />

HA has been used as a biomarker for liver injury for decades,<br />

the contribution of HA to alcoholic liver disease is not well<br />

understood. During acute and chronic inflammation or tissue<br />

injury, reactive oxygen species and matrix metalloproteinases<br />

increase HA turnover, resulting in local and systemic accumulation<br />

of HA fragments of different molecular weights. HA<br />

is recognized as an important damage associated pattern<br />

molecule (DAMP) that regulates innate immunity. However,<br />

depending on their size, HA fragments can have either pro-inflammatory<br />

or anti-inflammatory functions. Chronic alcohol<br />

consumption is associated with an increase in the sensitivity of<br />

hepatic macrophages to signaling via TLR4 associated with an<br />

M1 polarization and enhanced expression of pro-inflammatory<br />

cytokines. Here we investigated the impact of specific-sized<br />

small HA on TLR4 mediated signaling in Kupffer cells after<br />

chronic ethanol exposure to rats. Primary cultures of rat Kupffer<br />

cells were isolated from rats chronically exposed to ethanol<br />

via the Lieber-DeCarli diet or pair-fed control diets. Kupffer<br />

cells from ethanol-fed rats were more sensitive to stimulation<br />

with LPS, resulting in increased expression of mRNA for the<br />

pro-inflammatory cytokine, TNFα. When Kupffer cells were<br />

pre-treated with HA fragments for 5 hr prior to LPS challenge<br />

for 60 min, specific small-sized HA normalized TNFα mRNA<br />

expression in Kupffer cells from ethanol fed rats. In contrast,<br />

higher molecular weight HA fragments had no effect on TNF-α<br />

production of Kupffer cells. Normalization of TLR4 signaling<br />

after treatment with small-sized HA was associated with an<br />

increase expression of the M2 phenotypic gene arginase-1<br />

(arg-1) and a decrease expression of the M1 associated gene<br />

iNOS. In summary, these data demonstrate for the first time<br />

that specific small-sized HA fragments promote polarization<br />

of Kupffer cells to M2/anti-inflammatory phenotype to normalize<br />

chronic ethanol induced sensitization of TLR4 signaling in<br />

Kupffer cells.<br />

Disclosures:<br />

The following authors have nothing to disclose: Paramananda Saikia, Katherine<br />

A. Pollard, Megan R. McMullen, Laura E. Nagy<br />

2122<br />

Novel in vivo and in vitro models of Hepatitis E virus<br />

genotype 3 infectivity for chronic human HEV infection.<br />

Martijn D. van de Garde 1 , Suzan D. Pas 3 , Guido van der Net 3 ,<br />

Bart L. Haagmans 3 , Andre Boonstra 1 , Thomas Vanwolleghem 1,2 ;<br />

1 Department of Gastroenterology and Hepatology, Erasmus MC<br />

University Hospital, Rotterdam, Netherlands; 2 Gastroenterology<br />

and Hepatology, University Hospital Antwerp, Antwerp, Belgium;<br />

3 Viroscience, Erasmus Medical Center, Rotterdam, Netherlands<br />

Hepatitis E virus (HEV) genotype 3 (gt3) infections are emerging<br />

in western countries. The pathogenesis of HEV infection<br />

with associated liver pathology and clinical disease is poorly<br />

understood due to a lack of suitable model systems. Here we<br />

apply a novel in vivo and in vitro model system to characterize<br />

the infectivity of different HEV RNA containing human clinical<br />

samples. Human-liver chimeric mice (uPA +/+ Nod-SCID-IL2Rγ -/- )<br />

and human lung adenocarcinoma A549 cells were challenged<br />

with HEV gt3 obtained from human plasma, feces or liver-biopsy<br />

of 3 patients. Detection and quantification of HEV RNA<br />

was performed in mouse serum, feces, bile and liver or in<br />

A549 culture supernatant using RT-qPCR. High HEV RNA levels<br />

(up to 8 log IU HEV RNA/gr) were consistently detected in<br />

100% of chimeric mouse livers (n=10) from week 2-14 post<br />

inoculation with human feces-derived HEV (6-8 logs IU). Feces<br />

of these mice was positive for HEV RNA at least once during<br />

follow-up, while HEV viremia was inconsistently detectable,<br />

with maximum viral loads of 3.6 log IU/ml. HEV derived from<br />

a cryopreserved liver biopsy (5.5 log IU) similarly resulted in<br />

moderate to high HEV RNA levels in mouse feces, bile and<br />

liver (n=2). In contrast, anti-HEV IgG/IgM negative, HEV RNA<br />

positive plasma (6 log IU) was not infectious in any of the<br />

inoculated chimeric animals (n=8). In these animals human<br />

albumin plasma levels were stable and comparable to HEV-infected<br />

chimeric mice, indicating a functional hepatocyte graft.<br />

Infection of A549 cells with HEV RNA positive clinical samples<br />

showed rapid and increasing HEV RNA titers in culture<br />

supernatant within 7 days and could be maintained for more<br />

than 7 passages. Consistent in vitro HEV infectivity was seen<br />

with feces samples of all 3 patients, while only 2 of 3 human<br />

serum samples could be cultured in vitro, showing less efficient<br />

propagation. In conclusion, infectivity of feces derived human<br />

HEV is higher compared to serum derived HEV both in vitro<br />

and in vivo. The sustained HEV gt 3 liver infections induced in<br />

chimeric mice, show preferential viral shedding towards mouse<br />

bile and feces, mimicking the course of infection in humans.<br />

Besides antiviral efficacy <strong>studies</strong>, both models will guide future<br />

investigations into the biology of HEV infectivity with implications<br />

towards transfusion transmitted HEV in humans.<br />

Disclosures:<br />

Andre Boonstra - Grant/Research Support: BMS, Janssen Pharmaceutics, Merck,<br />

Roche, Gilead<br />

The following authors have nothing to disclose: Martijn D. van de Garde, Suzan<br />

D. Pas, Guido van der Net, Bart L. Haagmans, Thomas Vanwolleghem


1242A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2123<br />

Liver Plasmacytoid Dendritic Cells Expressing miR-181<br />

prolong graft survival by increasing T Regulatory cells<br />

and Decreasing B cells as Revealed by Mass Cytometry<br />

(CyTOF<br />

Audrey H. Lau 1 , Matthew J. Vitalone 2 , Xiumei Qu 2 , Todd Shawler 2 ,<br />

Olivia M. Martinez 2 , Carlos O. Esquivel 2 , Sheri M. Krams 2 ; 1 Pediatrics,<br />

Division of Gastroenterology, Hepatology, and Nutrition,<br />

Stanford/LPCH, San Francisco, CA; 2 Surgery, Stanford University,<br />

Stanford, CA<br />

Liver allografts are well tolerated and other solid organ<br />

allografts, when transplanted concurrently with livers, show<br />

improved outcomes. However, the mechanisms underlying<br />

“hepatic tolerance” have yet to be elucidated. Previous data<br />

show that hepatic dendritic cells (DC) have diminished antigen<br />

presentation and immune stimulatory function compared<br />

with DC in lymphoid tissue. Immature plasmacytoid (p)DC have<br />

been shown to induce graft prolongation and we have previously<br />

shown miR-181a1b1 is increased in pDC as compared<br />

to conventional DC. Wild-type (WT) BALB/c mice were transplanted<br />

with allogeneic (C57BL/6) vascularized heterotopic<br />

heart grafts. Animals received either no treatment (n=3) or were<br />

injected i.v. with 0.5×10 6 hepatic pDC (n=5) or miR-181a1b1 -/-<br />

(KO) pDC (n=3) at day (D)−7 from transplant. Recipients with<br />

no treatment rejected their allograft by D7 while recipients pretreated<br />

with hepatic pDC had significantly prolonged allograft<br />

survival to D15-21 (p1.5<br />

yrs) Opn -/- mice, which displayed significant hepatomegaly<br />

and splenomegaly compared to their age-matched WT littermates.<br />

Morphological features consisted of small aggregates<br />

of cells with intensely basophilic nuclei and clusters of immature<br />

and mature myeloid cells located in the sinusoids and in<br />

the portal areas. Accumulation of HSCs occurred also in the<br />

spleen but to a much lesser extent in the kidney and in the lung.<br />

Furthermore, Opn -/- mice showed premature exit of immature<br />

hematopoietic cells from the bone marrow to the spleen and<br />

to the liver, increased cellularity of lineage-negative cells and<br />

elevated hepatic GCSF, CXCL1, MCSF, MIP1α, RANTES and<br />

TNFα as well as serum MCSF and CXCL1 compared to WT<br />

littermates. Opn -/- mice also presented numerous iron deposits<br />

in the periportal zone of the liver, in the bone marrow and in<br />

the spleen but they were minimal in the lung and were absent<br />

in the heart and in the kidney, suggesting dysregulation of iron<br />

metabolism in the absence of Opn compared to WT littermates.<br />

CONCLUSION: These results suggest a critical role for OPN in<br />

HSC homeostasis by regulation of bone marrow homing and<br />

modulation of cytokine levels involved in their development.<br />

Disclosures:<br />

The following authors have nothing to disclose: Fernando Magdaleno, Xiadong<br />

Ge, Holger Fey, Costica Aloman, M. Isabel Fiel, Natalia Nieto<br />

2125<br />

IL-17A plays a pivotal role in LPS/GalN-induced fulminant<br />

hepatic injury in mice.<br />

Hiroshi Kono, Shinji Furuya, Chao Sun, Hideki Fujii; First Department<br />

of Surgry, University of Yamanashi, Chuo, Japan<br />

Background Lipopolysaccharide/D-Galactosamine (LPS/Gal-<br />

N)-induced hepatic injury is an experimental model of fulminant<br />

hepatic failure in which tumor necrosis factor-a (TNF-a) plays<br />

a pivotal role. Moreover, it was reported from our laboratory<br />

that interleukin (IL)-17A enhanced production of TNF-a by the<br />

Kupffer cell. Objective The purpose of this study was to determine<br />

the role of IL-17A in LPS/GalN-induced hepatic injury<br />

in mice. Methods LPS/GalN was injected into three mouse<br />

models: wild-type (WT) mice, IL-17A knockout (KO) mice, or<br />

IL-17A knockout mice treated with recombinant mouse (rm)<br />

IL-17A homodimer (KO + rmIL-17A). Survival was assessed<br />

for 24 hours after LPS/GalN injection, and histopathologic<br />

findings were evaluated at various time points after LPS/GalN<br />

injection for neutrophil and apoptosis markers. After LPS/GalN<br />

injection, expression of the inflammatory mediators TNF-a, IL-6,<br />

monocyte chemotactic protein (MCP)-1, IL-17A, high mobil-


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1243A<br />

ity group box 1 (HMGB1) and soluble intercellular adhesion<br />

molecule (sICAM)-1 was assessed in serum by enzyme-linked<br />

immunosorbent assay (ELISA). Results In the WT mice, mortality<br />

was 50 % at 24 hours after LPS/GalN injection. In contrast,<br />

all KO mice survived at 24 hours after injection. There were<br />

significant differences in the mortality between the WT mice<br />

and the KO mice. However, in KO + rmIL-17A mice, mortality<br />

was not significantly different compared to the other groups.<br />

Neutrophil infiltration and apoptosis were significantly greater<br />

in WT mice than KO mice. Furthermore, liver injury was severe<br />

in the WT mice compared with the KO mice; however, in KO<br />

+ rmIL-17A mice, injury was similar to those in the WT mice.<br />

Serum ALT levels were also significantly greater in WT mice<br />

than KO mice. In KO + rmIL-17A mice, these levels were similar<br />

to those in WT mice, as expected. Supporting the pathological<br />

findings, serum TNF-a, MCP-1, IL-17A, HMGB1 and sICAM-1<br />

levels were also significantly greater in WT mice than KO mice.<br />

In KO + rmIL-17A mice, these levels were similar to those in<br />

WT mice. Conclusions IL-17A is a key regulator in hepatic<br />

injury caused by neutrophil-induced inflammatory responses<br />

after LPS/GalN injection. The lack of IL-17A decreases the<br />

serum cytokine and chemokine levels, and all animals survived<br />

after LPS/GalN treatment. These results suggest that IL-17A<br />

may play one of the important role in fulminant hepatic injury.<br />

Disclosures:<br />

The following authors have nothing to disclose: Hiroshi Kono, Shinji Furuya,<br />

Chao Sun, Hideki Fujii<br />

2126<br />

Upregulated hepatic expression of pattern recognition<br />

receptors and Th1 type cytokine genes are associated<br />

with clearance of HEV in genotype 1 infected rhesus<br />

macaques<br />

Youkyung Choi, Xiugen Zhang, Brianna Skinner, Chong-Gee Teo;<br />

Center for Disease Control and Prevention, Atlanta, GA<br />

Infection by hepatitis E virus (HEV) genotype 1 causes acute<br />

liver disease that is usually self-limited but occasionally leads<br />

to fulminant hepatic failure. Immune responses, rather than<br />

virus-induced hepatocytolysis, have been shown to relate to<br />

the pathogenesis of acute hepatitis E and liver failure. Since<br />

HEV-specific CTL responses do not appear to be robustly<br />

induced in peripheral blood of HEV-infected patients during the<br />

acute and convalescent phases of infection, the liver is likely to<br />

be the major site of pathogenically relevant anti-HEV immune<br />

responses. In order to investigate the host immune responses<br />

induced by HEV infection, host gene expression profiles were<br />

analyzed in liver biopsy tissues of experimentally infected rhesus<br />

macaques. Three rhesus macaques were inoculated with<br />

HEV genotype 1 and observed for up to 150 days. All 3 exhibited<br />

the course typical of acute hepatitis. Thus, HEV RNA was<br />

detected in stools from 8 to 13 days after infection (DAI) and<br />

in serum from DAIs 22 to 32, becoming undetectable between<br />

DAIs 51 and 140; and elevated ALT activity and IgM and IgG<br />

anti-HEV antibody became detectable between DAIs 39 and<br />

56. Hepatic gene expression profiles were studied using 3 different<br />

real-time PCR based rhesus-macaque-specific arrays (Qiagen)<br />

that included 250 immune response-related genes. PCR<br />

arrays were done at 3 time-points: during and after the peak<br />

of stool HEV RNA detectability, and when HEV RNA became<br />

undetectable. Levels of gene expression were compared to<br />

normal liver tissues obtained from a macaque seronegative for<br />

anti-HEV. Genes showing ≥2-fold change in expression were<br />

considered to be significantly altered by HEV infection. Up to<br />

57 genes were up-regulated during and after the peak of stool<br />

HEV RNA detectability. These 57 genes could be grouped to<br />

belong to 3 broad categories: pattern recognition receptors<br />

(PRRs) (including TLR7 and TLR9); those associated with type 1<br />

interferon response (including ISG15, ISG20, IRF1, IRF5 and<br />

IRF7); and genes encoding Th1 type cytokines (including IL-12,<br />

CCR5, CXCR3, IFNg, STAT4 and TBX21). Conclusions: Genes<br />

upregulated in the liver during HEV infection include those<br />

associated not only with type 1 interferon response and Th1<br />

type cytokines but also PRRs. Upregulated hepatic PRR genes<br />

possibly mediate heightened interferon response and cytokine<br />

secretion, thereby contributing to resolution of acute infection.<br />

Studies are ongoing to study the kinetics, in plasma, of proteins<br />

associated with these responses.<br />

Disclosures:<br />

The following authors have nothing to disclose: Youkyung Choi, Xiugen Zhang,<br />

Brianna Skinner, Chong-Gee Teo<br />

2127<br />

Hepatocellular carcinoma is accelerated by modified<br />

western diet involving M2 macrophage polarization<br />

mediated by HIF-1a activation<br />

Aditya Ambade, Abhishek Satishchandran, Banishree Saha,<br />

Karen Kodys, Gyongyi Szabo; Medicine, UMass Medical School,<br />

Worcester, MA<br />

Background/Aims: Obesity is an independent risk factor for<br />

the development of hepatocellular carcinoma (HCC), the third<br />

leading cause of cancer related deaths worldwide. Tissue resident<br />

macrophages play a key role in promoting tumor progression.<br />

Transformed cells within the tumor tissue can induce<br />

macrophage polarization as a survival mechanism. However,<br />

the effect of western diet on macrophage polarization in tumor<br />

tissue microenvironment is not known. Here we sought to understand<br />

the role of western diet in promoting tumor development<br />

and inducing macrophage polarization after carcinogen<br />

pre-exposure. Methods: Four week-old male C57BL/6 mice<br />

were injected with 6 doses of N, N diethyl nitrosamine (DEN)<br />

intraperitoneally. 75 mg/kg/week DEN was administered for<br />

3 weeks followed by 100 mg/kg/week for 3 weeks. At 6<br />

weeks of age, mice were divided into modified western diet<br />

containing high fat-high cholesterol-high sugar (HF-HC-HS)<br />

and chow (control) diet groups. All mice were sacrificed at<br />

30 weeks of age; blood, liver, primary hepatocytes and liver<br />

mononuclear cells (LMNCs) were collected. In vitro, naïve primary<br />

hepatocytes were treated with 10 mM DEN and/or 330<br />

mM palmitic acid (PA) and supernatants were transferred to<br />

RAW 264.7 macrophages. Results: HF-HC-HS diet resulted<br />

in weight gain except for mice with DEN injection. ALT and<br />

liver to body weight ratio were significantly higher in HF-HC-<br />

HS+DEN compared to HF-HC-HS diet mice. Serum AFP levels<br />

were significantly higher in HF-HC-HS+DEN mice compared to<br />

all other groups (P


1244A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

hepatocyte supernatants induced HIF-1a DNA binding and<br />

up-regulated VEGF in RAW 264.7 macrophages. Conclusion:<br />

Modified western diet accelerated carcinogen initiated hepatocellular<br />

carcinoma, involving hepatocytes derived signals promoting<br />

M2 macrophage polarization.<br />

Disclosures:<br />

The following authors have nothing to disclose: Aditya Ambade, Abhishek Satishchandran,<br />

Banishree Saha, Karen Kodys, Gyongyi Szabo<br />

2128<br />

CD8 + T-cells drive liver injury and hepatocellular carcinoma<br />

in chronic liver disease<br />

Jessica Endig 1 , Laura E. Buitrago 1 , Silke Marhenke 1 , Anna Saborowski<br />

1 , jutta Schütt 2 , Florian Reisinger 3 , Florian Limbourg 1 ,<br />

Christian Koenecke 1 , Alina Schreder 1 , Robert Geffers 4 , Michael<br />

P. Manns 1 , Mathias Heikenwälder 5 , Thomas Longerich 6 , Arndt<br />

Vogel 1 ; 1 Hannover Medical School, Hannover, Germany;<br />

2 Philipps University Marburg, Marburg, Germany; 3 Technische<br />

Universität München, Munich, Germany; 4 Helmholtz Centre for<br />

Infection Research, Braunschweig, Germany; 5 Helmholtz Zentrum<br />

München, Munich, Germany; 6 Aachen University Hospital,<br />

Aachen, Germany<br />

Tumors frequently arise at sites of inflammation and chronic<br />

inflammation is increasingly recognized as a key factor in<br />

the pathogenesis of many malignancies. Hepatocellular carcinoma<br />

(HCC), one of the most lethal and prevalent cancers<br />

worldwide, represents a classic example of inflammation-linked<br />

cancer. Although activation of different cytokines has been<br />

reported in chronic liver disease, the critical components linking<br />

inflammation and hepatocarcinogenesis remain elusive. To<br />

characterize the role of the immune system in hepatic injury<br />

and tumor development, we comparatively studied the extent<br />

of liver disease and hepatocarcinogenesis in immunocompromised<br />

vs. immunocompetent Fah-deficient mice. Flares of liver<br />

injury were repeatedly induced which led to HCC formation<br />

within 3-4 months. Strikingly, chronic liver injury and tumor<br />

development were markedly suppressed in alymphoid Fah -/-<br />

mice despite an overall increased mortality. Mechanistically,<br />

we show that CD8 + Tcell-derived lymphotoxin-β (LTβ) is a<br />

central mediator of HCC formation in Fah -/- mice. Pharmacological<br />

inhibition of the lymphotoxin-β receptor (LTβR) indeed<br />

delays tumors development in mice with chronic liver injury<br />

without preventing liver damage. We here demonstrate that<br />

lymphocytes play a decisive, yet ambiguous role in chronic<br />

liver disease: while infiltrating lymphocytes mediate hepatocyte<br />

damage, liver fibrosis and tumor development, they also protect<br />

mice from acute on chronic liver failure and are required<br />

for activation of liver progenitor cells during regeneration. Our<br />

data illustrate that the immune system needs to be tightly regulated<br />

in a context-specific fashion, in order to balance immune<br />

surveillance and cancer risk. We propose that targeting specifically<br />

the tumor-promoting pathways such as LTβ might be an<br />

attractive chemopreventive strategy for patients at risk.<br />

Disclosures:<br />

Michael P. Manns - Consulting: Roche, BMS, Gilead, Boehringer Ingelheim,<br />

Novartis, Idenix, Achillion, GSK, Merck/MSD, Janssen, Medgenics; Grant/<br />

Research Support: Merck/MSD, Roche, Gilead, Novartis, Boehringer Ingelheim,<br />

BMS; Speaking and Teaching: Merck/MSD, Roche, BMS, Gilead, Janssen, GSK,<br />

Novartis<br />

The following authors have nothing to disclose: Jessica Endig, Laura E. Buitrago,<br />

Silke Marhenke, Anna Saborowski, jutta Schütt, Florian Reisinger, Florian<br />

Limbourg, Christian Koenecke, Alina Schreder, Robert Geffers, Mathias Heikenwälder,<br />

Thomas Longerich, Arndt Vogel<br />

2129<br />

Combined bone marrow plus liver transplantation in<br />

cynomolgus monkeys<br />

Joshua Weiner, Yojiro Kato, Sulemon Chaudhry, Raimon Duran-<br />

Struuck, Mercedes Martinez, Paula Alonso-Guallart, Jonah Zitsman,<br />

Jay H. Lefkowitch, Tomoaki Kato, Megan Sykes, Adam D.<br />

Griesemer; Columbia Center for Translational Immunology, New<br />

York Presbyterian-Columbia University Medical Center, New York,<br />

NY<br />

Introduction Combined kidney/bone marrow transplantation<br />

(CKBMT) induces tolerance in both primates and humans.<br />

Since liver is thought to be more tolerogenic than kidney, we<br />

hypothesized that combined liver/bone marrow transplantation<br />

(CLBMT) should be successful. We performed this pre-clinical<br />

study in cynomolgus monkeys. Methods Five pairs of animals<br />

were used. All were mismatched for class I, and the first 2 pairs<br />

were mismatched for class II. Liver and ~3 x 10^8 donor bone<br />

marrow cells/kg were transplanted on day 0. The induction<br />

regimen consisted of 1.5Gy total body irradiation on day-6,<br />

ATGAM (Pfizer) (50mg/kg) on days -2, -1, and 0, rituximab<br />

(Genentech) (20mg/kg) on day -6 (Recipients 2-5, also day<br />

6 for Recipients 4 and 5), 7Gy thymic irradiation on day -1,<br />

splenectomy on day 0, and 28 days of either IM cyclosporine<br />

(Novartis) or continuous IV tacrolimus (Astellas) followed by<br />

4-week taper (tacrolimus only). Recipients 4 and 5 additionally<br />

received LoCD2 (Mass Biologics) and anti-CD40 mAb (Mass<br />

Biologics). Anti-donor MLR was assessed by plating recipient<br />

and irradiated donor cells (1 x 10^6 each) for 5 days and<br />

pulsing with tritiated thymidine for 24 hours. Results Recipients<br />

1-3 survived 42, 69, and 57 days with multilineage mixed<br />

chimerism (MC) up to days 36, 40, and 23 respectively. They<br />

had normal to increased anti-donor cellular responses, except<br />

Recipient 1, and severe histological cellular rejection. Flow<br />

cytometry showed high percentages of CD8 T effector memory<br />

cells (TEM) in the blood, graft, and draining node but not<br />

peripheral nodes. Therefore, we added costimulatory blockade<br />

(anti-CD40) and depletion of CD2+ memory cells (LoCD2b)<br />

for Recipients 4 and 5, and subsequently this cell population<br />

did not expand. In contrast to the first three recipients, the<br />

fourth recipient survived 61 days (final 3 weeks with negligible<br />

immunosuppression levels) and died of diarrhea/electrolyte<br />

abnormalities on day 61 with normal LFTs, no histological<br />

rejection or GVHD, significant MC, and decreased anti-donor<br />

cellular responses. Recipient 5 is now 40 days post-transplantation<br />

with 70-80% chimerism in all lineages and no signs<br />

of rejection despite weaning immunosuppression. Conclusion<br />

CLBMT induces transient MC, but rejection eventually occurs.<br />

High levels of CD8 T effector memory cells likely contribute.<br />

However, when additional costimulatory blockade and CD2<br />

depletion are given, these cells are significantly depleted, preventing<br />

rejection, prolonging MC, decreasing anti-donor cellular<br />

responses, and possibly permitting tolerance.<br />

Disclosures:<br />

Tomoaki Kato - Grant/Research Support: Novartis<br />

The following authors have nothing to disclose: Joshua Weiner, Yojiro Kato,<br />

Sulemon Chaudhry, Raimon Duran-Struuck, Mercedes Martinez, Paula Alonso-Guallart,<br />

Jonah Zitsman, Jay H. Lefkowitch, Megan Sykes, Adam D. Griesemer


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1245A<br />

2130<br />

Allogeneic and xenogeneic transplantation of primary<br />

hepatocytes into the fetal liver in mice<br />

Yoshinobu Saito, Hayato Hikita, Yasutoshi Nozaki, Yugo Kai, Yuki<br />

Makino, Tasuku Nakabori, Satoshi Tanaka, Satoshi Aono, Ryotaro<br />

Sakamori, Naoki Hiramatsu, Tomohide Tatsumi, Tetsuo Takehara;<br />

Gastroenterology and Hepatology, Osaka University Graduate<br />

School of Medicine, Suita, Japan<br />

Background and Aim: Liver humanized mouse models using<br />

Alb-uPA SCID mice and Alb-HSVtk NOG mice have been<br />

reported to provide a lot of information on human liver physiology<br />

and pathology. However, these mouse models are not<br />

appropriate for investigating chronic inflammation and liver<br />

cancer because these mice are immune-deficient and could not<br />

survive more than one year. Immune tolerance is considered<br />

to be established during embryonic or neonatal stage. In this<br />

study, we investigated the engraftment efficacy of allogeneic or<br />

xenogeneic hepatocytes into the fetal liver by in utero transplantation<br />

in immune-competent mice. Method: As immune-competent<br />

recipient mice, we used hepatocyte-specific Mcl-1 knockout<br />

(KO) mice, which cause spontaneous hepatocyte apoptosis<br />

in life (Liver Injury mice), and hepatocyte-specific Mcl-1 and<br />

Bcl-xL double KO (DKO) mice, which show a decreased number<br />

of hepatocytes on embryonic (E) 18.5 day and die within<br />

one day after birth due to hepatic failure (Liver Impairment<br />

mice). Primary hepatocytes isolated from ubiquitously green fluorescent<br />

protein (GFP) transgenic (Tg) mice or primary human<br />

hepatocytes were administered into these recipient mice on E<br />

16.5 day via vitelline vein. These embryos were given birth by<br />

cesarean section and these livers were investigated at birth or<br />

at the age of 6 weeks. Results: Hepatocyte-transplanted wildtype<br />

mice and Liver Injury mice showed scattered embolization<br />

areas at birth by HE staining of the liver sections. The embolization<br />

was considered to be caused by transplanted hepatocytes.<br />

In contrast, Liver Impairment mouse livers escaped from<br />

embolization. In livers of wild-type mice and Liver Injury mice<br />

transplanted with GFP Tg hepatocytes, GFP-positive cells were<br />

observed forming a lot of clusters. While repopulation rate was<br />

not different between livers of wild-type mice and those of Liver<br />

Injury mice at birth (35.8% (3.6-48.3) and 29.8% (20.4-39.1),<br />

respectively), that of Liver Impairment was 7.8% (3.3-16.7).<br />

Meanwhile, repopulation rate at the age of 6 weeks was 0% in<br />

wild-type mice and 4.2% (2.1-8.0) in Liver Injury mice. In livers<br />

of wild-type and Liver Injury mice transplanted with primary<br />

human hepatocytes, embolization was similarly observed at<br />

birth by HE staining, suggesting that transplanted hepatocytes<br />

might be engrafted. Conclusion: Our results suggests that transplantation<br />

of primary hepatocytes into the fetal liver of Liver<br />

Injury mice, but not Liver Impairment mice, have the possibility<br />

of generating humanized liver mice without immunodeficiency.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yoshinobu Saito, Yasutoshi<br />

Nozaki, Yugo Kai, Yuki Makino, Tasuku Nakabori, Satoshi Tanaka, Satoshi<br />

Aono, Ryotaro Sakamori, Naoki Hiramatsu, Tomohide Tatsumi<br />

2131<br />

The gut-liver axis coordinate T-cell-mediated immunity<br />

by regulating the antigen-presenting activity of macrophages<br />

in murine liver<br />

Nobuhito Taniki 1 , Nobuhiro Nakamoto 1 , Hirotoshi Ebinuma 1 ,<br />

Po-sung Chu 1 , Takeru Amiya 1 , Shunsuke Shiba 1 , Akihiro Yamaguchi<br />

1 , Yuko Wakayama 1 , Hidetsugu Saito 2 , Takanori Kanai 1 ;<br />

1 Internal Medicine, Keio University, School Of Medicine, Tokyo,<br />

Japan; 2 Pharmaceutical Department, Keio University, Tokyo, Japan<br />

BACKGROUND & AIMS: The liver is a barrier for gut-derived<br />

antigens from degraded commensal bacteria. In colitic mice,<br />

live bacteria escaping the gut enter the liver through the portal<br />

vein. Specific subsets of innate immune cells suppress immune<br />

responses to such gut-derived antigens. The hygiene hypothesis<br />

proposes the recent increase in autoimmune hepatitis is the<br />

result of modern highly hygienic lifestyles. Reduced antigen<br />

exposure may cause dysfunctional hepatic immunosuppressive<br />

mechanisms. We hypothesized that increased antigen stimuli<br />

might activate immunosuppression. We investigated the role of<br />

innate immune cells in immunological tolerance in murine liver<br />

and the underlying mechanisms, focusing on the gut-liver axis.<br />

METHODS: We reported that dextran sulfate sodium (DSS)-<br />

treated mice (a colitis model) developed mild liver inflammation<br />

by histology and serology compared with normal mice following<br />

a sub-lethal dose of concanavalin A (ConA) (a T-cell-mediated<br />

hepatitis model). Colitis severity was inversely correlated<br />

with subsequent liver inflammation. We analyzed immune cell<br />

subsets in organs from experimental groups (WATER pre-ConA,<br />

WATER-ConA, DSS-pre ConA, and DSS-ConA). RESULTS: In<br />

DSS-ConA mouse liver, numbers of CD4 + and CD8 + T cells<br />

were decreased significantly, compared with the WATER-ConA<br />

group, but numbers of Foxp3 + CD4 + regulatory T cells (Tregs)<br />

were unchanged. Importantly, gene expression of the inflammatory<br />

cytokines interferon-gamma (IFN-γ) and tumor necrosis<br />

factor (TNF) in sorted T cells were significantly decreased in the<br />

DSS-ConA group. We recently reported that TNF-producing<br />

CCR9 + CD11b + macrophages were critical in the pathogenesis<br />

of ConA-induced acute liver injury by inducing T helper<br />

1 responses. In WATER-ConA-treated livers, CCR9 + CD11b +<br />

macrophages (WATER-ConA mac) were increased and CCR9 -<br />

CD11b + macrophages (DSS-ConA mac), a functionally distinct<br />

subset from WATER-ConA mac, were increased in DSS-ConAtreated<br />

livers. Sorted DSS-ConA mac had regulatory characteristics<br />

and produced IL-10, but low levels of TNF or IFN-γ after<br />

LPS stimulation in vitro. DSS-ConA mac expressed low MHC<br />

class II and CD80, and antigen presentation to naïve CD4 T<br />

cells derived from ovalbumin-specific αβ-TCR transgenic mice<br />

was deficient. CONCLUSIONS: CCR9 - CD11b + macrophages<br />

migrate to the inflamed liver under a colitic state and contribute<br />

to immunological tolerance in the liver by suppressing T-cell<br />

immunity. Our data provide a novel mechanism underlying<br />

immune tolerance in the liver through the gut-liver axis.<br />

Disclosures:<br />

Takanori Kanai - Grant/Research Support: Mitsubishi Tanabe Pharma Corporation<br />

The following authors have nothing to disclose: Nobuhito Taniki, Nobuhiro Nakamoto,<br />

Hirotoshi Ebinuma, Po-sung Chu, Takeru Amiya, Shunsuke Shiba, Akihiro<br />

Yamaguchi, Yuko Wakayama, Hidetsugu Saito


1246A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2132<br />

Small-specific-sized hyaluronic acid prevents inflammation<br />

and liver injury in chronic ethanol-fed mice<br />

Damien Bellos 1 , Megan R. McMullen 1 , Rebecca L. McCullough 1 ,<br />

Paramananda Saikia 1 , Sanjoy Roychowdhury 1 , Carol A. de la<br />

Motte 1 , Laura E. Nagy 1,2 ; 1 Lerner Research Institute, Cleveland<br />

Clinic, Cleveland, OH; 2 Center for Liver Disease Research, Case<br />

Western Reserve University, Cleveland, OH<br />

Chronic alcohol consumption leads to the activation of Kupffer<br />

cells and remodeling of the extracellular matrix to release hyaluronic<br />

acid fragments. Classically HA fragments are thought<br />

to act as DAMPs that promote inflammation; however, growing<br />

evidence indicates that small-specific-sized HA (SSS-HA)<br />

downregulate inflammatory signaling. We find that SSS-HA<br />

normalizes TNFα expression in primary Kupffer cells isolated<br />

from ethanol-fed rats. To investigate these findings in mice,<br />

female C57BL/6 mice were fed a Lieber DeCarli liquid diet<br />

with increasing concentrations of ethanol or pair-fed isocaloric<br />

maltodextrose over 25 days. SSS-HA was included or<br />

not in the liquid diet at 300ug/mouse/day. SSS-HA protected<br />

from ethanol-induced from increases in plasma ALT, AST and<br />

accumulation of hepatic triglycerides. SSS-HA also normalized<br />

lipid peroxidation and cell death in the liver as determined<br />

by 4-HNE and TUNEL staining, respectively. This decrease in<br />

liver injury was paralleled by lower expression of mRNA for<br />

inflammatory genes such as TNFα and IL-6. Semi-quantification<br />

of hepatic TNFα by immunohistochemistry showed that SSS-HA<br />

also prevented the ethanol-induced increase in TNFα in parenchymal<br />

and non-parenchymal cells. Taken together, these findings<br />

indicate that SSS-HA prevents liver injury by preventing the<br />

ethanol-induced increase in inflammatory signaling. Treatment<br />

with SSS-HA represents a potential pharmacologic intervention<br />

that could be targeted to normalize liver inflammation in ALD<br />

patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Damien Bellos, Megan R. McMullen,<br />

Rebecca L. McCullough, Paramananda Saikia, Sanjoy Roychowdhury, Carol<br />

A. de la Motte, Laura E. Nagy<br />

2133<br />

Sparstolinin B, a TLR4-antagonist Attenuates Early Steatohepatitic<br />

Injury in Progressive NonAlcoholic Steatohepatitis<br />

Diptadip Dattaroy 1 , Ratanesh K. Seth 1 , Suvarthi Das 1 , Firas Alhasson<br />

1 , Daping Fan 2 , Mitzi Nagarkatti 3 , Prakash Nagarkatti 3 , Saurabh<br />

Chatterjee 1 ; 1 Environmental Health Sciences, University of<br />

South Carolina, Columbia, SC; 2 Department of Cell Biology and<br />

Anatomy, University of South Carolin aSchool of Medicine, Columbia,<br />

SC; 3 PMI, University of South Carolina Shool of Medicine,<br />

Columbia, SC<br />

Nonalcoholic Steatoheaptitis, a hepatic manifestation of metabolic<br />

syndrome arises from an underlying condition of obesity.<br />

It is estimated that a roughly 5-8% (500,000) of obese<br />

population in USA alone have NASH. Sparstolonin B (SsnB)<br />

is derived from Chinese herb Spaganium stoloniferum and is<br />

a potent bioactive compound which has already been characterized<br />

to act as Toll like receptor 4 (TLR4) antagonist. In this<br />

research we hypothesize that SsnB attenuates early steatohepatitic<br />

injury and inflammation by inhibiting TLR4-Flotilin colocalization,<br />

decreased miR21 expression and PTEN activation. To<br />

prove our hypothesis, we have used an early steatohepatitic<br />

injury model of diet induced obese (DIO) mice administered<br />

with bromodichloromethane (BDCM) for 1 week. A similar<br />

group of mice were administered with SsnB for 1 week along<br />

with BDCM (SsnB treated group). Quantitative real time PCR<br />

(qRT PCR) results reveled that there was a significant decrease<br />

in CD68 (macrophage activation marker), MCP1 (monocyte<br />

chemoattractant protein 1), IL1β, TNFα, IL23 (inflammatory<br />

cytokines) gene expression in SsnB treated mice liver compared<br />

to BDCM treated group. SsnB treatment was found to decrease<br />

TLR4-flottilin colocalization compared to untreated group-thus<br />

attenuating TLR4 signaling. Our previous published research<br />

has shown that Micro RNA 21(miR21) upregulation causes<br />

inflammation and fibrosis in steatohepatitic injury by inhibiting<br />

its target proteins. Several other <strong>studies</strong> also show downregulation<br />

of PTEN and activation of PI3k/AKT pathway. Results in<br />

this model showed significantly decreased miR21 expression in<br />

SsnB treated mice liver compared to only BDCM treated mice.<br />

Increased PTEN activation is protective as it is a negative regulator<br />

of PI3K/AKT pathway and thus acts as a protective factor<br />

against steatohepatitic liver injury. Western blot data confirms<br />

significantly higher PTEN expression in SsnB treated mice liver<br />

as compared to BDCM treated mice. Serum ALT, a marker of<br />

liver injury was also decreased in SsnB treated mice as compared<br />

to BDCM-treated mice where liver injury was prominent.<br />

Taken together, we report a novel attenuation of steatohepatitic<br />

injury function of SsnB. Future <strong>studies</strong> will focus on molecular<br />

mechanisms in isolated cells to unravel the binding, kinetics<br />

and abrogation of TLR4 signaling in hepatic stellate cells.<br />

Disclosures:<br />

The following authors have nothing to disclose: Diptadip Dattaroy, Ratanesh<br />

K. Seth, Suvarthi Das, Firas Alhasson, Daping Fan, Mitzi Nagarkatti, Prakash<br />

Nagarkatti, Saurabh Chatterjee<br />

2134<br />

Macrophage activation through hypoxia inducible factor<br />

1 alpha (HIF1a) signaling is essential in parenteral<br />

nutrition associated cholestasis in mice<br />

Karim C. El Kasmi, Aimee Anderson, Michael W. Devereaux,<br />

Ronald J. Sokol; Pediatrics, UC Denver, Aurora, CO<br />

Background: Parenteral Nutrition Associated Cholestasis<br />

(PNAC) is most severe and progressive in children and adults<br />

with intestinal failure. We have previously reported that LPS<br />

absorption into the portal vein and transcription of pro-inflammatory<br />

cytokines (e.g., IL1β) is increased in hepatic macrophages<br />

in mice with intestinal injury that receive PN and<br />

develop PNAC. Furthermore, activation of macrophages and<br />

suppression of hepatic bile salt export pump Abcb11/BSEP<br />

and bilirubin exporter Abcc2/MRP2 required intact TLR4 signaling<br />

in this model. (Sci. Transl. Med. 2013 ;5:206ra137).<br />

An important role for the transcription factor HIF1a has<br />

been reported in pro-inflammatory macrophage activation in<br />

response to LPS-TLR4 signaling and that HIF1 directly increases<br />

transcription of IL-1b in response to LPS. We therefore hypothesized<br />

that HIF1α expression in macrophages played a critical<br />

role in the pathogenesis of PNAC. Methods and Results: PNAC<br />

model: Wild type C57/B6 mice were maintained on infusion<br />

of phytosterol-containing (soy lipid) PN solution through a central<br />

venous catheter for 14 days (DSS-PN mice) following exposure<br />

to dextran sulfate sodium (DSS) (to induce intestinal injury)<br />

in drinking water for 4 days. DSS-PN mice at 14 d developed<br />

cholestasis (increased serum bile acids and bilirubin),<br />

hepatocyte injury (increased AST and ALT), hepatic macrophage<br />

hyperplasia and hypertrophy by immunohistochemistry<br />

(F4/80 stain) and increased expression of IL1b mRNA in liver<br />

homogenate as well as in isolated intrahepatic macrophages.<br />

DSS-PN mice displayed significantly reduced hepatic mRNA<br />

for bile transporters Abcb11 and Abcc2 and the sterol transporter<br />

(Abcg5/g8). We next used genetic approaches to elucidate<br />

the roles of IL1b and HIF1a in this model. Global genetic


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1247A<br />

interruption of IL1b signaling (IL1R -/- mice) prevented PNAC<br />

in DSS-PN mice, and normalized hepatic macrophages on<br />

histology and the expression of bile and sterol transporters to<br />

levels observed in control mice. Macrophage specific genetic<br />

deletion of HIF1a (Hif1a-LysMcre) in DSS-PN treated mice also<br />

normalized AST, ALT, bile acids, and bilirubin, whereas similarly<br />

treated wild type mice showed evidence of hepatocyte<br />

injury and cholestasis. Moreover, hepatic gene expression of<br />

Abcb11, Abcc2, and Abcg5/g8 was not reduced in DSS-PN<br />

HIF1a-LysMcre mice. Conclusions: These data support a key<br />

role of macrophage activation through gut-derived LPS in triggering<br />

the development of PNAC. Moreover, these data are<br />

consistent with HIF1 as a critical upstream regulator of IL1b<br />

signaling in hepatic macrophages. Finally, blocking HIF1a<br />

signaling in macrophages is sufficient to prevent PNAC in a<br />

mouse model.<br />

Disclosures:<br />

Ronald J. Sokol - Advisory Committees or Review Panels: Yasoo Health, Inc.; Consulting:<br />

Roche, Ikaria, Otsuka American Pharmaceuticals, Alnylam, Retrophin;<br />

Grant/Research Support: Mead Johnson Nutritionals, Lumena, FFF Enterprises<br />

The following authors have nothing to disclose: Karim C. El Kasmi, Aimee Anderson,<br />

Michael W. Devereaux<br />

2135<br />

Knockdown of Receptor Interacting Protein kinase-1<br />

(RIPK1) markedly exacerbates immune-mediated liver<br />

injury and induces lethality through massive TNFα/<br />

caspase-dependent apoptosis of hepatocytes, independently<br />

of RIPK3/MLKL-mediated necroptosis.<br />

Jo Suda 1 , Lily Dara 1 , Luoluo Yang 1,3 , William A. Gaarde 2 , Neil<br />

Kaplowitz 1 , Zhang-Xu Liu 1 ; 1 Medicine, University of Southern<br />

California, Los Angeles, CA; 2 ISIS Pharmaceutics, Carlsbad, CA;<br />

3 Medicine, Bethune First Hospital of Jilin University, Changchuan,<br />

China<br />

Background: RIPK1 has an essential role in the signaling pathways<br />

triggered by death receptors (DR) through activation of<br />

NF-kB and regulation of caspase-dependent apoptosis and<br />

RIPK3/MLKL-mediated necroptosis. The pathophysiological<br />

roles of RIPK1 and RIPK3 are controversial in immune-mediated<br />

liver injury. Alpha-galactosylceramide (αGal), a specific activator<br />

for NKT cells, induces a well-established immune-mediated<br />

liver injury predominantly through DR pathways. Aim: examine<br />

the effect of RIPK1 knockdown on αGal-mediated liver injury.<br />

Methods: C57BL/6 mice received 4μg/mouse of αGal i.p. to<br />

induce immune-mediated liver injury. Injection of RIPK1 antisense<br />

(ASO), every other day, for 10 days efficiently silenced<br />

RIPK1 expression in liver. Results: αGal caused moderate liver<br />

injury in control ASO-treated mice characterized by small<br />

patchy inflammatory necrotic foci in histology and mild increase<br />

in serum ALT (190±110 at 6hr, n=10), with 100% survival.<br />

However, in RIPK1 ASO-treated (RIPK1-KD) mice, αGal caused<br />

severe liver injury evidenced by markedly increased serum ALT<br />

(12,403±3547 at 6hr, n=10, p


1248A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

TLR9 antagonists into DKO mice ameliorated serum ALT levels<br />

and the expression levels of IFN-beta. These results suggested<br />

that the activation of TLR9 was involved in the exacerbation<br />

of liver injury observed in DKO mice. Conclusion: Deficiency<br />

of DNase in apoptotic-prone hepatocytes with mitochondrial<br />

dysfunction is considered to induce necrotic cell death through<br />

the TL9 signaling pathway.<br />

Disclosures:<br />

Hayato Hikita - Grant/Research Support: Bristol-Myers Squibb<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yoshinobu Saito, Yasutoshi<br />

Nozaki, Yugo Kai, Yuki Makino, Tasuku Nakabori, Satoshi Tanaka, Satoshi<br />

Aono, Ryotaro Sakamori, Naoki Hiramatsu, Tomohide Tatsumi<br />

2137<br />

TRPV4 regulates inflammation and Kupffer cell activation<br />

in nonalcoholic steatohepatitis by attenuation of<br />

CYP2E1-mediated oxidative stress<br />

Ratanesh K. Seth 1 , Suvarthi Das 1 , Diptadip Dattaroy 1 , Firas Alhasson<br />

1 , Phillip D. Bell 2 , Gregory A. Michelotti 3 , Mitzi Nagarkatti 4 ,<br />

Prakash Nagarkatti 4 , Wolfgang B. Liedtke 5 , Anna Mae Diehl 3 ,<br />

Saurabh Chatterjee 1 ; 1 ENVIRONMENTAL HEALTH SCIENCES,<br />

UNIVERSITY OF SOUTH CAROLINA, Columbia, SC; 2 Department<br />

of Medicine, Medical University of South Carolina, Charleston,<br />

SC; 3 Division of Gastroenterology, Duke University, Durham, NC;<br />

4 Department of Pathology, Microbiology and Immunology, University<br />

of South Carolina School of Medicine, Columbia, SC; 5 Department<br />

of Neurology, Duke University Medical Center, Durham, NC<br />

Emerging evidence shows that oxidative stress via the activation<br />

of cytochrome p450 2E1 (CYP2E1) is key to progression<br />

of inflammation in nonalcoholic steatohepatitis (NASH). We<br />

have shown previously that CYP2E1 mediated oxidative stress<br />

and macrophage polarization in NASH livers were attenuated<br />

by administration of NO donor. However the molecular<br />

mediator and its pathways that regulate CYP2E1 mediated<br />

oxidative stress in NASH remains obscure. For this study we<br />

used a high fat diet induced obese mice (DIO), where CYP2E1<br />

substrate bromodichloromethane (BDCM) was used to induce<br />

CYP2E1 mediated oxidative stress, inflammation and NASH<br />

pathology. Results showed that the transient receptor potential<br />

vanilloid channel 4 (TRPV4) expression and protein levels were<br />

significantly elevated in parallel to increases in CYP2E1 and<br />

correlated well with increased lipid peroxidation, IL-1β, MCP-<br />

1, TNF-α and HMGB1 levels in the NASH livers and immortalized<br />

Kupffer cells. TRPV4 knockout (KO) mice exposed with<br />

BDCM showed increased CYP2E1 protein, lipid peroxidation,<br />

inflammatory cytokines, infiltration of leukocytes, endothelial<br />

dysfunction maker genes (CD34, cdh5, ICAM-1 and VEGFR2),<br />

HMGB1 levels, decreased endothelial nitric oxide synthase<br />

(NOS3) phosphorylation and exhibited early morbidity as<br />

compared to wild type mice with NASH. Mechanistically, diallyl<br />

sulfide (CYP2E1 inhibitor) or NO donor (DETANONOate)<br />

administration to TRPV4 KO mice completely abrogated<br />

enhanced NASH symptoms and morbidity. Interestingly, use<br />

of NO donor significantly decreased expression of CYP2E1,<br />

CYP2E1-mediated lipid peroxidation, an indirect measure of<br />

its activity, proinflammatory genes and endothelial dysfunction<br />

markers in TRPV4 KO mice. The results obtained show that<br />

TRPV4, a crucial protein responsible for sensing changes in<br />

osmotic pressure and Ca 2+ also regulates CYP2E1-mediated<br />

oxidative stress, inflammation and endothelial injury probably<br />

by activating NOS3 and release of nitric oxide. Based on the<br />

above, targeting TRPV4 or its downstream signaling cascade<br />

might be a promising therapeutic strategy in NASH.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ratanesh K. Seth, Suvarthi Das,<br />

Diptadip Dattaroy, Firas Alhasson, Phillip D. Bell, Gregory A. Michelotti, Mitzi<br />

Nagarkatti, Prakash Nagarkatti, Wolfgang B. Liedtke, Anna Mae Diehl, Saurabh<br />

Chatterjee<br />

2138<br />

Immune Regulation of Mesenchymal Stem Cells Is<br />

Dependent on Notch/Cox2/PGE2-Mediated Macrophage<br />

Differentiation in Liver Inflammatory Injury<br />

Changyong Li, Shi Yue, Ronald W. Busuttil, Jerzy Kupiec-Weglinski,<br />

Bibo Ke; The Dumont-UCLA Transplant Center, Department<br />

of Surgery, David Geffen School of Medicine at UCLA, Los Angeles,<br />

CA<br />

Background: Mesenchymal stem cells (MSCs) have been<br />

shown an immune modulatory properties and tissue repairing<br />

capability under inflammatory conditions. However, it is<br />

unknown how MSCs may affect immunity in liver inflammatory<br />

injury. This study was designed to dissect the molecular mechanisms<br />

of MSCs in regulating immune responses in ischemia and<br />

reperfusion (IR)-triggered liver inflammation. Methods: Bone<br />

marrow-derived MSCs were transfected with CRISPR/Cas9-mediated<br />

Notch1 or Cox2 knockout vector (p-Notch1 KO and<br />

p-Cox2 KO), and then co-cultured with bone marrow-derived<br />

macrophages using a transwell system followed by LPS (100<br />

ng/ml) stimulation. Using liver IRI model, mice (n=6/gr) were<br />

injected i.v. with 1x10 6 MSCs 24 h prior to liver ischemia. Mice<br />

were sacrificed after 6 h of reperfusion. Results: The expression<br />

of Notch1, Cox2, and PGE2 was enhanced in LPS-stimulated<br />

MSCs. Macrophages co-cultured with MSCs reduced M1 macrophage<br />

iNOS expression yet strongly augmented the expression<br />

of arginase-1, an M2 macrophage phenotype, which<br />

accompanied by increased expression of IL-10/TGF-β. However,<br />

transfection of p-Notch1 KO in MSCs decreased Cox2<br />

and PGE2 secretion, as well as reduced macrophage expression<br />

of arginase-1. Furthermore, knockdown of Cox2 by transfecting<br />

p-Cox2 KO in MSCs diminished PGE2 production and<br />

IL-10/TGF-β expression. Unlike in controls, adoptive transfer<br />

of MSCs in mice ameliorated IR-induced liver damage, which<br />

accompanied by increased expression of Notch1, Cox2, and<br />

PGE2 in ischemic livers. Furthermore, macrophages isolated<br />

from MSCs-treated mice displayed an increased expression of<br />

arginase-1 than those untreated controls. Conclusion: This study<br />

demonstrates that MSCs can reprogram host macrophage differentiation<br />

towards an anti-inflammatory M2 phenotype via a<br />

Notch/Cox2/PGE2-dependent manner. Our findings provide<br />

a novel therapeutic potential for the management of IR-triggered<br />

liver inflammatory injury.<br />

Disclosures:<br />

Ronald W. Busuttil - Grant/Research Support: Bayer, Fujisawa, Novartis, Roche/<br />

Genentech, NIH<br />

The following authors have nothing to disclose: Changyong Li, Shi Yue, Jerzy<br />

Kupiec-Weglinski, Bibo Ke


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1249A<br />

2139<br />

The attenuation of inflammatory liver injury by alkaline<br />

phosphatase is impaired in livers lacking asialoglycoprotein<br />

receptors: role of the major hepatic trafficking<br />

system in detoxification processes<br />

Robert S. Bridge 1 , John Gould 2 , Jacy L. Kubik 1 , Carol A. Casey 2,1 ,<br />

Dean J. Tuma 1 , Benita L. McVicker 2,1 ; 1 Internal Medicine, University<br />

of Nebraska Medical Center, Omaha, NE; 2 VA Nebraska-Western<br />

Iowa Health Care System, Omaha, NE<br />

Alkaline phosphatase (ALP) is a known biomarker during<br />

inflammatory liver diseases and associated lipopolysaccharide<br />

(LPS) toxicity. It has been noted that ALP expression during disease<br />

is a hepatoprotective response against LPS toxicity. Also,<br />

the therapeutic potential of ALP has been shown to involve LPS<br />

dephosphorylation and attenuation of liver injury following the<br />

administration of exogenous intestinal ALP (iALP). However,<br />

the mechanisms involved in ALP-mediated LPS detoxification<br />

are not defined. We hypothesize that the functional expression<br />

of a major hepatic trafficking protein, the asialoglycoprotein<br />

receptor (ASGPR), is a key factor involved in ALP regulation<br />

and LPS detoxification. It is known that ASGPRs bind, internalize<br />

and transport ALP in hepatocytes, yet the contribution<br />

of ASGPRs in ALP’s protective actions is not known. Here, the<br />

hepatoprotective potential of ALP was investigated in normal<br />

wild-type (WT) mice compared to ASGP receptor deficient (RD)<br />

animals subjected to LPS/D-galactosamine (Gal) liver injury<br />

with or without iALP treatment. Results: LPS/Gal administration<br />

resulted in enhanced liver injury in the RD animals shown<br />

by measures of apoptosis, serum transaminases, altered liver<br />

architecture and inflammatory parameters (6-fold increase<br />

(p


1250A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2141<br />

Inflammasome Activation Due to Vinyl Chloride Metabolite<br />

Exposure in NAFLD Caused by High Fat Diet in Mice.<br />

Lisanne C. Anders 1 , Adrienne M. Bushau 1 , Anna L. Lang 1 , Gavin<br />

E. Arteel 1 , Matthew C. Cave 2,1 , Craig J. McClain 2,1 , Juliane I.<br />

Beier 1 ; 1 Pharmacology and Toxicology, University of Louisville<br />

health Science Center, Louisville, KY; 2 Medicine, University of Louisville,<br />

Louisville, KY<br />

Background. Vinyl chloride (VC), a ubiquitous environmental<br />

contaminant, ranks 4 th on the ATSDR Hazardous Substances<br />

Priority List. A major paradigm shift in environmental research<br />

is to assess the impact of underlying disorders that may modify<br />

risk. A major health problem in the United States and worldwide<br />

is non-alcoholic fatty liver disease (NAFLD) due to dietary<br />

excess. NAFLD, the hepatic manifestation of metabolic complications<br />

due to obesity, may increase the sensitivity to other<br />

insults. Indeed, <strong>studies</strong> by our group and others suggest that<br />

obesity increases susceptibility to environmental hepatotoxicants<br />

(e.g., industrial solvents). Recent <strong>studies</strong> demonstrate a<br />

critical role of the inflammasome in macrophage activation<br />

during NAFLD. Inflammasome activation is induced by pathogen-associated<br />

molecular patterns (“PAMPs”), such as LPS,<br />

as well as by molecules released from dead or dying cells<br />

(damage-associated molecular patterns; “DAMPs”). Previously<br />

we have shown that VC metabolite chloroethanol (ClEtOH)<br />

exacerbated injury and inflammation leading to necrotic cell<br />

death in an experimental model of high-fat diet (HFD) induced<br />

NAFLD. The purpose of the current study was to investigate<br />

the interaction between NAFLD and VC metabolites in the context<br />

of inflammasome activation in an experimental model of<br />

HFD-induced obesity. Methods. Mice, fed a HFD (42% milk<br />

fat) or low fat control diet (LFD; 13% milk fat) for 10 weeks,<br />

were administered a bolus dose of ClEtOH or vehicle. Animals<br />

were sacrificed 0-24 hours after ClEtOH exposure. Plasma and<br />

tissue samples were harvested for determination of liver injury<br />

and inflammasome activation. Results. In LFD-fed control mice,<br />

ClEtOH caused no detectable liver damage, as determined by<br />

plasma parameters (AST and ALT) and histologic indices of<br />

damage. In HFD-fed mice, ClEtOH increased HFD-induced liver<br />

damage, steatosis, hepatocyte ballooning, infiltrating inflammatory<br />

cells, hepatic expression of proinflammatory cytokines<br />

and markers of endoplasmic reticulum (ER) stress. VC-metabolite<br />

induced cell death favors necrosis due to mitochondrial dysfunction<br />

and ATP depletion in this model. Moreover, in animals<br />

on a HFD, ClEtOH exacerbated expression of key markers<br />

involved in inflammasome activation, such as NLRP3 and IL-1β.<br />

Conclusions. Taken together, chloroethanol (as a surrogate VC<br />

exposure) can exacerbate liver injury and inflammasome activation<br />

in HFD-induced NAFLD. This serves as proof-of-concept<br />

that VC hepatotoxicity may be altered by risk-modifying factors<br />

such as diet-induced obesity and NAFLD. These data implicate<br />

exposure to VC as a risk factor in the development of liver disease<br />

in susceptible populations.<br />

Disclosures:<br />

Matthew C. Cave - Advisory Committees or Review Panels: Intercept, Abbvie;<br />

Consulting: Abbvie, Diapharma; Grant/Research Support: Merck, Gilead, Intercept,<br />

Conatus, Lumena, Cepheid, Tobira, Galectin, Bayer; Speaking and Teaching:<br />

BMS, Abbvie, Gilead, Janssen, Genentech<br />

Craig J. McClain - Consulting: Vertex, Gilead, Baxter, Celgene, Nestle, Danisco,<br />

Abbott, Genentech; Grant/Research Support: Ocera, Merck, Glaxo SmithKline;<br />

Speaking and Teaching: Roche<br />

The following authors have nothing to disclose: Lisanne C. Anders, Adrienne M.<br />

Bushau, Anna L. Lang, Gavin E. Arteel, Juliane I. Beier<br />

2142<br />

TNFα mediates the liver:lung axis in alcohol-enhanced<br />

acute lung injury in mice<br />

Lauren G. Poole 1 , Veronica L. Massey 1 , Edilson Torres-Gonzalez 2 ,<br />

Keith C. Falkner 2 , Deanna L. Siow 1 , Nikole L. Warner 3 , Robin H.<br />

Schmidt 1 , Jeffrey D. Ritzenthaler 2 , Jesse Roman 2 , Gavin E. Arteel 1 ;<br />

1 Pharmacology and Toxicology, University of Louisville, Louisville,<br />

KY; 2 Medicine, University of Louisville, Louisville, KY; 3 Microbiology<br />

and Immunology, University of Louisville, Louisville, KY<br />

Background: It is well known that liver and lung injury can<br />

occur simultaneously during severe inflammation (e.g. multiple<br />

organ failure). However, whether these are parallel or interdependent<br />

(i.e. liver:lung axis) mechanisms is unclear. Previous<br />

<strong>studies</strong> have shown that chronic ethanol exposure increases<br />

the incidence, severity, and mortality of sepsis-induced acute<br />

lung injury (ALI). There is a known liver:lung axis, and previous<br />

<strong>studies</strong> have suggested that hepatic cytokines can contribute to<br />

pulmonary inflammation; however, this hypothesis in the context<br />

of alcohol exposure has not been tested. Therefore, the purpose<br />

of the current study was to investigate the role of hepatic<br />

cytokine release in alcohol-enhanced lung injury. Methods:<br />

Male mice were exposed to ethanol-containing Lieber-DeCarli<br />

diet or pair-fed control diet for 6 weeks. Some animals were<br />

administered intraperitoneal lipopolysaccharide (LPS) 4 or 24<br />

hours prior to sacrifice to mimic sepsis-induced ALI. The effect<br />

of systemic TNFα depletion on lung injury was determined with<br />

a TNFα-inactivating antibody that is restricted to the plasma<br />

compartment (etanercept). The expression of cytokine mRNA<br />

in lung and liver tissue was determined by qPCR. Cytokine<br />

levels in the bronchoalveolar lavage fluid (BALF) and plasma<br />

were determined by Luminex assay. Results: The combination<br />

of ethanol and LPS caused enhanced liver injury, as indicated<br />

by significantly increased levels of the transaminases ALT/AST<br />

in the plasma and by changes in liver histology. In the lung,<br />

ethanol pre-exposure enhanced pulmonary inflammation and<br />

alveolar hemorrhage caused by LPS. As expected, ethanol<br />

enhanced LPS-induced TNFα expression in the liver; this effect<br />

of ethanol was not observed in the lung. In contrast, TNFα-dependent<br />

chemokines (MIP-2 and KC) were superinduced by the<br />

combination of ethanol and LPS. Systemic TNFα depletion with<br />

etanercept almost completely prevented the enhancement of<br />

lung damage. Furthermore, protecting against this damage correlated<br />

with an almost complete attenuation of the enhanced<br />

induction of TNFα-responsive chemokines MIP-2 and KC. Conclusions:<br />

Chronic ethanol pre-exposure enhanced both liver<br />

and lung injury caused by LPS. Enhanced organ injury corresponded<br />

with unique changes in the pro-inflammatory cytokine<br />

expression profiles in the liver and the lung. Although there<br />

are also likely pulmonary-specific changes caused by alcohol,<br />

these data suggest that systemic (hepatic) TNFα drives alcohol-enhanced<br />

ALI.<br />

Disclosures:<br />

Jesse Roman - Advisory Committees or Review Panels: Cellgene; Grant/Research<br />

Support: Actelion, Intermune, Novartis<br />

The following authors have nothing to disclose: Lauren G. Poole, Veronica L.<br />

Massey, Edilson Torres-Gonzalez, Keith C. Falkner, Deanna L. Siow, Nikole L.<br />

Warner, Robin H. Schmidt, Jeffrey D. Ritzenthaler, Gavin E. Arteel


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1251A<br />

2143<br />

Osteopontin promotes cholangiocyte chemokine secretion<br />

and macrophage accumulation in mice<br />

Jason D. Coombes 1 , Paul P. Manka 1,2 , Marzena Swiderska-Syn 3 ,<br />

Antonio Riva 4 , Danielle Reid 5 , Lee C. Claridge 6 , Laurent Dollé 7 ,<br />

Rasha Younis 1 , Marco A. Briones-Orta 1 , Naoto Kitamura 1 , Zhiyong<br />

Mi 8 , Paul C. Kuo 8 , Roger Williams 1,4 , Anna Mae Diehl 3 ,<br />

Shilpa Chokshi 4 , Bertus Eksteen 5 , Ali Canbay 2 , Wing-Kin Syn 1,8 ;<br />

1 Regeneration and Repair Group, Institute of Hepatology, Foundation<br />

for Liver Research, London, United Kingdom; 2 Department<br />

of Gastroenterology and Hepatology, Essen University Hospital,<br />

Essen, Germany; 3 Department of Medicine, Division of Gastroenterology,<br />

Duke University, Durham, NC; 4 Viral Hepatitis and<br />

Alcohol Research Group, The Institute of Hepatology, Foundation<br />

for Liver Research, London, United Kingdom; 5 Snyder Institute for<br />

Chronic Diseases, Health Research and Innovation Centre (HRIC),<br />

University of Calgary, Calgary, AB, Canada; 6 Department of<br />

Hepatology, Leeds Teaching Hospital NHS Trust, Leeds, United<br />

Kingdom; 7 Liver Cell Biology Lab (LIVR), Department of Cell Biology<br />

(CYTO),, Faculty of Medicine and Pharmacy, Vrije Universiteit<br />

Brussel, Brussels, Belgium; 8 Department of Surgery, Loyola University<br />

Chicago, Chicago, IL<br />

Background/Aim: Liver disease progression is characterized<br />

by recruitment and accumulation of inflammatory cells, which<br />

cluster around the peri-portal regions (i.e. ductular inflammatory<br />

response). We recently reported that repair-associated<br />

Hedgehog (Hh) ligands could induce cholangiocytes to secrete<br />

chemokines. The downstream Hh target Osteopontin (OPN)<br />

is widely upregulated during tissue injury and repair, and<br />

has been shown to regulate macrophage functions via pro-inflammatory<br />

chemoattractant properties. Although OPN is recognised<br />

as a key driver of liver fibrogenesis via activation of<br />

hepatic stellate cells and progenitor cells, the role of OPN in<br />

liver macrophage activation and recruitment remains unclear.<br />

We hypothesized that OPN stimulates cholangiocyte production<br />

of chemokines responsible for recruiting macrophage<br />

subsets that enhance liver fibrogenesis. Methods: In vivo: the<br />

role of OPN on total liver chemokine expression and macrophage<br />

numbers were assessed in two models of liver fibrosis<br />

(methionine-choline deficient diet and 3, 5,-diethoxycarbonyl-1,4-dihydrocollidine<br />

diet) by qRTPCR and/or FACS analysis<br />

(CD11b, Gr1, F4/80, CCR2) of liver infiltrating mononuclear<br />

cells. In vitro: OPN knockdown was achieved in murine 603B<br />

cholangiocytes using lentiviral mediated shRNA, and secreted<br />

OPN neutralized using specific aptamers. To determine if OPN<br />

modulates cholangiocyte chemokine production, conditioned<br />

media and cell lysates were harvested. Cytokine secretion was<br />

measured by cytometric bead array and mRNA by qRTPCR.<br />

The RAW264.7 cell line was used to study macrophage migration<br />

in a transwell assay. Results: At the end of treatment, MCD<br />

and DDC-fed mice developed liver fibrosis, upregulated total<br />

liver OPN, TGF-β, MCP-1, RANTES, and CXCL1 mRNA, and<br />

accumulated liver CD11b/F480(+) CCR2 (hi) macrophages.<br />

Conversely, MCD-fed mice that received OPN-aptamers (which<br />

neutralizes circulating OPN) downregulated MCP-1, RANTES,<br />

and CXCL1 mRNA, reduced liver CD11b/F4/80(+) CCR2 (hi)<br />

by 30%, and developed less fibrosis. In vitro, OPN knockdown<br />

reduced the secretion of chemokines RANTES, MCP-1<br />

and CXCL1 (by 75%, 73% and 41%, respectively). There were<br />

similar reductions in RANTES, MCP-1 and CXCL1 mRNA. Additionally,<br />

RAW264.7 macrophages cultured with conditioned<br />

media from 603B treated with OPN-neutralizing aptamer<br />

exhibited a 35% decrease in migration. Conclusions: OPN is<br />

overexpressed in progressive liver disease, enhances cholangiocyte<br />

production of key macrophage chemokines MCP-1,<br />

RANTES, and CXCL1, and promotes macrophage accumulation.<br />

OPN-neutralization leads to attenuated fibrogenic outcomes<br />

and is a promising anti-fibrotic strategy.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jason D. Coombes, Paul P.<br />

Manka, Marzena Swiderska-Syn, Antonio Riva, Danielle Reid, Lee C. Claridge,<br />

Laurent Dollé, Rasha Younis, Marco A. Briones-Orta, Naoto Kitamura, Zhiyong<br />

Mi, Paul C. Kuo, Roger Williams, Anna Mae Diehl, Shilpa Chokshi, Bertus<br />

Eksteen, Ali Canbay, Wing-Kin Syn<br />

2144<br />

Liver Enzymes Elevation In The Setting Of Chronic<br />

Graft-Versus-Host Disease Is A Non Specific Marker Of<br />

Inflammation That Does Not Accurately Predict Disease<br />

Related Hepatic Injury<br />

Ma Ai Thanda Han 4 , Niharika Samala 4 , Bisharah S. Rizvi 2 , Kenneth<br />

J. Wilkins 4 , Ohad Etzion 4 , Elizabeth C. Jones 5 , Christopher<br />

Koh 4 , David E. Kleiner 1 , Steven Pavletic 3 , Theo Heller 4 ; 1 LABORA-<br />

TORY OF PATHOLOGY, NCI, Bethesda, MD; 2 Internal Medicine,<br />

Canton medical education foundation, Canton, OH; 3 NIH-NCI,<br />

Bethesda, MD; 4 Liver Diseases Branch, NIH/NIDDK, Bethesda,<br />

MD; 5 Radiology, NIH-Clinical Center, Bethesda, MD<br />

Introduction: Diagnosis of chronic hepatic graft-versus-host disease<br />

(HcGVHD) is challenging, as the current scoring system<br />

(the NIH Scoring System or NSS) is based on total bilirubin<br />

(TB), alanine aminotransferase (ALT) and alkaline phosphatase<br />

(ALP). The challenge is that liver associated enzyme (LAE) and<br />

bilirubin elevations may be nonspecific. Once diagnosed with<br />

HcGVHD, patients are treated with immunosuppressive medications<br />

that have risk. Aims: To stratify patterns of LAE that could<br />

mimic HcGVHD, to assess for cytokine associations, and to<br />

assess the accuracy of the NSS for HcGVHD. Method: Patients<br />

with stable chronic GVHD (cGVHD) at any site were enrolled at<br />

variable times from bone marrow transplant. LAE, PT, TB, albumin,<br />

platelets, inflammatory markers (ESR, ferritin, C3), and<br />

cytokines were measured. Patients underwent liver biopsy at<br />

the physicians’ discretion. Box-Cox power transformation and t<br />

test were used for testing strength of associations. Results: 302<br />

patients (58% men, median age 43 y, ALT=39 IU/ml, ALP=96<br />

IU/ml, platelet=245 x 10 3 /ml) with cGVHD were included.<br />

.Abnormal ALT and AST were significantly associated with<br />

lower platelets, and higher TB and PT, in addition to higher ferritin,<br />

C3, MDC, TGFα, IFNα, IL15 and amyloid A (p= 0.0009,<br />

0.02, 0.007, 0.02, 0.04, 0.02, 0.04 respectively). Abnormal<br />

γ glutamyl transferase (GGT) was associated higher TB and<br />

PT but not with lower platelets. In addition, higher GGT was<br />

associated with elevated ESR (p=0.04) and C3 (p=0.0002),<br />

as well as with MDC, TGFα, IL 6, IP10 and IFNγ (p= 0.009,<br />

0.02, 0.008, 0.04, 0.02 respectively). Abnormal ALT, AST<br />

and GGT together were associated with abnormal platelet,<br />

TB, PT, C3 and ferritin (p=0.03, 0.01, 0.0003, 0.0007,<br />

0.02 respectively), as well as with MDC, and IL15 (p=0.008,<br />

0.01 respectively). 151/302 patients were diagnosed with<br />

HcGVHD by NSS. 24/302 patients underwent liver biopsy.<br />

Among them, 12 patients were diagnosed with HcGVHD by<br />

NSS of whom only 5 had histologic features that were consistent<br />

with HcGVHD, while others had characteristics of non-alcoholic<br />

steatohepatitis or nodular regenerative hyperplasia.<br />

Of the non-HcGVHD group by NSS, 4 had HcGVHD. Overall<br />

sensitivity and specificity for HcGVHD was 55% and 53%.<br />

There was no statistical differences between biopsy proven<br />

HcGVHD versus non-HcGVHD for AST, ALT, ALP, GGT, platelet,<br />

TB and albumin. Conclusion: Abnormal LAE in cGVHD is a<br />

marker of inflammation rather than a reflection of any specific<br />

pattern of liver injury. The NSS is neither sensitive nor specific<br />

for HcGVHD diagnosis. Hepatic histology may therefore be


1252A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

crucial to distinguish the causes of hepatic inflammation in<br />

cGVHD.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ma Ai Thanda Han, Niharika<br />

Samala, Bisharah S. Rizvi, Kenneth J. Wilkins, Ohad Etzion, Elizabeth C. Jones,<br />

Christopher Koh, David E. Kleiner, Steven Pavletic, Theo Heller<br />

2145<br />

The hepatic and extra-hepatic profile of resolution of<br />

steatohepatitis induced by GFT-505<br />

Arun J. Sanyal 2 , Stephen A. Harrison 3 , Sven M. Francque 4 , Pierre<br />

Bedossa 5 , Lawrence Serfaty 6 , Manuel Romero-Gomez 7 , Paul<br />

Cales 8 , Manal F. Abdelmalek 9 , Stephen H. Caldwell 10 , Joost<br />

Drenth 11 , Quentin M. Anstee 12 , Dean W. Hum 13 , Rémy Hanf 13 ,<br />

Alice Roudot 13 , Sophie Megnien 13 , Bart Staels 14 , Vlad Ratziu 1 ;<br />

1 Hepatology, Hopital Pitie Salpetriere, Paris, France; 2 Internal<br />

Medicine/Division of Gastroenterology, Hepatology and Nutrition,<br />

Virginia Commonwealth University, Richmond, VA; 3 Gastroenterology,<br />

Brooke Army Medical Center, Fort Sam Houston,<br />

TX; 4 Gastroenterology Hepatology, Antwerp Univesrity Hospital,<br />

Antwerp, Belgium; 5 Pathology, Hopital Beaujon, Clichy, France;<br />

6 Hepatology Department, Saint-Antoine Hospital, Paris, France;<br />

7 Digestive Diseases, Valme University Hospital, Sevilla, Spain;<br />

8 Hepatology Department, University Hospital & LUNAM University,<br />

Angers, France; 9 Medicine, Duke University, Durham, NC;<br />

10 Hepatology Department, University of Virginia, Charlotesville,<br />

VA; 11 Department of Gastroenterology and Hepatology, RUNMC,<br />

Nijmegen, Netherlands; 12 Institute of cellular Medicine, Newcastle<br />

University, Newcastle, United Kingdom; 13 Genfit SA, Loos,<br />

France; 14 INSERM U1011, European Genomic Institute for Diabetes<br />

(EGID), Institut Pasteur de Lille, Lille, France<br />

Background: Non-alcoholic steatohepatitis (NASH) is associated<br />

with increased liver- and cardio-metabolic risks. It is<br />

unknown whether resolution of steatohepatitis (SH) is associated<br />

with fibrosis regression and changes in cardio-metabolic<br />

risk profile, and whether drug-induced resolution of SH is similar<br />

to spontaneous resolution of SH. AIMS: We compared the<br />

hepatic and extra-hepatic profile associated with resolution<br />

of SH in GOLDEN505, a phase 2 randomized placebo-controlled<br />

trial of GFT505, a PPAR α/δ agonist. METHODS: Subjects<br />

with SH and NAS≥4 who received GFT505 120 mg/<br />

day (G120) or placebo (PBO) were studied. SH was defined<br />

by the presence of steatosis, inflammation and hepatocyte ballooning.<br />

Resolution of SH was defined by loss of any one of<br />

these components. Responders (R) were defined as resolution<br />

of SH without worsening of fibrosis. R and non-responders<br />

(NR) to G120 were compared for hepatic and extra-hepatic<br />

activity, and G120 R were compared to PBO R to explore the<br />

drug-induced effect of reversal of SH on extra-hepatic improvements.<br />

RESULTS: 202 subjects were included. The proportion of<br />

R was greater in G120 vs. PBO (22.4% vs. 12.7%; RR 2.01,<br />

p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1253A<br />

(advanced NASH: a-NASH)). 13 C-labeled palmitate (an SFA)<br />

was administered directly into the duodenum using gastrointestinal<br />

endoscopy to avoid delays resulting from delivery by<br />

stomach. Breath 13 CO 2<br />

levels were then measured to quantify<br />

metabolized SFA, and apolipoprotein B-48 concentrations in<br />

postprandial serum after test meal were measured to quantify<br />

absorbed chylomicrons in the intestine. Expression of SFA<br />

transporters in intestinal specimens from these groups was also<br />

examined. Results: Overall, 13 CO 2<br />

excretion was significantly<br />

higher in the e-NASH group than in the control and a-NASH<br />

groups (P


1254A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

centers (Hôpital Beaujon, Clichy and Hospital Clinic, Barcelona).<br />

Plasma samples were obtained from peripheral blood<br />

and M65, M30, CCL20 and TREM1 levels were measured by<br />

ELISA. The fraction of M65 and M30 carried by microparticles<br />

(MP) was determined. Plasma levels of biomarkers were compared<br />

between patients with and without histological diagnosis<br />

of AH. RESULTS AH was histologically confirmed in 47 of 88<br />

patients from the test cohort. The ABIC score was B or C in<br />

71/88 patients (81%), and 40/47 (85%) in those with AH.<br />

Patients with AH had higher median plasma levels of circulating<br />

free and MP-bound M65 (8.0 and 12.2-fold), free and<br />

MP-bound M30 (9.4 and 21.6-fold) and CCL20 (4.0-fold) than<br />

patients without AH (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1255A<br />

2150<br />

Cross-sectional and longitudinal agreement of magnetic<br />

resonance imaging proton density fat fraction with<br />

pathologist grading of hepatic steatosis in adults with<br />

nonalcoholic steatohepatitis in a multi-center trial<br />

Michael S. Middleton 1 , Elhamy Heba 1 , Catherine A. Hooker 1 ,<br />

Brent A. Neuschwander-Tetri 2 , Mustafa R. Bashir 3 , Kathryn<br />

Fowler 4 , Kumar E. Sandrasegaran 11 , Elizabeth M. Brunt 5 , David<br />

E. Kleiner 6 , Edward Doo 7 , Mark L. Van Natta 8 , James Tonascia 9 ,<br />

Rohit Loomba 10 , Claude B. Sirlin 1 ; 1 Radiology, UCSD, San Diego,<br />

CA; 2 Medicine, Saint Louis University, St. Louis, MO; 3 Radiology,<br />

Duke University, Durham, NC; 4 Radiology, Washington University,<br />

St. Louis, MO; 5 Pathology and Immunology, Washington University,<br />

St. Louis, MO; 6 Center for Cancer Research, NIH, Bethesda,<br />

MD; 7 Liver Disease Research Branch, NIH, Bethesda, MD; 8 Epidemiology,<br />

Johns Hopkins University, Baltimore, MD; 9 Biostatistics<br />

and Epidemiology, Johns Hopkins University, Baltimore, MD;<br />

10 Medicine, UCSD, San Diego, CA; 11 Radiology, Indiana University,<br />

Indianapolis, IN<br />

Materials and Methods Magnetic resonance imaging (MRI)<br />

was offered at baseline and end of treatment (EOT) to adults<br />

participating in the Farnesoid X Receptor Ligand Obeticholic<br />

Acid in NASH Treatment (FLINT) trial. Proton density fat fraction<br />

(PDFF), a non-invasive MRI biomarker for hepatic steatosis,<br />

was estimated using an advanced gradient-recalled-echo<br />

sequence that avoids or corrects confounding factors that can<br />

cause hepatic fat quantification to be inaccurate or scanner<br />

dependent. Imaging quality control was managed centrally<br />

by the NASH CRN Radiology Coordinating Center. Histologic<br />

steatosis grade, scored centrally by the NASH CRN Pathology<br />

Committee, served as the reference standard for PDFF.<br />

Cross-validated receiver operating characteristic (ROC) analyses<br />

were performed. Results Of 283 adults enrolled in FLINT,<br />

113 had MRI and biopsy at baseline, 85 had MRI and biopsy<br />

at EOT, and 78 had MRI and biopsy at both time points. MRIs<br />

were obtained at seven clinics using MRI scanners from three<br />

manufacturers, operating at one of two field strengths. Timing<br />

of MRIs ranged from 29 days before to 89 days after<br />

biopsy. At baseline, 34% of biopsies had steatosis grade 0<br />

or 1, 39% had grade 2, and 27% had grade 3 with corresponding<br />

mean(SD) PDFF(%) of 9.8(3.7), 18.1(4.3), and<br />

30.1(8.1), respectively. The area under ROC (AUROC) from<br />

logistic regression using PDFF as a surrogate for classifying<br />

steatosis grade 2+ vs. < 2 was 0.95 (95% CI: 0.91, 0.98) and<br />

for classifying steatosis grade 3 vs. < 3 was 0.96 (95 % CI:<br />

0.93, 0.99). PDFF cut-off values at 90% specificity were 16.3%<br />

for grade 2+ and 21.7% for grade 3 discrimination; sensitivities<br />

were 83% and 84%, respectively. At EOT, 42% of paired<br />

biopsies had improvement in steatosis grade, 49% had no<br />

change, and 9% had worsening with corresponding mean(SD)<br />

change from baseline in PDFF(%) of -7.4(8.7), 0.3(6.3) and<br />

7.7(6.0), respectively. The AUROC using PDFF change to classify<br />

steatosis grade improvement and worsening, respectively,<br />

were 0.81 (95% CI: 0.71, 0.91) and 0.81 (95% CI: 0.63,<br />

0.99). Cut-off values for PDFF change at 90% specificity were<br />

-5.1% for improvement and +5.6% for worsening; sensitivities<br />

were 58% and 57%, respectively. Conclusions In a multi-center<br />

setting, MRI PDFF showed high agreement with histologic<br />

hepatic steatosis grades on scanners at different sites using<br />

different scanner manufacturers and of different field strengths.<br />

Importantly, change in PDFF accurately classified change in<br />

72-week histologic hepatic steatosis grade. These findings support<br />

wider use of MRI PDFF in multi-center trials as a biomarker<br />

for hepatic steatosis at baseline, and for change in hepatic<br />

steatosis with treatment.<br />

Disclosures:<br />

Michael S. Middleton - Consulting: Bracco; Grant/Research Support: Gilead,<br />

Isis, Genzyme, Pfizer, Siemens, Bayer, Synageva, Merck, Janssen; Stock Shareholder:<br />

General Electric<br />

Brent A. Neuschwander-Tetri - Advisory Committees or Review Panels: Nimbus<br />

Therapeutics, Bristol Myers Squibb, Janssen, Mitsubishi Tanabe, Conatus,<br />

Scholar Rock<br />

Elizabeth M. Brunt - Consulting: Synageva; Independent Contractor: Rottapharm<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

Claude B. Sirlin - Grant/Research Support: GE, Pfizer, Bayer, Guerbet, Siemens<br />

The following authors have nothing to disclose: Elhamy Heba, Catherine A.<br />

Hooker, Mustafa R. Bashir, Kathryn Fowler, Kumar E. Sandrasegaran, David E.<br />

Kleiner, Edward Doo, Mark L. Van Natta, James Tonascia<br />

2151<br />

Non-invasive Diagnosis of Non-alcoholic Steatohepatitis<br />

(NASH) using a Panel of Fatty Acids<br />

Donjeta Gjuka 1 , Xiaoling Song 2 , Wonbeak Yoo 1 , Suk Young<br />

Yoo 3 , Jing Wang 3 , Michael B. Fallon 4 , George N. Ioannou 5 , Stephen<br />

A. Harrison 6 , Laura Beretta 1 ; 1 Molecular and Cellular Oncology,<br />

MD Anderson Cancer Center, Houston, TX; 2 Fred Hutchinson<br />

Cancer Research Center, Seattle, WA; 3 Bioinformatics and Computational<br />

Biology, The University of Texas MD Anderson Cancer<br />

Center, Houston, TX; 4 Division of Gastroenterology, The University<br />

of Texas Health Science Center at Houston, Houston, TX; 5 Division<br />

of Gastroenterology, Veterans Affairs Puget Sound Health Care<br />

System and University of Washington, Seattle, WA; 6 Department<br />

of Medicine, Brooke Army Medical Center, Houston, TX<br />

Introduction: Non-alcoholic fatty liver disease (NAFLD) is rapidly<br />

becoming the most common form of liver disease worldwide.<br />

NAFLD represents a spectrum of disease states ranging<br />

from isolated steatosis to non-alcoholic steatohepatitis (NASH).<br />

Diagnosis of NASH is important as this form of the disease<br />

has been shown to progress to cirrhosis and hepatocellular<br />

carcinoma. To date, liver biopsy is required for the diagnosis<br />

of NASH. Methods: We measured 46 fatty acids in sera from<br />

106 patients with NAFLD who underwent biopsy at Brooke Military<br />

Hospital in San Antonio (n=75) and the Veterans Affairs<br />

Puget Sound Health Care System in Seattle (n=31). For fatty<br />

acid profiling, total lipids were extracted from 100 μL of serum,<br />

the phospholipid fraction was isolated by one-dimensional<br />

thin-layer chromatography and methyl esters of phospholipid<br />

fatty acids were prepared using direct transesterification and<br />

separated by gas chromatography. Results: Levels of 15:0,<br />

17:0 and 16:1n7t negatively correlated with NAFLD activity<br />

scores (NAS) and hepatocyte ballooning scores, while levels<br />

of 18:1n7c strongly correlated with fibrosis scores and significantly<br />

discriminated patients with mild fibrosis and patients<br />

with intermediate or advanced fibrosis. Levels of 18:1n7c also<br />

correlated with liver inflammation. In addition, 15:0 and 17:0<br />

levels negatively correlated with fasting glucose and AST, while<br />

levels of 16:1n7c, 18:1n7c and 22:5n3 positively correlated<br />

with AST, ferritin and albumin, respectively. None of the six<br />

fatty acids correlated with BMI. Inclusion of age, ferritin, AST<br />

or AST-to-Platelet ratio Index (APRI) scores improved the performance<br />

of the fatty acid panels in detecting patients with NAS<br />

≥5 or patients with intermediate to severe fibrosis. The panel<br />

composed of 15:0, 16:1n7t, 18:1n7c, 22:5n3, age, ferritin<br />

and APRI best predicted intermediate or advanced fibrosis in<br />

NAFLD patients, with an area under ROC curve (AUROC) of<br />

0.92 and a 82% sensitivity at 90% specificity. Conclusion: The<br />

comprehensive analysis of fatty acid composition through the<br />

different stages of NAFLD and their correlation to histological


1256A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

hallmarks of NASH described in this study have allowed for<br />

the identification of 6 circulating fatty acids that could serve<br />

as non-invasive markers of disease severity in NAFLD patients.<br />

It also provides potential insights into mechanisms of disease<br />

progression. This panel should be further evaluated in prospective<br />

cohorts for its utility in distinguishing histological severity<br />

in patients with NAFLD. In addition, the biological effects of<br />

the fatty acids identified in this study should be investigated to<br />

evaluate their potential therapeutic utility.<br />

Disclosures:<br />

Michael B. Fallon - Grant/Research Support: Bayer/Onyx, Eaisi, Gilead, Grifolis<br />

Stephen A. Harrison - Advisory Committees or Review Panels: Merck, Nimbus<br />

Discovery, Fibrogen, RuiYi, CLDF; Consulting: NGM Biopharmaceuticals; Speaking<br />

and Teaching: Gilead, Abbvie, Janssen, CLDF<br />

The following authors have nothing to disclose: Donjeta Gjuka, Xiaoling Song,<br />

Wonbeak Yoo, Suk Young Yoo, Jing Wang, George N. Ioannou, Laura Beretta<br />

2152<br />

1 H-Magnetic Resonance Spectroscopy is superior to<br />

Controlled Attenuation Parameter (CAP) in assessing<br />

liver fat content in human non-alcoholic fatty liver disease<br />

(NAFLD)<br />

Jurgen H. Runge 1 , Loek Smits 4 , Joanne Verheij 3 , Aart Nederveen 1 ,<br />

Ulrich Beuers 2 , Jaap Stoker 1 ; 1 Radiology, Academic Medical Center,<br />

Amsterdam, Netherlands; 2 Gastroenterology and Hepatology,<br />

Academic Medical Center, Amsterdam, Netherlands; 3 Pathology,<br />

Academic Medical Center, Amsterdam, Netherlands; 4 Vascular<br />

Medicine, Academic Medical Center, Amsterdam, Netherlands<br />

PURPOSE Non-alcoholic fatty liver disease (NAFLD) is an<br />

increasingly recognized global health problem. Liver biopsy<br />

is the diagnostic standard, but given its drawbacks the liver<br />

fat content is preferably assessed noninvasively. The FibroScan®<br />

now provides the Controlled Attenuation Parameter<br />

(CAP) as a noninvasive measure of the liver fat content. However,<br />

only limited data are available regarding its accuracy<br />

and reproducibility compared to established and validated<br />

quantitative measures such as H-Magnetic Resonance Spectroscopy<br />

(H-MRS). Therefore, we prospectively compared<br />

CAP and H-MRS against liver biopsy in a cohort of NAFLD<br />

patients. PATIENTS AND METHODS Forty-four NAFLD patients<br />

(M/F: 33/11) with median (IQR) age of 52.3 (43.8–56.5)<br />

years and BMI of 27.2 (25.4–33.1) kg/m 2 were included in<br />

this prospective board-approved study. Same-day 3T MRI and<br />

CAP measurement were performed within 28 (17–50) days of<br />

biopsy, the latter assessed by a single hepatopathologist. Within-weeks<br />

and within-test reproducibility were assessed for CAP/<br />

MRI and CAP-only in a subgroup using Bland-Altman limits of<br />

agreement (BALA). H-MRS fat fractions (FF) and CAP values<br />

were compared between Brunt steatosis grades S0–S3 using<br />

Kruskall-Wallis and Mann-Whitney-U tests. Correlations were<br />

assessed with Spearman’s and diagnostic accuracies of CAP<br />

and FF to identify ≥S1 on biopsy with ROC analyses. RESULTS<br />

CAP showed considerable overlap between steatosis grades<br />

(see Table 1) and only differed between S0 and S2 (p=0.015)<br />

and S1 and S2 (p=0.004). FF differed (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1257A<br />

length appears to be an important factor impacting histological<br />

assessments and has important implications for the interpretation<br />

of placebo-controlled trials in NASH. Shorter biopsy<br />

length can increase placebo response in a trial and therefore,<br />

biopsy length should be closely monitored and factored into<br />

sample-size assessment.<br />

Disclosures:<br />

Manal F. Abdelmalek - Consulting: Islet Sciences; Grant/Research Support:<br />

Tobira, Gilead Sciences, NIH/NIDDK, Synageva, Genfit Pharmaceuticals,<br />

Immuron, Galmed, TaiwanJ Pharma, Intercept, NGM Pharmaceuticals<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

Brent A. Neuschwander-Tetri - Advisory Committees or Review Panels: Nimbus<br />

Therapeutics, Bristol Myers Squibb, Janssen, Mitsubishi Tanabe, Conatus,<br />

Scholar Rock<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

Reshma Shringarpure - Employment: Intercept<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

The following authors have nothing to disclose: Xiaohong Yan, Arthur J.<br />

McCullough<br />

2154<br />

Predictors of improvement in NAFLD Activity Score to<br />

obeticholic acid: A secondary analysis of FLINT Trial<br />

Rohit Loomba 1 , Arun J. Sanyal 2 , Kris V. Kowdley 3 , Norah Terrault<br />

4 , Naga P. Chalasani 5 , Manal F. Abdelmalek 6 , Arthur J.<br />

McCullough 7 , Xiaohong Yan 8 , Reshma Shringarpure 8 , Beatrice<br />

Ferguson 8 , David Shapiro 8 , Brent A. Neuschwander-Tetri 9 ; 1 University<br />

of California, San Diego, La Jolla, CA; 2 Virginia Commonwealth<br />

University, Richmond, VA; 3 Swedish Medical Center,<br />

Seattle, WA; 4 University of California, San Francisco, San Francisco,<br />

CA; 5 Indiana University, Indianapolis, IN; 6 Duke University,<br />

Durham, NC; 7 Cleveland Clinic, Cleveland, OH; 8 Intercept<br />

Pharmaceuticals, San Diego, CA; 9 Saint Louis University Health<br />

Sciences Center, St. Louis, MO<br />

Background: The Farnesoid X Receptor (FXR) Ligand Obeticholic<br />

Acid (OCA) in NASH Treatment (FLINT) Trial showed<br />

that OCA led to significantly higher rates of histologic response<br />

(defined as an improvement in NAFLD activity score [NAS] by<br />

≥2 with no worsening of fibrosis). Aim: The aims of this secondary<br />

analysis of the FLINT trial was to identify predictors of histologic<br />

response after 72 weeks of OCA treatment. Methods:<br />

Logistic regression model with stepwise selection procedure<br />

was performed to identify potential predictors of response after<br />

72 weeks of OCA treatment. Model selection was based on<br />

Akaike information criterion (AIC) and Bayesian information<br />

criterion (BIC). All patients included in analysis had biopsies<br />

at baseline (BL) and 72 weeks. Predictor data collected at BL,<br />

weeks 12, 24, and 36 were evaluated in this model. 59 predictors<br />

were assessed including BL values for demographics,<br />

histology, liver and serum biochemistry as well as the changes<br />

in those parameters over the course of the trial. The selected<br />

parameters were used to build a predictive model in OCAtreated<br />

patients for predicting response defined as improvement<br />

in the NAS by ≥2 with no worsening of fibrosis. The model was<br />

cross-validated by jackknife method, area under the receiver<br />

operating characteristic curve (AUROC), and 95% CIs. Results:<br />

Among 102 OCA-treated patients, 50 (49%) showed ≥2 point<br />

improvement in NAS without worsening fibrosis. This predictive<br />

model showed that older age, higher BL NAS, lower BL BMI,<br />

higher BL uric acid, and decrease in AST at week 24, increase<br />

in creatinine at week 24 and greater weight loss at week 24<br />

was associated with histologic response on OCA treatment<br />

(Table). The predictive model for placebo-treated patients was<br />

distinct from the OCA-treated patients (not shown). Conclusions:<br />

Baseline characteristics and changes in clinical parameters<br />

may be helpful in predicting histological response to OCA<br />

therapy in adults with NASH and further evaluation with larger<br />

patient numbers is warranted.<br />

Disclosures:<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Kris V. Kowdley - Advisory Committees or Review Panels: Achillion, BMS, Evidera,<br />

Gilead, Merck, Novartis, Trio Health, Abbvie; Grant/Research Support:<br />

Evidera, Gilead, Immuron, Intercept, Tobira; Speaking and Teaching: Abbvie,<br />

Gilead<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

Naga P. Chalasani - Consulting: Abbvie, Lilly, Celgene, Tobira, NuSirt, Takeda,<br />

Merck/Anthem, Salix; Grant/Research Support: Intercept, Gilead, Galectin<br />

Manal F. Abdelmalek - Consulting: Islet Sciences; Grant/Research Support:<br />

Tobira, Gilead Sciences, NIH/NIDDK, Synageva, Genfit Pharmaceuticals,<br />

Immuron, Galmed, TaiwanJ Pharma, Intercept, NGM Pharmaceuticals<br />

Reshma Shringarpure - Employment: Intercept<br />

Beatrice Ferguson - Employment: Intercept Pharmaceuticals<br />

David Shapiro - Employment: Inttercept Pharmaceuticals; Management Position:<br />

Intercept Pharmaceuticals; Stock Shareholder: Intercept Pharmaceuticals<br />

Brent A. Neuschwander-Tetri - Advisory Committees or Review Panels: Nimbus<br />

Therapeutics, Bristol Myers Squibb, Janssen, Mitsubishi Tanabe, Conatus,<br />

Scholar Rock<br />

The following authors have nothing to disclose: Arthur J. McCullough, Xiaohong<br />

Yan


1258A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2155<br />

Short Sleep Duration is associated with Nonalcoholic<br />

Fatty Liver Disease in US Adults<br />

Donghee Kim 1 , Hwa Jung Kim 2 , Alina M. Allen 3 , Aijaz Ahmed 1 ,<br />

Nae-Yun Heo 1 , Prowpanga Udompap 1 , Ajitha Mannalithara 1 , W.<br />

Ray Kim 1 ; 1 Division of Gastroenterology and Hepatology, Stanford<br />

University School of Medicine, Stanford, CA; 2 Department<br />

of Clinical Epidemiology and Biostatistics, Asan Medical Center,<br />

Seoul, Korea (the Republic of); 3 Division of Gastroenterology and<br />

Hepatology, Mayo Clinic, Rochester, MN<br />

Background/Aims: Short sleep duration has been associated<br />

with increased risk of hypertension, type 2 diabetes and obesity.<br />

We investigate association between duration and quality<br />

of sleep and nonalcoholic fatty liver disease (NAFLD). Methods:<br />

Among adult participants in the National Health and<br />

Nutrition Examination Survey in 2005-2012, subjects with<br />

a potential diagnosis of NAFLD were selected by excluding<br />

individuals with excessive alcohol consumption (> 30 g/day<br />

for men and 20 g/day for women) or with evidence of viral<br />

hepatitis. Suspected NAFLD (sNAFLD) was diagnosed if serum<br />

alanine aminotransferase was >30 U/L for men and >19 U/L<br />

for women in the absence of other causes. Predicted NAFLD<br />

(pNAFLD) was diagnosed using fatty liver index (FLI) of ≥ 60<br />

(based on body mass index (BMI), waist circumference, triglycerides<br />

and gamma glutamyl transpeptidase). Duration of<br />

sleep, categorized into ≤ 5, 6, 7, 8, and ≥ 9 hours per night<br />

and sleep quality, graded as good, moderate and poor, were<br />

assessed using the Sleep Disorders’ Questionnaire. Results:<br />

There were 17,245 participants (mean age 46 years, 48.7%<br />

male) represented in the data set. The most common duration<br />

of sleep was 7 hours (27.1%), followed by 8 hours (26.3%),<br />

6 hours (23.8%), ≤ 5 hour (15.9%), and ≥ 9 hour (7.0%).<br />

Respondents rated their quality of sleep to be good (78.0%),<br />

moderate (11.7%), or poor (10.3%). The prevalence of sNA-<br />

FLD was 37.3%, which was inversely correlated with sleep<br />

duration: 39.2% of respondents with sleep ≤ 5 hours had sNA-<br />

FLD, 38.0% with 6 hours, 37.3% with 7 hours, 36.6% with 8<br />

hours, and 33.3% with ≥ 9 hours of sleep (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1259A<br />

larger generalized population. Conclusions: The combination<br />

of serum Fuc-Hpt and Mac2bp can distinguish NASH from<br />

NAFLD patients. Our noninvasive model using serum two glycobiomarkers<br />

contribute to a novel NASH diagnostic methodology<br />

instead of liver biopsy.<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

Norifumi Kawada - Grant/Research Support: BMS, Chugai, Kowa; Speaking<br />

and Teaching: MSD, Janssen<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Yoshihiro Kamada, Masafumi<br />

Ono, Hideyuki Hyogo, Hideki Fujii, Yoshio Sumida, Kohjiroh Mori, Saiyu<br />

Tanaka, Makoto Yamada, Maaya Akita, Kayo Mizutani, Hironobu Fujii, Akiko<br />

Yamamoto, Shinji Takamatsu, Yuichi Yoshida, Toshiji Saibara, Eiji Miyoshi<br />

2157<br />

Increased Immune Responses of Patients with Lean<br />

Non-Alcoholic Fatty Liver Diseases Over Obese Non-Alcoholic<br />

Fatty Liver Diseases in Response to Saturated<br />

Fatty Acid<br />

Ayub Al Mamun 2 , Mamun A. Mahtab 2 , Sheikh Mohammad Fazle<br />

Akbar 1 ; 1 Department of Medical Sciences, Toshiba General Hospital,<br />

Tokyo, Japan; 2 Department of Hepatology, Bangbandhu<br />

Sheikh Mujib Medical University, Dhaka, Bangladesh<br />

Purpose: To develop insights about mechanism of pathogenesis<br />

of obesity in lean persons, this study was initiated to dissect<br />

immunological events in lean patients with non-alcoholic fatty<br />

liver diseases (NAFLD) versus obese patients with NAFLD. Methods:<br />

Out of total 450 patients with NAFLD, clinical parameters<br />

of age and sex-matched 50 patients with lean NAFLD (body<br />

mass index (body mass index, BMI) 30.0) were compared. Peripheral<br />

blood mononuclear cells (PBMC) and antigen-presenting dendritic<br />

cells (DC) from both lean NAFLD and obese NAFLD<br />

patients were isolated and production of pro-inflammatory cytokines<br />

and nitric oxide (NO) in response to various stimuli were<br />

evaluated. Results: As mentioned, the BMI of lean NAFLD patents<br />

were significantly lower than that of obese NAFLD patients<br />

(p0.05). The levels of proinflammatory cytokines [interferon<br />

(IFN)-gamma, tumor necrosis factor (TNF)-alpha, and interleukin<br />

(IL)-6] produced by PBMC and DC were also comparable<br />

between lean NAFLD and obese NAFLD patients when these<br />

were stimulated by polyclonal stimulators like concanavalin<br />

A or lipopolysachcharides (p>0.05). However, significantly<br />

higher levels of IFN-gamma, TNF-alpha, and IL-6 were produced<br />

by PBMC and DC of lean NAFLD patients compared to<br />

those produced by obese NAFLD patients in response to palmitic<br />

acid (p0.05).<br />

This outcome was reproducible in three separate experiments<br />

with sera collected at three separate points. Also, nitric oxide<br />

(NO) production was significantly higher from PBMC of lean<br />

NAFLD patients compared to obese NAFLD patients (p


1260A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

DM-HCC individuals than DM-non-HCC individuals (OR 3.44,<br />

p = 0.0002). The influence of PNPLA3 and JAZF1 was independent<br />

of age, sex, and body mass index after multivariable<br />

logistic regression analysis. There were no differences in the<br />

allele frequencies of PNPLA3 and JAZF1 between the HCV-<br />

HCC and HCV-non-HCC groups. Conclusions: The PNPLA3<br />

rs738409 was determined to be associated with HCC development<br />

in non-viral hepatitis patients with T2DM. Additionally,<br />

the JAZF1 rs864745 was identified as a risk factor for<br />

HCC among T2DM patients with the GG genotype at PNPLA3<br />

rs738409.<br />

Disclosures:<br />

The following authors have nothing to disclose: Misuzu Ueyama, Masaaki<br />

Korenaga, Nao Nishida, Keiko Korenaga, Erina Kumagai, Masaya Sugiyama,<br />

Yoshihiko Aoki, Masatoshi Imamura, Kazumoto Murata, Naohiko Masaki,<br />

Takumi Kawaguchi, Takuji Torimura, Hideyuki Hyogo, Hiroshi Aikata, Kiyoaki<br />

Ito, Yoshio Sumida, Tatsuya Kanto, Sumio Watanabe, Masashi Mizokami<br />

2159<br />

Older Age, Increasing Body Mass Index, and Concurrent<br />

Diabetes Mellitus Increases Risk of Advanced Fibrosis<br />

Among Nonalcoholic Fatty Liver Disease Patients<br />

Maria Aguilar 2 , Edward W. Holt 3 , Taft Bhuket 1 , Benny Liu 1 , Zobair<br />

M. Younossi 4,5 , Aijaz Ahmed 6 , Robert J. Wong 1 ; 1 Gastroenterology<br />

and Hepatology, Alameda Health System - Highland Hospital,<br />

Oakland, CA; 2 Medicine, Alameda Health System - Highland<br />

Hospital, Oakland, CA; 3 Transplantation - Division of Hepatology,<br />

California Pacific Medical Center, San Francisco, CA; 4 Medicine<br />

- Center for Liver Diseases, Inova Fairfax Hospital, Falls Church,<br />

VA; 5 Betty and Guy Beatty Center for Integrated Research, Inova<br />

Health System, Falls Church, VA; 6 Division of Gastroenterology<br />

and Hepatology, Stanford University School of Medicine, Stanford,<br />

CA<br />

Background:Nonalcoholic fatty liver disease (NAFLD) is the<br />

leading cause of chronic liver disease (CLD) in the U.S. and<br />

will become the leading etiology of hepatocellular carcinoma<br />

and liver transplantation. However, the true prevalence of<br />

NAFLD and NAFLD with advanced fibrosis (NAFLD-AF) is not<br />

well understood. Aim:To determine the current prevalence and<br />

predictors of NAFLD and NAFLD-AF among U.S. adults. Methods:Using<br />

the most recent 2011-2012 U.S. National Health<br />

and Nutrition Examination Survey, we defined NAFLD as presence<br />

of both abnormal alanine aminotransferase (ALT>20 U/L<br />

in women, ALT>30 U/L in men) and metabolic syndrome after<br />

excluding other CLD etiologies (eg. hepatitis C, hepatitis B,<br />

alcohol). NAFLD-AF was defined as NAFLD fibrosis score ><br />

0.676. Prevalence estimates were stratified by sex, ethnicity,<br />

and age. Multivariate logistic regression models evaluated predictors<br />

of NAFLD and NAFLD-AF. Results:Overall NAFLD prevalence<br />

was 21.1%, representing 49.2 million persons. NAFLD<br />

prevalence was higher in females compared to males (31.4%<br />

vs. 14.0%, p60: 48.2% vs. age


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1261A<br />

regression analysis of biomarkers indicated that HA, CK18<br />

and TIMP1 were independent predictor of advanced fibrosis<br />

(odds ratio [OR] HA=1.057, CK18=1.002, TIMP1=0.99,<br />

95% CI, p


1262A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

hypertension, fasting plasma insulin and vitamin B12. In contrast,<br />

the fibrosis score was higher by 1.4 units in the high<br />

folate group (RBC folate > 620 ng/ml) compared to low folate<br />

group (p=0.003) in patients with the homozygous group. The<br />

fibrosis score increases by 0.06 units for every 25mg increase<br />

in serum folate level in homozygous group (p=0.014). Plasma<br />

vitamin B12 was not associated with histological severity. Conclusions.<br />

While folate has a protective effect on ballooning and<br />

the overall NAS score in NAFLD, in those NAFLD patients with<br />

the MTHFR homozygous mutation high folate levels had more<br />

advanced fibrosis. Despite the reported decreased activity of<br />

mutated MTHFR, folate supplementation may contribute to progression<br />

of fibrosis and should be used with caution.<br />

Disclosures:<br />

The following authors have nothing to disclose: Jaividhya Dasarathy, Rony Varghese,<br />

Iryna Kalinina, Rocio Lopez, Arthur J. McCullough, Srinivasan Dasarathy<br />

2163<br />

The Lipidomic Signature Of Disease Progression In Nonalcoholic<br />

Fatty Liver Disease (NAFLD)<br />

Cristina Alonso 1 , Amon Asgharpour 2 , Ibon Martinez-Arranz 1 ,<br />

Puneet Puri 2 , Mohammad S. Siddiqui 2 , Pablo Ortiz 1 , Jose M.<br />

Mato 3 , Arun J. Sanyal 2 ; 1 OWL, Derio, Spain; 2 VCU Medical Center<br />

Richmond, Division of Gastroenterology and Hepatology, Richmond,<br />

VA; 3 CIC bioGUNE, Ciberehd, Derio, Spain<br />

BACKGROUND:NAFLD includes a spectrum of histological<br />

phenotypes which can progress to cirrhosis at variable rates.<br />

While fibrosis assessment with a liver biopsy is the gold standard<br />

for assessment of disease progression it is limited by<br />

sampling variability and its invasive nature. There is therefore<br />

a major unmet need to identify a signature of disease progression<br />

that does not rely on a liver biopsy. Metabolomics<br />

provides an unbiased approach to obtain an assessment of<br />

whole-body metabolic response to disease progression. Previous<br />

attempts to describe a lipid signature focused on fatty liver<br />

vs.NASH (Hepatol 2009,50:1827, J Prot Res 2012,11:2521,<br />

J Lipid Res 2015,56:185) or NASH vs.cirrhosis (J Lipid Res<br />

2015,56:722). The lipidomic signature of disease progression<br />

over earlier stages remains unknown. AIM:To characterize<br />

the changes in the circulating lipidome with disease<br />

progression from no fibrosis to advanced fibrosis in subjects<br />

with NAFLD. METHODS:Plasma samples collected at the time<br />

of liver biopsy were analyzed by UPLC-MS. Specifically, amino<br />

acids, fatty acyls, bile acids, glycerolipids, glycerophospholipids,<br />

sterols and sphingolipids were analyzed. RESULTS:<br />

11controls, 29NAFLD with no fibrosis, 43NAFLD with early<br />

stage fibrosis (stage1-2) and 22NAFLD with advanced fibrosis<br />

(stage3-4) were studied. 22metabolites were associated<br />

with disease progression (Figure). While cholesteryl esters and<br />

phosphatidylcholines, decreased with the progression [t-test<br />

Control vs.NAFLD with no fibrosis; early stage or advanced<br />

fibrosis: PC(O-24:1/0:0) and PC(18:2/20:4):p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1263A<br />

0.84), 0.84(0.76-0.92) and 0.88(0.82-0.94), respectively.<br />

NFS had the highest NPV (98%), while NPVs of AAR, APRI,<br />

BARD and FIB-4 were 87%, 85%, 94.5% and 94%, respectively.<br />

56% to 83% of cohort 1 pts were correctly classified as<br />

having or not AF. In cohort 2 (mean age 46±13, males 65%,<br />

diabetes 28%, mean BMI 30±5, AF 23.5%), AUROCs of AAR,<br />

APRI, BARD, FIB-4 and NFS were 0.58(0.50-0.66), 0.55(0.47-<br />

0.63), 0.64(0.57-0.71), 0.68(0.61-0.75) and 0.66(0.59-<br />

0.73), respectively. NFS had the highest NPV (84%), while<br />

NPVs of AAR, APRI, BARD and FIB-4 were 78%, 77.5%, 83%<br />

and 82%. 64% to 71% of cohort 2 pts were correctly classified<br />

as having or not AF. In cohort 3 (mean age 49±13, males<br />

61%, diabetes 38%, mean BMI 31±5, AF 22.7%), AUROCs<br />

of AAR, APRI, BARD, FIB-4 and NFS were 0.65(0.58-0.71),<br />

0.73(0.67-0.79), 0.71(0.65-0.76), 0.74(0.68-0.8) and<br />

0.71(0.65-0.77), respectively. BARD had the higher NPV<br />

(89%), while NPVs of AAR, APRI, FIB-4 and NFS were 84%,<br />

81%, 88% and 86%. 61% to 78% of cohort 3 pts were correctly<br />

classified as having or not AF. Conclusions: There is a<br />

substantial variability across different cohorts in the diagnostic<br />

accuracy of non-invasive scoring systems for ruling-out AF in<br />

NAFLD. Complex scores (BARD, FIB-4, NFS) perform better<br />

than simple ones (AAR and APRI), but none of them is consistently<br />

within the acceptable range for clinical-decision making<br />

in an individual patient. Combination with imaging methods<br />

should be tested in future <strong>studies</strong>.<br />

Disclosures:<br />

Manuel Romero-Gomez - Advisory Committees or Review Panels: Roche Farma,SA.,<br />

MSD, S.A., Janssen, S.A., Abbott, S.A.; Grant/Research Support: Ferrer,<br />

S.A.<br />

Giulio Marchesini - Advisory Committees or Review Panels: Sanofi-Synthelabo;<br />

Board Membership: GENFIT, Gilead, Glaxo, Novartis; Grant/Research Support:<br />

Merck Sharp & Dome; Speaking and Teaching: Novo Nordisk, Merck Sharp &<br />

Dome, Boerhinger Ingelheim, Eli Lilly, Astra-Zeneca<br />

Michael Trauner - Advisory Committees or Review Panels: MSD, Janssen, Gilead,<br />

Abbvie; Consulting: Phenex; Grant/Research Support: Intercept, Falk Pharma,<br />

Albireo; Patent Held/Filed: Med Uni Graz (norUDCA); Speaking and Teaching:<br />

Falk Foundation, Roche, Gilead<br />

Jean-Francois Dufour - Advisory Committees or Review Panels: Bayer, BMS, Gilead,<br />

AbbVie, Novartis, Sillagen, Genfit<br />

Vlad Ratziu - Advisory Committees or Review Panels: GalMed, Abbott, Genfit,<br />

Enterome, Gilead; Consulting: Tobira, Intercept, Exalenz, Sanofi-Synthelabo,<br />

Boehringer-Ingelheim<br />

The following authors have nothing to disclose: Fabio Nascimbeni, Salvatore<br />

Petta, Elisabetta Bugianesi, Stefano Bellentani, Christopher P. Day, Pierre<br />

Bedossa, Claudia P. Oliveira, Raluca Pais<br />

2165<br />

DJ-1 labeling index is a novel diagnostic biomarker for<br />

predicting progression of nonalcoholic steatohepatitis<br />

Masaaki Takamura 1 , Satoshi Yamagiwa 1 , Yasunobu Matsuda 2 ,<br />

Minoru Nomoto 1 , Toshifumi Wakai 1 , Shuji Terai 1 ; 1 Niigata University<br />

Graduate School of Medical and Dental Sciences, Niigata,<br />

Japan; 2 Niigata University Graduate School of Health Sciences,<br />

Niigata, Japan<br />

Background & Aims: Nonalcoholic steatohepatitis (NASH) is<br />

a common cause of chronic liver disease that can progress to<br />

cirrhosis and hepatocellular carcinoma. We have previously<br />

reported that hepatic proteins that were dysregulated in a<br />

murine NASH model functioned in a wide variety of pathways,<br />

including oxidative stress, glycolysis, the urea cycle, fatty acid<br />

oxidation, and apoptosis, based on a proteomic approach.<br />

Among the candidate proteins, we focused on the antioxidant<br />

protein DJ-1, which was upregulated in the livers of NASH<br />

model mice. In this study we analyzed the clinical implications<br />

of DJ-1 in patients with NASH. Methods: Liver biopsies and<br />

blood samples were obtained from 72 and 38 patients with<br />

nonalcoholic fatty liver disease (NAFLD) and 8 and 16 control<br />

subjects, respectively. The hepatic DJ-1 expression and serum<br />

DJ-1 levels were investigated using immunohistochemistry and<br />

enzyme-linked immunosorbent assay, respectively. The percentage<br />

of immunoreactive hepatocyte nuclei in the total number<br />

of hepatocyte nuclei was defined as DJ-1 labeling index (LI).<br />

NASH was assessed according to the Brunt classification.<br />

Results: DJ-1 was localized in the hepatocyte cytoplasm in the<br />

control and was incrementally translocated to the nucleus in<br />

NAFLD. A stepwise increase in DJ-1 LI (median [interquartile<br />

range]) was found from the control group (0 % [0-0.16]) to the<br />

nonalcoholic fatty liver (NAFL) group (4.34 % [1.50-15.4]) to<br />

the NASH group (38.5 % [29.6-48.7]; p


1264A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

viduals with NAFLD and HCC were significantly older (63.5<br />

± 9.7 vs. 57.4 ± 10.5 years, P=0.0004) and more frequently<br />

male (67% vs. 38%, P=0.0004). In addition, those with HCC<br />

were more likely to have esophageal varices (57% vs. 40%,<br />

P=0.047) and a trend toward a higher frequency of ascites<br />

(50% vs. 37%, P=0.1). However, the average MELD score was<br />

lower in those with HCC compared to controls (11.3 ± 3.6 vs.<br />

13.3 ± 4.6, P=0.05). Individuals with HCC had a trend toward<br />

a higher prevalence of dyslipidemia (52% vs. 38%, P=0.09)<br />

and cardiovascular disease (28% vs. 16%, P=0.09) and were<br />

more likely to use statins (31% vs. 16%, P=0.04). On multivariate<br />

regression analysis, increasing age (OR 1.09, 95% CI<br />

1.03-1.15, P=0.002), male gender (OR 4.31, 95% CI 1.69-<br />

10.97, P=0.002) and the presence of varices (OR 3.93, 95%<br />

CI 1.43-10.82, P=0.008) were significantly associated with<br />

HCC. Conclusion: Male gender, increasing age, and the presence<br />

of varices are associated with HCC in NAFLD cirrhosis.<br />

Table. Selected characteristics of NAFLD cirrhosis patients with<br />

and without HCC<br />

2167<br />

Disturbance of regulatory T cells, MDSCs and NK cells is<br />

involved in NASH and mouse model of NASH<br />

Tatsuki Morosawa, Yasuteru Kondo, Takayuki Kogure, Jun Inoue,<br />

Yu Nakagome, Osamu Kimura, Yuta Wakui, Tomoaki Iwata,<br />

Yasuyuki Fujisaka, Teruyuki Umetsu, Tooru Shimosegawa; Tohoku<br />

University Hospital, Sendai, Japan<br />

Background and aim: The involvement of lipotoxicity, insulin<br />

resistance, intestinal bacteria, and disturbance of the immune<br />

system has been reported to be involved in the pathogenesis<br />

of non-alcoholic steatohepatitis (NASH). However, the immunological<br />

dis-regulation in NASH and mouse models of NASH<br />

has not been clarified yet. The aim of this study is to analyze<br />

the immunological responses in NASH patients and mouse<br />

models of NASH. Material and method: Patients: Permission<br />

for the study was obtained from the ethics committee at our<br />

university. We analyzed various kinds of lymphocytes in the<br />

peripheral blood from 35 NASH patients who had been diagnosis<br />

histologically, eight patients with non-alcoholic fatty liver<br />

disease, 35 patients with chronic hepatitis C, 20 patients with<br />

chronic hepatitis B, and 18 healthy subjects. We subjected the<br />

samples to BD FACS Canto flow cytometry and analyzed;<br />

(1) myeloid dendritic cells (mDC)1, mDC2, plasmacytoid DC<br />

(pDC), (2) T cells (regulatory T cells (Tregs), Th1, Th2, Th17),<br />

(3) Natural Killer (NK) cells, NK-T cells, (4) Myeloid-Derived<br />

Suppressor Cells (MDSCs). Further, the expression of PD-1,<br />

PD-L1, NKG2D and CD69 on these immune cells was analyzed<br />

to detect the level of functional activity. In addition, we<br />

separated the immune cells from the liver and spleen tissue of<br />

STAM mice, MCD mice, and C57BL/6 mice (normal diet group<br />

or high-fat diet group) and conducted an analysis of immune<br />

functions. Co-cultivation analysis: HepG2 cells with or without<br />

fatty acid were used for co-cultivation with PBMC to determine<br />

the effect of fatty hepatocytes on various immune cells. Results:<br />

We observed a significant increase of Tregs and MDSCs in the<br />

immune cells of the NASH patients compared with the NAFLD<br />

patients and healthy subjects (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1265A<br />

(IRR 5.7, p=0.001), but the presence of USS detected hepatic<br />

steatosis was not. Assessment of diagnostic test accuracy found<br />

area under the receiver operator curves of 0.7-0.8 for the liver<br />

fibrosis markers, with optimal cut-offs giving sensitivities of up<br />

to 90% but positive predictive values of 10% of total body weight lost (OR<br />

21.67; 95% CI 3.73-125.77; p=0.0006), and change in<br />

weight measured as mean kilograms lost (mean difference of<br />

21.0 kg; 95% CI 8.40-33.60; p=0.001), regression of fibrosis<br />

(OR 18.90; 95% CI 2.10-170.39; p=0.009), reduction in<br />

ALT (-3.87, 95%CI -3.89-3.85, p=


1266A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

and 9 reported a change in fibrosis score (D weighted mean<br />

-0.34). Conclusion: In this meta-analysis, patients with NAFLD<br />

who underwent bariatric surgery showed significant reduction<br />

of all-cause mortality, weight loss, fibrosis score and ALT. Further<br />

<strong>studies</strong> are needed to assess the longterm benefit of bariatric<br />

surgery in the treatment of patients with NAFLD.<br />

Disclosures:<br />

Thomas D. Schiano - Advisory Committees or Review Panels: salix, merck, gilead,<br />

pfizer; Grant/Research Support: galectin, massbiologics, biotest<br />

The following authors have nothing to disclose: Amanda Schneier, Naveen Ganjoo,<br />

Rachel Pinotti, Malcolm M. Wells<br />

2171<br />

Clinical features and risk factors associated with<br />

Non-alcoholic steatohepatitis (NASH) in Spain. On<br />

behalf of HEPAmet Registry<br />

Rocío Gallego-Durán 12 , Rocío Aller 5 , Javier Ampuero 12 , Carmelo<br />

García-Monzón 1 , Jesus M. Banales 2 , Javier Crespo 3 , Víctor Aguilar-Urbano<br />

6 , María Luisa García-Torres 7 , José Luís Calleja 4 , Jose<br />

Luis Olcoz 8 , Javier Salmeron 9 , Martin Prieto 10 , Javier García-Samaniego<br />

11 , Moisés Diago 7 , Helena Pastor 12 , María J. Pareja 12 ,<br />

Miguel Fernandez-Bermejo 13 , Judith Gomez-Camarero 14 , Raul<br />

J. Andrade 15 , Pamela Estévez 16 , Conrado M. Fernández-Rodríguez<br />

17 , Lourdes Grande 18 , Jose María Moreno-Planas 19 , Marta<br />

Maraver Zamora 20 , Agustin Albillos 21 , Manuel Romero-Gomez 12 ;<br />

1 Santa Cristina University Hospital, Madrid, Spain; 2 Biodonostia<br />

Research Institute, San Sebastián, Spain; 3 HUM Valdecillas.<br />

IDIVAL, Santander, Spain; 4 Puerta de Hierro Hospital, Madrid,<br />

Spain; 5 Clínico Valladolid University Hospital, Valladolid, Spain;<br />

6 Costa del Sol Hospital, Málaga, Spain; 7 General de Valencia<br />

Hospital, Valencia, Spain; 8 León University Hospital, León, Spain;<br />

9 San Cecilio University Hospital, Granada, Spain; 10 La Fe University<br />

Hospital, Valencia, Spain; 11 Carlos III Hospital, Madrid,<br />

Spain; 12 Valme University Hospital and CIBERehd, Sevilla, Spain;<br />

13 San Pedro de Alcántara Hospital, Cáceres, Spain; 14 Burgos University<br />

Hospital, Burgos, Spain; 15 Virgen de la Victoria University<br />

Hospital, Málaga, Spain; 16 Meixoxeiro Hospital, Vigo, Spain;<br />

17 Fundación Alcorcón University Hospital, Madrid, Spain; 18 Virgen<br />

del Rocío University Hospital, Sevilla, Spain; 19 Albacete University<br />

Hospital, Albacete, Spain; 20 Juan Ramón Jiménez University<br />

Hospital, Huelva, Spain; 21 Ramón y Cajal University Hospital,<br />

Madrid, Spain<br />

Aims: To describe the individual profile of NASH Spanish<br />

patients and to identify risk factors associated with this condition.<br />

Material and methods: Spanish Association for the<br />

Study of the Liver (AEEH) HEPAmet Registry is a multicentric<br />

monitored patient database, including Spanish NAFLD-biopsy<br />

patients that fulfilled at least 2/4 inclusion criteria (steatosis<br />

by ultrasound(US); ALT or AST above upper limits of normality(ULN);<br />

HOMA-IR>4 or metabolic syndrome(MetS) defined<br />

by ATPIII criteria). Demographic, anthropometric, concomitant<br />

diseases and medication, US, transient elastography, analytical<br />

and anatomopathological data were recorded.Univariate<br />

and multivariate analyses were performed to identify risk factors<br />

for NASH defined by histology. Results: Seven hundred<br />

and twenty-six patients were included, 43.8%(318/726)<br />

men, mean age 51.5+12.9 years old. 50%(370/726) presented<br />

MetS, 34%(250/726) morbid obese underwent bariatric<br />

surgery, 73%(528/726) overweight (BMI>25 kg/<br />

m 2 ) and 1.2%(9/689) suffered from any cerebrovascular<br />

event. NASH at liver biopsy was found in 47%(343/726) of<br />

patients. Fibrosis stage distribution was F0:50%(364/726),<br />

F1:30%(218/726), F2:9%(65/726), F3:8%(57/726) and<br />

F4:3%(22/726). Extrahepatic conditions associated with<br />

NASH vs simple steatosis were T2DM(30.9% vs 21.4%,<br />

p=0.005) and hyperuricemia(7.8% vs 3.6%, p=0.043). No<br />

differences were observed between both groups according<br />

to MetS(55%(182/330) vs 51%(188/371); p=ns) and overweight<br />

prevalence(76%(250/329) vs 74%(278/376);p=ns).<br />

In univariate analysis, age,T2DM, arterial hypertension,<br />

hypercholesterolemia, BMI, AST, ALT, HOMA-IR, transferrin<br />

saturation index(TSI), platelets and ferritin were significantly<br />

associated with NASH. In multivariate analysis, HOMA-IR, ALT<br />

and TSI were confirmed as independent risk factors for NASH.<br />

Conclusions: The results of this large cohort confirmed insulin<br />

sensitivity, hepatic inflammation and iron metabolism as key<br />

factors for NASH. Interestingly, MetS and overweight seems to<br />

be equally distributed.<br />

Table 1. Univariate and multivariate analyses in NASH patients<br />

Disclosures:<br />

Martin Prieto - Advisory Committees or Review Panels: Bristol, Gilead<br />

Javier García-Samaniego - Consulting: Bristol-Myers-Squibb, Gilead, Janssen,<br />

Abbvie<br />

Moisés Diago - Grant/Research Support: MSD, JANSSEN; Speaking and Teaching:<br />

BMS, ABBVIE, GILEAD<br />

Manuel Romero-Gomez - Advisory Committees or Review Panels: Roche Farma,SA.,<br />

MSD, S.A., Janssen, S.A., Abbott, S.A.; Grant/Research Support: Ferrer,<br />

S.A.<br />

The following authors have nothing to disclose: Rocío Gallego-Durán, Rocío Aller,<br />

Javier Ampuero, Carmelo García-Monzón, Jesus M. Banales, Javier Crespo, Víctor<br />

Aguilar-Urbano, María Luisa García-Torres, José Luís Calleja, Jose Luis Olcoz,<br />

Javier Salmeron, Helena Pastor, María J. Pareja, Miguel Fernandez-Bermejo,<br />

Judith Gomez-Camarero, Raul J. Andrade, Pamela Estévez, Conrado M. Fernández-Rodríguez,<br />

Lourdes Grande, Jose María Moreno-Planas, Marta Maraver<br />

Zamora, Agustin Albillos


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1267A<br />

2172<br />

WITHDRAWN<br />

2173<br />

MHC Class I Related Gene A (MICA) Alleles are Independently<br />

Associated with Advanced Pathogenic Features<br />

of Non Alcoholic Fatty Liver Disease (NAFLD)<br />

Azza Karrar 1 , Mohamad Houry 1 , Irfan Ali 2 , Dinan Abdelatif 2 ,<br />

Pegah Golabi 1 , Mehmet Sayiner 1 , Zahra Younoszai 1 , Lakshmi<br />

Alaparthi 2 , James N. Cooper 1 , Munkhzul Otgonsuren 1 , Zachary<br />

D. Goodman 2 , Zobair M. Younossi 2 ; 1 Betty and Guy Beatty Center<br />

for Integrated Research, Inova Health Systems, Falls Church, VA;<br />

2 Center For Liver Disease, Department of Medicine, Inova Fairfax<br />

Hospital, Falls Church, VA<br />

Background and Aims: MICA is a highly polymorphic gene<br />

that modulates immune surveillance by binding to its receptor<br />

on natural killer cells, and its genetic polymorphisms have been<br />

associated with chronic immune-mediated diseases. NAFLD is<br />

characterized by fat accumulation in the liver and inflammation<br />

that can potentially progress to Non Alcoholic Steatohepatitis<br />

(NASH) and fibrosis. To-date, no data exists describing the role<br />

of MICA in the pathogenesis of NAFLD. The overall aim is to<br />

study the association between MICA polymorphism and NASH<br />

and its histological features. Methodology: DNA from biopsy-proven<br />

NAFLD patients were genotyped using (PCR-SSO) for<br />

MICA alleles. Liver biopsies were assessed for the presence of<br />

NASH, degree of fibrosis and inflammation. Multivariate analysis<br />

was performed to draw associations between MICA alleles<br />

and the different variables; p values ≤ 0.05 were considered<br />

to be significant. Results: The study cohort was comprised of<br />

134 patients; NAFLD: [NASH (n=32), Age 45.7±10.6 years;<br />

Male 37.5% and non NASH (n=49), Age 42.5±10.6 years;<br />

Male 10.2%] and Control: [(n=53), Age 55.1±16 years; Male<br />

59.6%]. Univariate analysis showed that MICA*007[4(4.9%)]<br />

and MICA*012[1(1.2%)] were less frequent in NAFLD<br />

patients, than healthy controls [8(15.1%), p=0.044], [5(9.4%),<br />

p=0.025] respectively. Furthermore, MICA*011 Odds Ratio<br />

[OR: 7.14(1.24-41.0), p=0.028] is associated with higher risk<br />

for NASH. Histologically, MICA*002 was found to be associated<br />

with lower risk for focal necrosis [OR: 0.34 (0.12-0.92),<br />

p=0.033], lymphocyte-mediated inflammation [OR: 0.13(0.02-<br />

1.05), p=0.056] and advanced fibrosis [OR: 0.20(0.04-<br />

0.92), p=0.039]. In addition, MICA*004 was associated with<br />

lower risk for portal fibrosis [OR: 0.30(0.09-1.02), p=0.054]<br />

and portal inflammation [OR: 0.29(0.08-1.05), p=0.059]. On<br />

the other hand, MICA*017 was associated with higher risk<br />

for lymphocyte-mediated inflammation [OR: 5.08(1.11-23.2),<br />

p=0.036]. Multivariate analysis showed that MICA*011<br />

[OR: 7.38(1.10-49.6), p=0.04] was still independently associated<br />

with higher risk for NASH. MICA*002 was found to<br />

be independently associated with lower risk for focal necrosis<br />

[OR: 0.24(0.08-0.74), p=0.013] and advanced fibrosis [OR:<br />

0.11(0.02-0.70), p=0.019]. MICA*017 continued to be independently<br />

associated with higher risk for lymphocyte-mediated<br />

inflammation [OR: 5.12(1.12-23.5), p=0.035]. Conclusions:<br />

MICA alleles were found to be independently associated with<br />

NASH, inflammation and fibrosis. Thus, they may contribute to<br />

the injury and scarring of liver. Additional research is required<br />

to investigate the potential role of MICA in suceptability and<br />

protection to NAFLD.<br />

Disclosures:<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Azza Karrar, Mohamad Houry,<br />

Irfan Ali, Dinan Abdelatif, Pegah Golabi, Mehmet Sayiner, Zahra Younoszai,<br />

Lakshmi Alaparthi, James N. Cooper, Munkhzul Otgonsuren<br />

2174<br />

A 32-gene poor prognosis signature identifies a subset<br />

of NASH patients at high risk of long-term mortality<br />

after bariatric surgery<br />

Nicolas Goossens 1,2 , Won-Min Song 3 , Minoa K. Jung 4 , Philippe<br />

Morel 4 , Shigeki Nakagawa 5 , Xiaochen Sun 1 , Anu Venkatesh 5 ,<br />

Anna Koh 5 , Jean-Louis Frossard 2 , Laurent Spahr 2 , Francesco<br />

Negro 2,6 , Scott L. Friedman 1 , Laura Rubbia-Brandt 6 , Emiliano Giostra<br />

2 , Yujin Hoshida 5 ; 1 Division of Liver Diseases, Department of<br />

Medicine, Icahn School of Medicine at Mount Sinai, New York,<br />

NY; 2 Division of Gastroenterology and Hepatology, Geneva University<br />

Hospital, Geneva, Switzerland; 3 Department of Genetics<br />

and Genomic Sciences, Icahn Institute of Genomics and Multiscale<br />

Biology, Icahn School of Medicine at Mount Sinai, New York, NY;<br />

4 Division of Surgery, Geneva University Hospital, Geneva, Switzerland;<br />

5 Division of Liver Diseases, Department of Medicine, Liver<br />

Cancer Program, Tisch Cancer Institute, Icahn School of Medicine<br />

at Mount Sinai, New York, NY; 6 Division of Pathology, Geneva<br />

University Hospital, Geneva, Switzerland<br />

Background & Aims Non-alcoholic steatohepatitis (NASH) is<br />

associated with increased mortality in subjects undergoing bariatric<br />

surgery (Goossens et al., EASL 2015). We aimed to<br />

determine whether bariatric surgery improved survival of obese<br />

patients with NASH and to identify molecular-based prognostic<br />

biomarkers in high-risk NASH patients undergoing bariatric<br />

surgery. Methods We included 492 subjects undergoing Rouxen-Y<br />

gastric bypass bariatric surgery in a single center between<br />

January 1997 and December 2004 with a median follow-up<br />

of up to 17 years. Survival of the entire cohort and of the<br />

subset of NASH subjects (defined here as steatosis and raised<br />

ALT for all bariatric surgery and NHANES III subjects) was<br />

compared to propensity score-matched subjects from the third<br />

National Health and Nutrition Examination Survey (NHANES<br />

III). We performed gene expression profiling in formalin-fixed,<br />

paraffin-embedded hepatic tissue of 47 NASH subjects from<br />

the bariatric surgery cohort. A prognostic 32-gene signature<br />

(PMID: 25143343) was evaluated in the NASH subjects and<br />

gene regulatory networks were identified by using Multiscale<br />

Embedded Gene Co-expression Network Analysis (MEGENA).<br />

Results In the bariatric surgery cohort, median body mass index<br />

(BMI), age, proportion of histologically documented NASH<br />

were 44kg/m 2 , 41 years, and 12%, respectively. Bariatric<br />

surgery was associated with reduced mortality compared to<br />

propensity-score matched NHANES III subjects (HR 0.54,<br />

p=0.035), however this benefit was not observed in the<br />

subgroup of NASH patients (HR 0.97, p=0.94). Within the<br />

47 NASH patients with gene expression analysis (baseline<br />

BMI 44.6kg/m 2 , 78% with fibrosis, 15% death during follow-up),<br />

a 32-gene poor prognosis signature prediction was<br />

associated with increased mortality in multivariable analysis<br />

(HR 7.7, p=0.045). The 32-gene poor prognosis NASH subjects<br />

(49%) had increased mortality compared to NHANES<br />

III NASH subjects (HR 4.4, p=0.005). MEGENA identified<br />

9 distinct gene regulatory modules associated with TNF-α/<br />

NF-κB signaling (p


1268A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

long-term mortality compared to control obese NASH subjects.<br />

Gene expression profiling of hepatic tissue in NASH subjects<br />

revealed the induction of TNF-α /NF-κB and fibrogenic pathways<br />

as potential targets.<br />

Disclosures:<br />

Francesco Negro - Advisory Committees or Review Panels: Bristol-Myers Squibb,<br />

MSD, Gilead, AbbVie; Grant/Research Support: Gilead<br />

Scott L. Friedman - Advisory Committees or Review Panels: Pfizer Pharmaceutical;<br />

Consulting: Conatus Pharm, Exalenz, Genfit, Exalenz Biosciences, Eli Lilly PHarmaceuticals,<br />

Fibrogen, Boehringer Ingelheim, Nitto Corp., Immune Therapeutics,<br />

Synageva, Roche/Genentech Pharmaceuticals, DeuteRx, Abbvie, Novartis,<br />

RuiYi, Kinemed, Sanofi Aventis, Takeda Pharmaceuticals, Nimbus Therapeutics,<br />

Bristol Myers Squibb, Astra Zeneca, Sandhill Medical Devices, Galmed, Northern<br />

Biologics, Enanta Pharmaceuticals, Regado Bioscience, Raptor Pharmaceuticals,<br />

Teva Pharmaceuticals, Zafgen Pharmaceuticals, Merck Pharmaceuticals,<br />

Debio Pharmaceuticals; Grant/Research Support: Galectin Therapeutics, Tobira<br />

Pharm; Stock Shareholder: Angion Biomedica, Intercept Pharma<br />

The following authors have nothing to disclose: Nicolas Goossens, Won-Min<br />

Song, Minoa K. Jung, Philippe Morel, Shigeki Nakagawa, Xiaochen Sun, Anu<br />

Venkatesh, Anna Koh, Jean-Louis Frossard, Laurent Spahr, Laura Rubbia-Brandt,<br />

Emiliano Giostra, Yujin Hoshida<br />

2175<br />

Naltrexone/Bupropion Extended-Release 32 mg/360<br />

mg significantly improves liver enzymes in obese/overweight<br />

individuals with elevated liver enzymes<br />

Allison Winokur 1 , Amy Halseth 2 , Claire Dybala 1 , Hung Lam 3 ,<br />

Steve Chen 1 , Naga P. Chalasani 4 ; 1 Takeda Pharmaceuticals<br />

U.S.A., Inc., Deerfield, IL; 2 Orexigen Therapeutics, Inc., La Jolla,<br />

CA; 3 Takeda Development Center Americas, Inc., Deerfield, IL;<br />

4 Division of Gastroenterology and Hepatology, Indiana University<br />

School of Medicine, Indianapolis, IN<br />

Background: Non-alcoholic fatty liver disease (NAFLD) can be<br />

a consequence of obesity. Alanine aminotransferase (ALT) is<br />

often increased in patients with NAFLD and is used as a measure<br />

of liver dysfunction. NAFLD can improve with weight loss.<br />

Aim: This posthoc analysis examined the effect of extendedrelease<br />

naltrexone/bupropion (NB), approved in the US and<br />

Europe for obesity management, on ALT in non-diabetic, obese<br />

patients. Methods: Subjects from 3 Phase 3 <strong>studies</strong> who were<br />

randomly assigned to receive NB or placebo (PBO), and who<br />

remained in the trials for 56 weeks were included. NB subjects<br />

had to lose at least 5% of their body weight at week 16 to<br />

be included, consistent with prescribing information. Analyses<br />

by baseline (BL) ALT quartile included change from BL in<br />

ALT, % of subjects achieving a 25% or 50% reduction in ALT,<br />

and % subjects achieving Prati criteria (≤19 IU/L for women;<br />

≤30 IU/L for men) at Week 56. Results: Study population consisted<br />

of 781 NB subjects (mean age 46 years, BMI 36 kg/<br />

m 2 , 85% female, median ALT 22 IU/L) and 663 PBO subjects<br />

(mean age 46 years, BMI 36 kg/m 2 , 84% female, median ALT<br />

22 IU/L). NB responders completing 56 weeks of treatment<br />

experienced significantly greater weight loss vs. PBO (-12.4%<br />

vs. -2.7%, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1269A<br />

ALT from randomisation to the end of treatment in the metronidazole<br />

and inulin treatment arm (Group 3) compared to<br />

the placebo group (group 1), (mean change in ALT -18.6 vs<br />

-0.5 U/L respectively; p=0.04). The reduction in ALT in group<br />

2 was not significantly different to the placebo group (mean<br />

change ALT -2.6 vs 0.5U/L respectively; p=0.5). At the end of<br />

treatment reduction in BMI, fasting lipids, glycaemia, BP and<br />

Fibroscan ® CAP scores did not significantly differ between the<br />

three groups. Conclusions: This is the first clinical trial evidence<br />

that supplementation with prebiotic inulin following brief metronidazole<br />

therapy can further reduce liver aminotransferase<br />

after 4 weeks VLCD therapy in patients with NAFLD. This is a<br />

potentially viable clinical therapy which requires validation in<br />

larger clinical <strong>studies</strong> of longer duration.<br />

Disclosures:<br />

The following authors have nothing to disclose: David W. Orr, Rinki Murphy<br />

2177<br />

Diagnostic accuracy for non-alcoholic fatty liver disease<br />

(NAFLD) with controlled attenuation parameter (CAP)<br />

measured by transient elastography<br />

Masahiro Kikuchi 1,2 , Rumiko Umeda 2 , Kota Tsuruya 2 , Hirokazu<br />

Shiozawa 2 , Miho Kikuchi 1 , Masahiko Takahashi 1 , Yoshinori<br />

Horie 4 , Yasuhiro Nishizaki 3 ; 1 Gastroenterology, National Hospital<br />

Organization Tokyo Medical Center, Tokyo, Japan; 2 Gastroenterology,<br />

Tokai University Tokyo Hospital, Tokyo, Japan; 3 Life Care<br />

Center, Tokai University Tokyo Hospital, Tokyo, Japan; 4 International<br />

University of Health and Welfare, Research Centre of Clinical<br />

Medicine, Sanno Medical Center, Tokyo, Japan<br />

Purpose: We reported that pulse wave velocity (PWV) as arteriosclerosis<br />

index accelerate and remnant like particles cholesterol<br />

(RLP-C), triglyceride (TG), and free fatty acid (FFA) which<br />

are arteriosclerosis promoting factors increase in non-alcoholic<br />

fatty liver disease (NAFLD). As NAFLD is closely related to<br />

cardiovascular deaths, a detection of the early stage of NAFLD<br />

patient is important. In this study, we aimed to investigate the<br />

diagnostic performance of controlled attenuation parameter<br />

(CAP) based on FibroScan ® 502 (Echosens,France) in NAFLD<br />

patients. Method: 341 non-viral hepatitis patients who intake<br />

alcohol less than 20g ethanol per day were enrolled. The<br />

existence of the fatty liver was assessed by B-mode ultrasonography,<br />

and it divided into two groups. (male: NAFLD (+)<br />

126cases and NAFLD (-) 97cases, female: NAFLD (+) 34cases<br />

and NAFLD (-) 84cases), and CAP was performed simultaneously.<br />

Moreover, fatty liver index (FLI) (Bedogni G.2006),<br />

hepatic steatosis index (HSI) (Lee JH.2010) and lipid accumulation<br />

product (LAP) (Chinag Jui-Kun.2012) known as liver steatosis<br />

prediction formulas from various physical measurements<br />

and multiple blood markers were investigated, and we compared<br />

the usefulness for detecting NAFLD with ROC analysis.<br />

Result: CAP value, FLI, HSI, and LAP were significantly high in<br />

the NAFLD (+) group compared with the (-) group (P< 0.0001).<br />

In male the area under ROC curve (AUROC) for NAFLD diagnostic<br />

tool was CAP: 0.890, FLI: 0.779, HSI: 0.811, and LAP:<br />

0.718 respectively. In female it was CAP: 0.894, FLI: 0.939,<br />

HSI: 0.911, and LAP: 0.928. The CAP was correlated with<br />

diagnosis of NAFLD especially in male (cut-off value=257d-<br />

B/m, sensitivity=87%, specificity=79%). Conclusion: The utility<br />

of CAP value was verified as accurate diagnosis tool for<br />

NAFLD. The detection of CAP value would influence the more<br />

aggressive treatment intervention and the raised motivations of<br />

therapy for NAFLD patients. This instrument must be contributed<br />

to not only the detection of NAFLD but also the prevention of<br />

arteriosclerosis related cardiovascular deaths.<br />

Disclosures:<br />

The following authors have nothing to disclose: Masahiro Kikuchi, Rumiko<br />

Umeda, Kota Tsuruya, Hirokazu Shiozawa, Miho Kikuchi, Masahiko Takahashi,<br />

Yoshinori Horie, Yasuhiro Nishizaki<br />

2178<br />

DACRAs are novel therapeutic candidates for the treatment<br />

of NAFLD and NASH<br />

Mette J. Nielsen 1,2 , Sara T. Hjuler 1 , Sofie Gydesen 1 , Morten A.<br />

Karsdal 1 , Kim Henriksen 1 ; 1 Nordic Bioscience A/S, Herlev, Denmark;<br />

2 Department of Gastroenterology and Hepatology, Odense<br />

University Hospital, University of Southern Denmark, Odense, Denmark<br />

Background and aim: Nonalcoholic fatty liver disease (NAFLD)<br />

is a significant complication of obesity and in many cases<br />

it leads to the development of nonalcoholic steatohepatitis<br />

(NASH). NAFLD is defined by the presence of liver fat accumulation<br />

exceeding 5% of hepatocytes and is strongly associated<br />

with the metabolic syndrome. While weight loss is beneficial,<br />

there are no pharmacological treatments for NAFLD and<br />

NASH, and novel candidates are intensely sought. DACRAs<br />

(dual amylin and calcitonin receptor agonists) are novel peptide-based<br />

drug candidates, which have shown promise in the<br />

treatment of type 2 diabetes due to potent weight reduction<br />

and alleviation of insulin resistance. Methods: In this study, we<br />

used High-Fat Diet (HFD)-induced obese Sprague-Dawley rats<br />

treated with different doses (0.625-10mg/kg) the DACRA KBP-<br />

042 for 8 weeks to elucidate the effects of a novel DACRA on<br />

liver fat accumulation, body weight, insulin/glucagon balance<br />

and glucose homeostasis. Results: A dose-dependent and sustained<br />

weight loss was obtained, resulting in a 4.7 % weight<br />

reduction (20 % vehicle-corrected) in the KBP-042 10 mg/kg<br />

group with a transient reduction in food intake. The weight<br />

loss was clearly superior to calorie-matched groups. Corresponding<br />

with the weight loss, visceral fat depots were significantly<br />

reduced (perirenal Adipose Tissue (AT): p


1270A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2179<br />

Digoxin provides hepatoprotection through inhibition of<br />

HIF-1a-NOX4-ROS active pathway<br />

Xinshou Ouyang 1 , Ji-Yuan Zhang 2 , Dechun Feng 3 , Shi-Ying Cai 1 ,<br />

Irma Garcia-Martinez 1 , Sheng-Na Han 1 , Rafaz Hoque 1 , Yonglin<br />

Chen 1 , Fu-Sheng Wang 2 , Bin Gao 3 , Natalie J. Torok 4 , Wajahat Z.<br />

Mehal 1 ; 1 Yale University, New Haven, CT; 2 Liver Failure Therapy<br />

and Research Center, Beijing 302 Hospital (PLA 302 Hospital),<br />

Beijing, China; 3 NIAAA, NIH, Bethesda, MD; 4 Gastroenterology<br />

and Hepatology, UC Davis, Sacramento, CA<br />

Introduction: Reactive oxygen species (ROS) driven tissue damage<br />

is common to liver diseases. NADPH oxidase family members<br />

have shown important contribution to the production of<br />

hepatic ROS, and this dependent in part to HIF-1a. The cardiac<br />

glycoside digoxin was identified as potent HIF-1a antagonist<br />

but its role in liver disease is not well defined. Aim: To assess<br />

whether digoxin has therapeutic effects in Non Alcoholic Fatty<br />

Liver Disease (NASH) and alcoholic liver disease (ALD) in<br />

mice, and investigate the molecular mechanisms in both mouse<br />

and human cell cultures. Methods: C57BL/6J male mice were<br />

placed on a 45% high fat diet (HFD) for 11weeks with and without<br />

digoxin (ip 1, 0.2 and 0.05 mg/kg twice a week). Digoxin<br />

1mg/kg ip daily in mice results in the therapeutic serum levels<br />

achieved in humans (0.5-2 ng/ml). Plasma ALT, liver histology,<br />

neutrophil staining, leukocytes profiling, mitochondrial reactive<br />

oxygen species (ROS) generation, and gene transcriptome<br />

microarrays were analyzed. The chronic plus binge model of<br />

ALD was performed. The mechanism of digoxin on inhibition<br />

of HIF-1a mediated NADPH oxidase 4 (NOX4) transcription<br />

and ROS production was tested by RT-PCR, reporter luciferase<br />

and ChIP-PCR assay. Results: Digoxin dose-dependently<br />

reduced histological injury, neutrophilic infiltrate and serum<br />

ALT values in both co-treatment (starting digoxin same time<br />

with HFD, ALT, 417 +/- 398 U/L in HFD vs 91 +/- 73 U/L<br />

in HFD+DIG, P< 0.001) and post-treatment (starting digoxin<br />

after 4 weeks HFD, neutrophil 24.6% in HFD vs 14.3% in<br />

HFD+DIG; monocytes 31.6% in HFD vs 19.1% in HFD+DIG;<br />

ALT, 400 +/- 130 U/L in HFD vs 80 +/- 17 U/L in HFD+DIG,<br />

P< 0.001) without a reduction in food intake. A low dose of<br />

digoxin (0.05 mg/kg) also shows significant protective effects<br />

again injury and inflammation in both NASH and ALD models.<br />

Further microarray analysis in HFD liver tissues revealed<br />

that digoxin down-regulated ROS metabolism and antioxidant<br />

pathway including NOX4. In vitro data showed that digoxin<br />

dose-dependently inhibited mitochondrial ROS production<br />

under TLR and hydrogen peroxide stimulation in multiple mouse<br />

and human cell cultures. Digoxin increases VHL protein levels,<br />

resulting in a corresponding reduction HIF-1a protein levels. In<br />

addition digoxin blocks HIF-1a mediated NOX4 gene expression,<br />

promoter activity and NOX4 mediated ROS. Digoxin<br />

also reduces HIF-1a binding to the NOX4 promoter region.<br />

Conclusions: Digoxin has strong effects in reducing liver steatosis<br />

and inflammation in experimental models of NASH and<br />

ALD via a HIF-1a-NOX4-ROS pathway. Low dose digoxin may<br />

provide significant therapeutic value in the treatment of NASH<br />

and ALD patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Xinshou Ouyang, Ji-Yuan Zhang,<br />

Dechun Feng, Shi-Ying Cai, Irma Garcia-Martinez, Sheng-Na Han, Rafaz Hoque,<br />

Yonglin Chen, Fu-Sheng Wang, Bin Gao, Natalie J. Torok, Wajahat Z. Mehal<br />

2180<br />

Can FibroMeter VCTE be useful for detecting hepatic<br />

fibrosis in patients with non-alcoholic fatty liver disease?<br />

Elif Dincses 3 , Yusuf Yilmaz 1,2 ; 1 Department of Gastroenterology,<br />

Marmara University, School of Medicine, Istanbul, Turkey; 2 Liver<br />

Research Unit, Institute of Gastroenterology, Marmara University,<br />

Istanbul, Turkey; 3 Department of Internal Medicine, Marmara University,<br />

Istanbul, Turkey<br />

Hepatic fibrosis plays a paramount role in determining the<br />

prognosis of patients with non-alcoholic fatty liver disease<br />

(NAFLD). Originally developed for the non-invasive detection<br />

of fibrosis in patients with chronic viral hepatitis, the FibroMeter<br />

VTCE is a diagnostic tool comprising both biochemical<br />

markers and transient elastography (TE). We conducted this<br />

pilot study to investigate the diagnostic performance of the<br />

FibroMeter VTCE tool for diagnosing fibrosis in patients with<br />

biopsy-proven NAFLD and compare its diagnostic accuracy<br />

with those of the NAFLD fibrosis score (NFSA) and TE alone.<br />

We determined FibroMeter VTCE, NFSA, and TE in 52<br />

NAFLD patients. The results of liver biopsies were considered<br />

as the gold standard. Areas under the ROC curve (AUROCs)<br />

were used to express the diagnostic accuracy of each test.<br />

We detected significant (F≥2) and severe (F≥3) fibrosis in 20<br />

(38%) and 10 (19%) patients, respectively. The sensitivity of<br />

FibroMeter VTCE, NFSA, and TE for detecting significant<br />

fibrosis was 70%, 65%, and 75%, respectively, whereas specificity<br />

was 88%, 81%, and 78%. The sensitivity of FibroMeter<br />

VTCE, NFSA, and TE for diagnosing severe fibrosis was<br />

90%, 90%, and 100%, respectively, whereas specificity was<br />

93%, 78%, and 76%. ROC analysis showed that FibroMeter<br />

VTCE had a significantly larger AUROC (0.968) compared<br />

to both NFSA (0.833, P < 0.001) and TE (0.922, P < 0.05) for<br />

the detection of severe fibrosis (Table). To the best of our knowledge,<br />

this is the first study showing that FibroMeter VTCE – a<br />

tool originally developed for patients with viral hepatitis – can<br />

be superior to both NFSA and TE for the detection of severe<br />

hepatic fibrosis in patients with NAFLD. Financial disclosures:<br />

The authors declare no conficts of interest.<br />

AUROCs and sensitivities/specificities of FibroMeter VTCE,<br />

NFSA, and TE for the detection of severe fibrosis (F≥3) in patients<br />

with biopsy-proven non-alcoholic fatty liver disease<br />

Abbreviations: AUROC = area under the receiver operating<br />

characteristic curve; NFSA = NAFLD fibrosis score; TE = transient<br />

elastography.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elif Dincses, Yusuf Yilmaz


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1271A<br />

2181<br />

Controlled Attenuation parameter (CAP) for the diagnosis<br />

of steatosis in NonAlcoholic Fatty Liver Disease.<br />

Jean-Baptiste Hiriart 1 , Grace Wong 2 , Julien Vergniol 1 , Henry Lik-<br />

Yuen Chan 2 , Brigitte Le Bail 3 , Anthony W. Chan 4 , Faiza Chermak<br />

1 , Juliette Foucher 1 , Wassil Merrouche 1 , Angel ML Chim 2 ,<br />

Victor de Ledinghen 1 , Vincent W. Wong 2 ; 1 Hepatology, Hôpital<br />

Haut-Lévêque, Pessac, France; 2 Medicine and Therapeutics, Prince<br />

of Wales Hospital, Hong Kong, China; 3 Pathology, Hôpital Pellegrin,<br />

Bordeaux, France; 4 Anatomical and Cellular Pathology,<br />

Prince of Wales Hospital, Hong Kong, China<br />

Background & Aims: Controlled attenuation parameter (CAP)<br />

evaluated with transient elastography (FibroScan) is a recent<br />

method for non-invasive assessment of steatosis. Its usefulness<br />

in NonAlcoholic Fatty Liver Disease (NAFLD) is unknown. We<br />

prospectively investigated the performance of CAP for the<br />

diagnosis of steatosis in NAFLD. Methods: From April 2009 to<br />

June 2014, all CAP examinations performed in adult NAFLD<br />

patients with a liver biopsy performed within one week of CAP<br />

measurement were included. Liver biopsies were assessed for<br />

necro-inflammatory activity and fibrosis stage, NAFLD activity<br />

score (NAS), and steatosis graded as follows: S0, steatosis<br />

≤5% of hepatocytes; S1, 6-33%; S2, 34-66%; S3, >66%.<br />

The Fatty Liver Index (FLI) was calculated on the day of liver<br />

biopsy. Receiver-operating characteristics (ROC) curves were<br />

constructed to assess the overall accuracy of CAP. Results:<br />

A total of 261 consecutive patients (59% males, mean age<br />

56 years) from two ethnic groups with CAP measurement and<br />

satisfactory liver biopsy specimens were included. Mean BMI<br />

was 30,2 ± 5,1 kg/m 2 , 59% of patients had diabetes and<br />

65,9% had metabolic syndrome (according to the International<br />

Diabetes Federation criteria). Mean length of liver biopsy was<br />

25.1 ± 8.6 mm. 46% of patients had NAS score ≥ 5. Fibrosis<br />

stages were distributed as follows: 37,9% F0F1, 21,5% F2,<br />

21,8% F3 and 18,8% F4. All patients had steatosis > 5%,<br />

29,9% were graded S1, 38,3% S2 and 31,8% S3. The area<br />

under the ROC curve of CAP for steatosis ≥ S1, ≥ S2 and S3<br />

was 0.85, 0.80, and 0.66 respectively. At a cutoff value of<br />

310 dB/m, the sensitivity, specificity, positive and negative<br />

predictive values for ≥ S2 steatosis were 79%, 71%, 86% and<br />

71%, respectively. CAP had significantly better performance<br />

than FLI for the diagnosis of S2S3 steatosis (0,8 [0,73 – 0,86]<br />

vs 0,66 [0,59 – 0,74]; p < 0,001). Discordance of at least<br />

one grade between CAP and steatosis was observed in 81<br />

(31%) patients. By multivariate analysis, only steatosis S2S3<br />

was associated with no discordance. Only 15% of patients<br />

with NAS ≥ 5 had discordance between CAP and steatosis<br />

grade. By multivariate analysis, only BMI ≥ 30 kg/m 2 was<br />

significantly associated with CAP > 310 dB/m. Conclusion:<br />

CAP provides an immediate assessment of steatosis simultaneously<br />

with liver stiffness measurement. The strong association<br />

of CAP with steaosis, especially in patients with nonalcoholic<br />

steatohepatitis, and with elevated BMI could be of interest for<br />

the diagnosis and follow-up of NAFLD patients.<br />

Disclosures:<br />

Grace Wong - Speaking and Teaching: Echosens, Echosens, Echosens, Echosens<br />

Henry Lik-Yuen Chan - Advisory Committees or Review Panels: Gilead, MSD,<br />

Bristol-Myers Squibb, Roche, Novartis Pharmaceutical, Abbvie; Speaking and<br />

Teaching: Echosens<br />

Juliette Foucher - Board Membership: BMS, Gilead, Abbvie; Speaking and<br />

Teaching: BMS, Gilead, Janssen, AbbVie<br />

Victor de Ledinghen - Board Membership: Janssen, Gilead, BMS, Abbvie; Speaking<br />

and Teaching: AbbVie, Merck, BMS, Gilead<br />

Vincent W. Wong - Advisory Committees or Review Panels: AbbVie, Gilead,<br />

Janssen; Consulting: Merck, NovaMedica; Speaking and Teaching: Gilead,<br />

Echosens<br />

The following authors have nothing to disclose: Jean-Baptiste Hiriart, Julien<br />

Vergniol, Brigitte Le Bail, Anthony W. Chan, Faiza Chermak, Wassil Merrouche,<br />

Angel ML Chim<br />

2182<br />

Obesity fibrosis score: A new model to predict<br />

advanced fibrosis in morbidly obese patients with nonalcoholic<br />

fatty liver disease (NAFLD)<br />

Tavankit Singh, Srinivasan Dasarathy, Rocio Lopez, Arthur J.<br />

McCullough; Cleveland Clinic Foundation, Cleveland, OH<br />

Introduction: The prevalence of NAFLD is high in obese<br />

patients. Current diagnostic tests include a liver biopsy that<br />

is invasive and fibroscan whose cost and lack of validation in<br />

the morbidly obese population makes it necessary to evaluate<br />

subjects by other non-invasive measures. Although a number<br />

of prediction models (APRI, BARD, FIB-4 and NAFLD fibrosis<br />

scores) have been developed, they have not been validated<br />

in morbidly obese patients. We designed this study to develop<br />

a predictive model for advanced fibrosis in morbidly obese<br />

patients and to compare it with the other known models. Methods:<br />

Patients with morbid obesity (BMI >40 kg/m 2 or >35<br />

kg/m 2 with diabetes) and NAFLD proven by biopsy between<br />

2005-2014 were included and randomly split into two groups<br />

for model building (training cohort) and model validation (validation<br />

cohort) in a 2:1 ratio. To build the prediction model,<br />

multivariable logistic regression analysis was performed. Internal<br />

model validation was measured by area under receiver<br />

operating curve (AUROC). Results: 612 patients (408 and<br />

204 patients in training and validation cohorts respectively)<br />

were included in the study. Both cohorts had similar clinical<br />

profiles, laboratory values, patient demographics. 25%<br />

patients had advanced fibrosis. After analysis, the model to<br />

predict advanced fibrosis was defined as follows: z = -27303<br />

+(1.2671 if diabetic) +(0.0162XALP) - (0.0072X platelets)<br />

+(0.0165XAST) +0.6359 if AST/ALT ≥0.8 This value was then<br />

converted into a probability distribution with a value between<br />

0 and 100. AUROC was 0.756 and 0.808 in training and<br />

validation cohorts respectively. Scores of 44.2 and 14.6 were<br />

used to include or exclude the presence of advanced fibrosis<br />

and provided a sensitivity and specificity of 90% and 44%<br />

for the lower cut-off and 26% and 90% for the upper cut-off,<br />

respectively. This model was superior to APRI, FIB-4, BARD,<br />

NAFLD fibrosis scores in both the cohorts. Compared to FIB-<br />

4, the difference was not significantly different. However, the<br />

FIB-4 score could only identify 8/102 patients with advanced<br />

fibrosis thus making it less useful. Conclusion: In the obese<br />

population, the obesity fibrosis score is superior to the currently<br />

existing prediction models to detect advanced fibrosis and will<br />

permit careful selection of patients for further diagnostic evaluation.<br />

Prediction Scores for Advanced Fibrosis:ROC Analysis<br />

Disclosures:<br />

The following authors have nothing to disclose: Tavankit Singh, Srinivasan<br />

Dasarathy, Rocio Lopez, Arthur J. McCullough


1272A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2183<br />

Exercise reduces liver fat, visceral fat and circulating<br />

triglycerides, but not inflammatory markers in non-alcoholic<br />

steatohepatitis<br />

David Houghton, Christian Thoma, Kate Hallsworth, Kieren G.<br />

Hollingsworth, Christopher P. Day, Quentin M. Anstee, Michael I.<br />

Trenell; Institute of Cellular Medicine, Newcastle University, Newcastle<br />

Upon Tyne, United Kingdom<br />

Background: Non-alcoholic steatohepatitis (NASH) represents<br />

steatosis and ballooning degeneration, inflammation and fibrosis.<br />

Current treatments for NASH are limited, with no <strong>studies</strong><br />

to date reporting the effects of exercise in people with NASH.<br />

Studies in NAFLD show promising effects on liver lipid but the<br />

effect on inflammation and visceral fat, key mediators of fibrosis,<br />

are not known. Aims: Determine the effect of a 12-week<br />

exercise intervention on liver lipid and circulatory inflammation<br />

in people with NASH. Methods: 24 participants (mean age<br />

52 ± 14 years, BMI 33 ± 6) with histologically characterised<br />

NASH (NAFLD Activity Score ≥ 5) received either: exercise<br />

(n = 12) or continue standard care (n = 12) over 12 weeks<br />

and maintained baseline weight. Participants were not undergoing<br />

insulin sensitising treatment, dietary change or regular<br />

activity. Subjects with heart/kidney disease or in vivo ferrous<br />

material were excluded. Liver lipid content, subcutaneous and<br />

visceral adiposity were assessed using magnetic resonance<br />

techniques, inflammation, fibrosis markers, insulin sensitivity<br />

were assessed at baseline and at 12 weeks. Results: Resistance<br />

exercise produced a significant reduction in liver lipid content<br />

(-13 ± 24 vs. 6 ± 16%, P0.05), HOMA<br />

IR (2.3 ± 1.4 to 1.9 ± 0.8 vs. 1.9 ± 1.1 to 1.7 ± 1.1, P>0.05)<br />

or 2hour glucose levels or liver enzymes (ALT: 53 ± 25U/L to<br />

52 ± 18U/L vs. 81 ± 59U/L to 71 ± 52U/L; AST: 41 ± 14U/L<br />

to 45 ± 12U/L vs. 59 ± 41U/L to 58 ± 45U/L, P>0.05). Conclusion:<br />

This is the first study reporting the effects of exercise on<br />

liver lipid, body composition and circulating inflammation in<br />

patients with histologically defined NASH. Exercise produces<br />

a significant reduction in liver lipid, visceral fat and plasma triglycerides<br />

but no effect on body weight, abdominal adiposity,<br />

or circulatory markers of inflammation. These results suggest<br />

that exercise alone may be insufficient to target the mediators<br />

of NASH and warrants further exploration.<br />

Disclosures:<br />

Quentin M. Anstee - Advisory Committees or Review Panels: Genfit, Intercept,<br />

Raptor; Grant/Research Support: GSK; Speaking and Teaching: Abbott Laboratories<br />

The following authors have nothing to disclose: David Houghton, Christian<br />

Thoma, Kate Hallsworth, Kieren G. Hollingsworth, Christopher P. Day, Michael<br />

I. Trenell<br />

2184<br />

The link between intestinal microbiota, bile acids and<br />

non-alcoholic fatty liver disease<br />

Alice Y. Wang 1 , Marialena Mouzaki 2,3 , Bianca M. Arendt 4,5 ,<br />

Robert Bandsma 2,3 , Scott Fung 4,5 , Sandra E. Fischer 4,6 , Johane<br />

P. Allard 4,5 ; 1 Physiology and Experimental Medicine Program,<br />

Research Institute, The Hospital for Sick Children, Toronto, ON,<br />

Canada; 2 Division of Gastroenterology, Hepatology and Nutrition,<br />

The Hospital for Sick Children, Toronto, ON, Canada; 3 University<br />

of Toronto, Toronto, ON, Canada; 4 Toronto General Hospital,<br />

University Health Network, Toronto, ON, Canada; 5 Department of<br />

Medicine, University of Toronto, Toronto, ON, Canada; 6 Department<br />

of Laboratory Medicine and Pathobiology, University of<br />

Toronto, Toronto, ON, Canada<br />

Purpose: Intestinal microbiota (IM) may be implicated in the<br />

pathogenesis of non-alcoholic fatty liver disease (NAFLD)<br />

through their effects on bile acid (BA) homeostasis. To better<br />

elucidate the associations, this study characterizes the fecal<br />

BA profiles of patients with NAFLD and examines the relationship<br />

between IM and fecal BA composition. Methods: This was<br />

a prospective, cross-sectional study. The fecal BA profiles of<br />

adults with biopsy-confirmed NAFLD (non-alcoholic fatty liver:<br />

NAFL and non-alcoholic steatohepatitis: NASH) and healthy<br />

controls (HC) were examined. In a subset of subjects, IM composition<br />

was assessed. Clinical and laboratory data, stool and<br />

blood samples and 7-day food records were collected prior<br />

to liver biopsy. Fecal BAs were quantified by liquid chromatography-tandem<br />

mass spectrometry . Quantitative real-time<br />

polymerase chain reaction was used to measure total bacterial<br />

counts, Bacteroides/Prevotella (Bacteroidetes), C. leptum,<br />

C. coccoides, Bifidobacteria, Escherichia coli and Archaea<br />

in stool. Statistical analyses were conducted using SPSS v.22<br />

(IBM Corp 2013). Results: This study included 53 participants<br />

(25 HC, 11 NAFL, 17 NASH), 56.5% male, with a mean<br />

(±SD) age of 41.0 ± 10.9 years and a median (IQR) BMI of<br />

28.4 (23.4-32.4) kg/m 2 . Total fecal BA levels were higher<br />

in patients with NAFLD (NASH and NAFL combined) compared<br />

to HC (log values: 3.65 and 3.50 pmol/mg, respectively;<br />

p=0.021). The ratio of conjugated to unconjugated BA<br />

was not different between the groups. Patients with NAFLD<br />

had a higher ratio of primary to secondary BA compared to<br />

HC. Levels of unconjugated cholic acid (CA) were higher in<br />

both NASH and NAFL compared to HC (p=0.005, p=0.026,<br />

respectively), whereas chenodeoxycholic acid (CDCA) was<br />

higher in NASH vs. HC (p=0.011). The primary conjugated BA<br />

were not different between the groups; however, patients with<br />

NASH had higher levels of conjugated lithocholic acid (LCA)<br />

than patients with NAFL (p=0.035). In a subset of 29 patients<br />

(18 HC, 11 NAFLD) with IM data, patients with NAFLD had<br />

decreased levels of Bacteroidetes and C. leptum (p=0.041;<br />

p=0. 006), which remained significant after adjusting for BMI<br />

and percent of fat intake. Bacteroidetes levels were inversely<br />

correlated with glycoLCA (r=-0.550, p=0.002) and tauroLCA<br />

(r=-0.408, p=0.048). C. leptum levels were positively correlated<br />

with fecal LCA (r=0.526, p = 0.003) and inversely with<br />

CA (r=-0.669, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1273A<br />

2185<br />

Serum marker of inflammasome activity correlates with<br />

liver injury in nonalcoholic fatty liver disease and is<br />

influenced by genetic polymorphisms.<br />

Leon A. Adams 1,2 , Alexander Wree 3 , Phillip Melton 4 , Gary P.<br />

Jeffrey 1,2 , Helena Ching 2,1 , Bastiaan de Boer 5 , John K. Olynyk 6 ,<br />

Oyekoya T. Ayonrinde 6 , Trevor A. Mori 1 , Lawrence Beilin 1 , Patricia<br />

Price 7 , Craig E. Pennell 9 , Mohammed Eslam 8 , Jacob George 8 ,<br />

Ariel E. Feldstein 3 ; 1 Medicine and Pharmacology, University of<br />

Western Australia, Nedlands, WA, Australia; 2 Hepatology, Sir<br />

Charles Gairdner Hospital, Nedlands, WA, Australia; 3 Pediatrics,<br />

University of California, San Diego, San Diego, CA; 4 Center of<br />

Genetic Origins of Health and Disease, The University of Western<br />

Australia, Nedlands, WA, Australia; 5 Anatomical Pathology,<br />

PathWest, Nedlands, WA, Australia; 6 Gastroenterology, Fiona<br />

Stanley Hospital, Perth, WA, Australia; 7 School of Biomedical Sciences,<br />

Curtin University, Perth, WA, Australia; 8 Medicine, University<br />

of Sydney, Sydney, NSW, Australia; 9 Womens and Childrens<br />

Health, The University of Western Australia, Perth, WA, Australia<br />

Background and Aims: The NLRP3 inflammasome has been<br />

implicated in the pathogenesis of liver injury in nonalcoholic<br />

fatty liver disease (NAFLD). In humans, mutations of Nlrp3<br />

have functional consequences and result in auto-inflammatory<br />

syndromes. We examined whether single nucleotide polymorphisms<br />

(SNP’s) of the NLRP3 and CARD8 genes of the NLRP3<br />

inflammasome influenced serum inflammatory markers and<br />

fatty liver in a general population based cohort. Secondly, we<br />

examined serum caspase-1 as a novel marker of liver injury<br />

and determined the association between inflammasome SNP’s,<br />

serum caspase-1 and liver injury. Methods: Inflammatory markers<br />

(CRP, IP-10, TNFR1, TNFR2, IL18), liver enzymes and<br />

ultrasound diagnosed fatty liver were assessed in a population-based<br />

cohort (n=1046) of 17-year old adolescents. Serum<br />

caspase-1 activity and protein levels were determined in a<br />

second cohort (n=154) and sub-group (n=20) of subjects with<br />

biopsy proven NAFLD by flourometric assay (Abcam, Cambridge,<br />

MA) and western blot respectively. Liver histology was<br />

evaluated according to the NASH CRN scoring system. Eleven<br />

NLRP3 and CARD8 SNPs were genotyped in both cohorts and<br />

assessed for phenotypic associations. Results: In cohort 1, the<br />

rs2043211 SNP in the CARD8 gene was associated with serum<br />

levels of IP10 (p=0.002), IL18 (p=0.009), sTNFR1 (p=0.04),<br />

sTNFR2 (p=0.04), CRP (p=0.056) and a diagnosis of fatty<br />

liver (adj. odds ratio 1.65, 95% CI 1.09-2.48, dominant allelic<br />

model). Serum IP10 and TNFR2 levels but not SNP genotypes<br />

were positively correlated with hepatic aminotransaminase<br />

levels. In cohort 2, serum caspase-1 activity was significantly<br />

increased in the presence of hepatocellular ballooning, inflammation<br />

and fibrosis (p


1274A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2187<br />

Liver Fat and Inflammation Are Associated with Circulating<br />

PCSK9 Levels<br />

Paola Dongiovanni 1 , Massimiliano Ruscica 2 , Nicola Ferri 2 , Claudia<br />

Lanti 2 , Raffaela Rametta 1 , Anna Ludovica Fracanzani 1 , Paolo<br />

Magni 2 , Silvia Fargion 1 , Luca Valenti 1 ; 1 Internal Medicine, University<br />

of Milano, Fondazione IRCCS Ca Granda, Milano, Italy;<br />

2 University of Milan, Milan, Italy<br />

Background and aims: Hepatocellular fat accumulation defines<br />

nonalcoholic fatty liver disease (NAFLD), also known as hepatic<br />

steatosis. Lipids accumulation derives from dietary fatty acids,<br />

increased peripheral lipolysis due to adipose tissue insulin resistance<br />

and elevated hepatic de novo lipogenesis due to hyperinsulinemia.<br />

Altered secretion of very low-density lipoproteins<br />

(VLDL) and free cholesterol accumulation also mediate hepatocellular<br />

damage. However, NAFLD also confers increased risk<br />

of cardiovascular events independently of classic risk factors.<br />

Proprotein Convertase Subtilisin/Kexin type 9 (PCSK9) is an<br />

important player in lipoprotein metabolism. Cleaved PCSK9 is<br />

secreted by hepatocytes with VLDL and circulate with lipoproteins.<br />

Soluble PCSK9 inhibits the uptake of low-density lipoproteins<br />

(LDL) by targeting LDL receptor (LDL-R) to degradation, and<br />

loss-of-function variants are associated with reduced circulating<br />

cholesterol and protection from cardiovascular disease. However,<br />

it is not known whether hepatic fat content and inflammation<br />

modify PCSK9 levels. Aim was to evaluate whether hepatic<br />

fat content affect circulating PCSK9 levels in individuals, who<br />

underwent liver biopsy for suspected nonalcoholic steatohepatitis<br />

(NASH). Methods: We considered 203 patients of Italian<br />

ancestry who underwent biopsy for suspected NASH (n=126<br />

for persistently elevated serum ALT, n=77 for severe obesity).<br />

Serum PCSK9 levels were measured using a commercial ELISA<br />

kit (R&D Systems, Minneapolis, MN, USA). Results: In NAFLD<br />

patients, at univariate analysis PCSK9 levels were significantly<br />

associated with steatosis severity (estimate +0.14±0.05<br />

log ng/mL; p=0.006), inflammation (estimate +0.15±0.06;<br />

p=0.01), and fibrosis (estimate +0.13±0.04; p=0.001). At<br />

multivariate analysis, PCSK9 serum levels were significantly<br />

associated with steatosis grade (estimate +0.13±0.06 log ng/<br />

mL; p=0.03), age (estimate +0.008±0.004; p=0.03), BMI<br />

(estimate -0.016±0.006; p=0.01) independently of sex, and<br />

liver disease severity. In a subset of 30 obese patients, there<br />

was no significant association between circulating PCSK9 concentration<br />

and hepatic mRNA levels. Conclusions: Modulation<br />

of serum PCSK9 by post-transcriptional mechanisms might be<br />

involved in mediating the susceptibility to dyslipidemia and atherosclerosis<br />

progression in patients with NAFLD, with potential<br />

therapeutic implications.<br />

Disclosures:<br />

The following authors have nothing to disclose: Paola Dongiovanni, Massimiliano<br />

Ruscica, Nicola Ferri, Claudia Lanti, Raffaela Rametta, Anna Ludovica Fracanzani,<br />

Paolo Magni, Silvia Fargion, Luca Valenti<br />

2188<br />

Increasing severity of NAFLD is associated with gut dysbiosis<br />

and modification of the metabolic function of the<br />

gut microbiota<br />

Jerome Boursier 1,3 , Olaf Mueller 2 , Matthieu Barret 4 , Mariana V.<br />

Machado 5 , Lionel Fizanne 3 , Cynthia D. Guy 6 , John F. Rawls 2 , Lawrence<br />

David 2 , Gilles Hunault 3 , Frederic Oberti 1,3 , Paul Cales 1,3 ,<br />

Anna Mae Diehl 5 ; 1 Hepato-Gastroenterology Department, University<br />

Hospital, Angers, France; 2 Center for Genomics of Microbial<br />

Systems, Duke University, Durham, NC; 3 Laboratoire HIFIH<br />

EA3859, University, Angers, France; 4 UMR1345, INRA, Beaucouzé,<br />

France; 5 Division of Gastroenterology, Duke University<br />

Medical Center, Durham, NC; 6 Department of Pathology, Duke<br />

University, Durham, NC<br />

Background: Several animal <strong>studies</strong> have suggested the role of<br />

gut microbiota in Non-Alcoholic Fatty Liver Disease (NAFLD).<br />

However, data about gut dysbiosis in human NAFLD remains<br />

scarce in the literature, especially <strong>studies</strong> including the whole<br />

spectrum of NAFLD lesions. We aimed to evaluate the association<br />

between gut dysbiosis and the severity of NAFLD,<br />

i.e. Non-Alcoholic Steatohepatitis (NASH) and fibrosis, in a<br />

well-characterized cohort of adult NAFLD. Methods: 57 patients<br />

with biopsy-proven NAFLD were enrolled. The taxonomic composition<br />

of gut microbiota was determined using 16S ribosomal<br />

RNA sequencing of stool samples. Results: 30 patients<br />

had F0/1 fibrosis stage at liver biopsy (10 with NASH), and<br />

27 patients had significant F ≥2 fibrosis (25 with NASH).<br />

Bacteroides were significantly increased in NASH and F ≥2<br />

patients, whereas Prevotella were decreased. Ruminococcus<br />

were significantly higher in F ≥2 patients. Multivariate analysis<br />

including age, sex, metabolic syndrome, serum transaminases,<br />

gammaGT, ultrasensitive CRP, and gut microbiota data showed<br />

that Bacteroides abundance was independently associated<br />

with NASH and Ruminococcus with F≥2 fibrosis. Stratification<br />

according to the abundance of these 2 bacteria generated<br />

3 patient subgroups with increasing severity of NAFLD<br />

lesions (low NASH/low fibrosis, high NASH/low fibrosis, high<br />

NASH/ high fibrosis). Patients with a metabolic syndrome had<br />

more severe liver lesions and, in those patients, NAFLD severity<br />

increased according to the 3 gut microbiota subgroups. Based<br />

on predicted metagenomics profile, the severity of NAFLD was<br />

associated with significant microbiota enrichment in KEGG<br />

pathways involved in carbohydrate metabolism (starch and<br />

sucrose metabolism, pentose phosphate pathway, pentose and<br />

glucoronate interconversions, pyruvate metabolism, glyoxylate<br />

and dicarboxylate metabolism), lipid metabolism (lipid biosynthesis<br />

proteins, sphingolipid metabolism), and amino acid<br />

metabolism (arginine and proline metabolism, cyanoamino<br />

acid metabolism). Conclusion: NAFLD severity associates with<br />

gut dysbiosis and altered metabolic functions of the gut microbiota.<br />

Bacteroides is an independent predictor for NASH and<br />

Ruminococcus for fibrosis. Thus, gut microbiota analysis adds<br />

information to classical predictors of NAFLD severity and suggests<br />

novel metabolic targets for pre/pro-biotic therapies.<br />

Disclosures:<br />

Frederic Oberti - Speaking and Teaching: LFB, gore<br />

Paul Cales - Consulting: BioLiveScale<br />

The following authors have nothing to disclose: Jerome Boursier, Olaf Mueller,<br />

Matthieu Barret, Mariana V. Machado, Lionel Fizanne, Cynthia D. Guy, John F.<br />

Rawls, Lawrence David, Gilles Hunault, Anna Mae Diehl


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1275A<br />

2189<br />

Prevalence of non-alcoholic fatty liver disease (NAFLD)<br />

and its subtypes obtained by liver biopsy in patients<br />

with type 2 diabetes not selected by ultrasound and/or<br />

aminotransferase levels<br />

Lilian M. Silva 2 , Fatima F. Figueiredo 2,3 , Frederico F. Campos 4 ,<br />

Marcio Q. Miguez 2 , Valéria P. Lanzoni 5 , Marilia B. Gomes 6 , Carlos<br />

Terra 2 , Hugo Perazzo 1 , Renata M. Perez 2,7 ; 1 National Institute<br />

of Infectious Diseases Evandro Chagas, Rio de Janeiro, Brazil;<br />

2 Liver Unit, University of the State of Rio de Janeiro, Rio de Janeiro,<br />

Brazil; 3 Internal Medicine, Federal University of Rio de Janeiro,<br />

Rio de Janeiro, Brazil; 4 Pathology, University of the State of Rio<br />

de Janeiro, Rio de Janeiro, Brazil; 5 Pathology, Federal University<br />

of São Paulo, São Paulo, Brazil; 6 Diabetology, University of the<br />

State of Rio de Janeiro, Rio deJaneiro, Brazil; 7 D’Or Institute for<br />

Research and Education, Rio de Janeiro, Brazil<br />

Background: Non-alcoholic fatty liver disease (NAFLD) and<br />

its severe form, steatohepatitis (NASH), have been described<br />

as highly prevalent in patients with metabolic diseases. The<br />

aims were: (i) to estimate the prevalence of NAFLD and its<br />

spectrum by liver biopsy in type 2 diabetes patients; (ii) to<br />

quantify steatosis and fibrosis; (iii) to identify variables predictive<br />

of NASH and fibrosis. Patients and methods: Patients with<br />

type 2 diabetes were submitted to percutaneous liver biopsy<br />

regardless of alanine aminotransferase (ALT) levels or abdominal<br />

ultrasound results. The exclusion criteria were presence of<br />

other chronic liver disease or serious systemic diseases; coagulation<br />

disturbances; morbid obesity; alcohol abuse; use of<br />

hepatotoxic drugs; pregnancy; or refusal to participate. Slides<br />

of liver biopsy were evaluated by two independent pathologists<br />

using the NASH Activity Score. Factors associated with NASH<br />

and significant fibrosis (stage F≥2) were defined by logistic<br />

regression analysis. Results: 85 patients [82% female, mean<br />

age=54 yrs, mean BMI=31 Kg/m 2 , 61% with normal ALT,<br />

53% treated by insulin] were included. Median (range) liver<br />

specimen length was 30 (25-75) mm. Prevalence of NAFLD in<br />

type 2 diabetes was 92% [n=78/85], 54% (n=42) of whom<br />

with simple steatosis and 46% (n=36) with NASH. Severe steatosis<br />

(>33% of hepatocytes) was more prevalent in patients<br />

with NASH compared to those with simple steatosis (65% vs<br />

7%; p


1276A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Rajarshi Banerjee - Board Membership: Perspectum Diagnostics; Employment:<br />

Perspectum Diagnostics; Grant/Research Support: Perspectum Diagnostics; Patent<br />

Held/Filed: Perspectum Diagnostics Ltd, University of Oxford; Stock Shareholder:<br />

Perspectum Diagnostics<br />

Elizabeth M. Tunnicliffe - Patent Held/Filed: Perspectum Diagnostics; Stock Shareholder:<br />

Perspectum Diagnostics<br />

Stefan Neubauer - Board Membership: Perspectum Diagnostics; Patent Held/<br />

Filed: University of Oxford<br />

The following authors have nothing to disclose: Jane Collier, Lai Mun Wang,<br />

Fleming A. Kenneth, Jeremy F. Cobbold, Eleanor Barnes<br />

2191<br />

Non-invasive hepatic fibrosis markers and cardiometabolic<br />

risk among adults in the Framingham Heart Study<br />

Michelle T. Long 1,2 , Alison Pedley 2 , Joseph M. Massaro 3,2 , Udo<br />

Hoffmann 4 , Caroline S. Fox 2,5 ; 1 Gastroenterology, Boston Medical<br />

Center, Boston, MA; 2 National Heart, Lung, and Blood Institute’s<br />

Framingham Heart Study, Framingham, MA; 3 Biostatistics, Boston<br />

University School of Public Health, Boston, MA; 4 Radiology,<br />

Massachusetts General Hospital, Boston, MA; 5 Endocrinology,<br />

Brigham and Women’s Hospital, Boston, MA<br />

Background & Aims: Non-alcoholic fatty liver disease (NAFLD)<br />

is associated with an increased risk of cardiovascular related<br />

death, particularly in those with a high predicted risk for hepatic<br />

fibrosis. We determined (a) the prevalence of predicted hepatic<br />

fibrosis based on various non-invasive fibrosis markers and (b)<br />

the association of predicted hepatic fibrosis with cardiovascular<br />

risk factors. Methods: We conducted a cross-sectional<br />

study of 575 Framingham Heart Study participants with NAFLD<br />

based on computed tomography after excluding excess alcohol<br />

use. We estimated the prevalence of predicted hepatic fibrosis<br />

based on the aspartate aminotransferase (AST)/alanine aminotransferase<br />

(ALT) ratio, AST to platelet ratio index (APRI),<br />

and the NAFLD Fibrosis Score (NFS). Using multivariable logistic<br />

regression models, we examined the association between<br />

NAFLD with low, intermediate, or high risk for advanced fibrosis<br />

according to the NFS and various cardiometabolic risk factors.<br />

Results: Among participants with NAFLD, the prevalence<br />

of predicted hepatic fibrosis was 12%, 5%, and 32% based<br />

on the NFS, APRI, and AST/ALT ratio, respectively. Since the<br />

NFS predicted a high risk for advanced fibrosis most similar<br />

to our a priori hypothesis, we continued the analysis using<br />

this model to categorized participants as low, intermediate,<br />

or high risk for advanced fibrosis. In multivariable models,<br />

participants with NAFLD and a high risk for advanced fibrosis<br />

by the NFS had an average 6.87 mm Hg wider pulse pressure<br />

(95% CI 3.33-10.42 mm Hg; p=0.0002) and an increased<br />

odds of hypertension (OR 2.92 95% CI 1.35-6.34; p = 0.007)<br />

compared to those with NAFLD and low risk of fibrosis. There<br />

were no statistically significant differences between the systolic<br />

blood pressure, HDL cholesterol or triglycerides for those<br />

with NAFLD and a high risk of advanced fibrosis by the NFS<br />

compared to those with NAFLD and a low risk of advanced<br />

fibrosis. Conclusions: The AST/ALT ratio, APRI, and NFS give<br />

widely disparate predictions of hepatic fibrosis when applied<br />

to a community-based cohort. Participants with NAFLD and a<br />

high risk for advanced fibrosis based on the NFS had a wider<br />

pulse pressure and increased odds of hypertension. Whether<br />

modifying these risk factors impacts cardiovascular endpoints<br />

in NAFLD patients remains unknown and should be explored<br />

in future <strong>studies</strong>.<br />

Disclosures:<br />

Alison Pedley - Employment: Merck<br />

The following authors have nothing to disclose: Michelle T. Long, Joseph M.<br />

Massaro, Udo Hoffmann, Caroline S. Fox<br />

2192<br />

Macrophages (CD68+ Cells) and CD8+ T Cytotoxic Cells<br />

are Independently Associated with Advanced Fibrosis in<br />

Patients with Non Alcoholic Fatty Liver Disease (NAFLD)<br />

Azza Karrar 1 , Dinan Abdelatif 2 , Irfan Ali 2 , Yousef Fazel 1 , Pegah<br />

Golabi 1 , Mohamad Houry 1 , Zahra Younoszai 1 , Fanny Monge 2 ,<br />

Munkhzul Otgonsuren 1 , Zachary D. Goodman 2 , Lakshmi Alaparthi<br />

2 , Zobair M. Younossi 1,2 ; 1 Betty and Guy Beatty Center for Integrated<br />

Research, Inova Health System, Falls Church, VA; 2 Center<br />

for Liver Diseases, Inova Fairfax Hospital, Falls Church, VA<br />

Background and Aims: Macrophages (CD68+ cells) and CD8+<br />

T cytotoxic cells dominate the inflammatory infiltrate in NAFLD.<br />

The crosstalk between liver and adipose tissue with regard to<br />

inflammatory cells has not been studied in patients with NAFLD.<br />

We aimed to quantify macrophages and T Helpers cells in liver<br />

and adipose of patients with NAFLD. Methods: Liver biopsy<br />

samples (49) were stained with Hematoxylin and eosin (H&E)<br />

and Masson Trichrome for assessment of NAFLD/NASH and<br />

scored for fibrosis. Digitized images of CD3, CD4, CD8 and<br />

CD68 stained liver and adipose sections were acquired using<br />

The Scan scope XT. An IHC Nuclear Image Analysis algorithm<br />

was used to quantify the percentage of CD3+, CD4+ and<br />

CD8+ cells and a positive pixel count algorithm was used to<br />

quantify the CD68+ cells. Multivariate analysis was performed;<br />

p values ≤ 0.05 were considered to be significant. Results: The<br />

study cohort (N=49) included 33 patients with NASH NAFLD<br />

[Median(IQR) Age=49.00(42.00,55.00); Male=11(33%);<br />

White=25(75.8%); BMI(Kg/m 2 =49.22(40.27,54.67)],<br />

and 16 non NASH NAFLD [Age=45.00(31.00,51.50);<br />

Male=4(25.0%); White=14(87.5%); BMI Kg/<br />

m 2 =44.53(42.23,51.49)]. In univariate analysis, percentages<br />

of CD3+ cells [Odds Ratio (OR): 1.44(1.02, 2.04),<br />

p=0.041], CD8+ cells [OR: 1.29(1.06, 1.58), p=0.011], and<br />

CD68+ cells [OR: 6.39 (1.11, 36.9), p=0.038] in the liver<br />

are significantly associated with higher risk for NASH NAFLD.<br />

Additionally, a greater percentage of CD8+ T cells [OR:<br />

1.33(1.08-1.64), p=0.006] in liver is correlated with higher<br />

risk for pericellular fibrosis, while in adipose, a greater percentage<br />

of CD3+ cells [OR: 0.67 (0.49-0.92), p=0.012] and<br />

CD8+ cells [OR: 0.78 (0.60-1.01), p=0.058] are correlated<br />

with lower risk of pericellular fibrosis. In liver, it is also found<br />

that a greater percentage of CD8+ cells [OR: 1.33(1.10-1.59),<br />

p=0.002] and CD68+ cells [OR: 5.87(1.38-25.1), p=0.017]<br />

are associated with higher risk of advanced fibrosis. In multivariate<br />

analysis, in the liver, a greater percentage of CD8+<br />

cells [OR: 1.28 (1.04-1.58), p=0.020) are independently<br />

associated with higher risk for pericellular fibrosis and greater<br />

percentage of CD68+ cells [OR: 7.37 (1.58-34.4), p=0.011]<br />

are independently associated with higher risk of advanced<br />

fibrosis. Conclusion: We show that CD8+ T cells are strongly<br />

associated with pericellular fibrosis while CD68+ cells are<br />

strongly associated with advanced fibrosis. Therefore macrophages<br />

and CD8+ cells are key cells in NAFLD progression.<br />

Disclosures:<br />

Zachary D. Goodman - Consulting: Gilead Sciences, Abbvie; Grant/Research<br />

Support: Gilead Sciences, Fibrogen, Galectin Therapeutics, Intercept, Synageva,<br />

Conatus, Tobira, Exalenz<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Azza Karrar, Dinan Abdelatif,<br />

Irfan Ali, Yousef Fazel, Pegah Golabi, Mohamad Houry, Zahra Younoszai,<br />

Fanny Monge, Munkhzul Otgonsuren, Lakshmi Alaparthi


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1277A<br />

2193<br />

Usefulness of mac-2 binding protein glycosylation isomer<br />

(M2BPGi) for advanced fibrosis and hepatocellular<br />

carcinoma in patients with non-alcoholic fatty liver disease<br />

Erina Kumagai 1,2 , Masaaki Korenaga 1 , Keiko Korenaga 1 , Misuzu<br />

Ueyama 1,2 , Masatoshi Imamura 1 , Masaya Sugiyama 1 , Kazumoto<br />

Murata 1 , Naohiko Masaki 1 , Tatsuya Kanto 1 , Sumio Watanabe 2 ,<br />

Masashi Mizokami 1 ; 1 The research center for hepatitis and Immunology,<br />

National Center of Global Health and Medicine (NCGM)<br />

at Kohnodai, Ichikawa, Japan; 2 Department of Gastroenterology,<br />

Juntendo University School of Medicine, Bunkyo-ku, Hongo, Japan<br />

Background and Aim: Evaluation of liver fibrosis in non-alcoholic<br />

fatty liver disease (NAFLD) patients is important for<br />

identifying those who may progress to liver cirrhosis and hepatocellular<br />

carcinoma (HCC). Although liver biopsies are the<br />

gold standard for diagnosing NAFLD-associated liver fibrosis,<br />

there is an urgent need for a non-invasive method for estimating<br />

the stage of liver fibrosis in NAFLD patients.The serum<br />

M2BPGi value using a glycan-based immunoassay may provide<br />

an accurate and reliable method for assessing the liver<br />

fibrosis stage in not only chronic hepatitis C patients but also<br />

NAFLD patients; however, there is no data regarding its clinical<br />

usefulness and the association with HCC. The aim of this study<br />

was to assess the diagnostic performances of M2BPGi comparison<br />

with liver stiffness and Fib4-index for the diagnosis of<br />

advanced liver fibrosis and HCC in patients with NAFLD. Methods:<br />

A total of 358 consecutive patients (mean Age [ys]:63,<br />

male [%]:43, BMI [kg/m2]:26.3, and HCC n [%]: 25[7])<br />

with NAFLD who underwent ultrasonography were prospectively<br />

enrolled in this study. M2BPGi (COI), Fib-4 index, and<br />

liver stiffness by FibroScan (kPa) were measured on the same<br />

day. Advanced fibrosis group was determined by cut of level<br />

in each marker from previous reports. (M2BPGi>1.46 COI,<br />

Fib4 >2.67, and FibroScan>9.8kPa) Liver stiffness measurement<br />

(LSM) failure was defined as zero valid shots; unreliable<br />

examination was defined as


1278A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2195<br />

Combined effects of the prosteatotic TM6SF2 and<br />

PNPLA3 mutations on NALFD severity: multicentre biopsy-based<br />

study<br />

Marcin Krawczyk 9,1 , Monika Rau 2 , Jörn Schattenberg 3 , Heike Bantel<br />

4 , Anita Pathil 5 , Münevver Demir 6 , Johannes Kluwe 7 , Tobias<br />

Boettler 8 , Frank Lammert 9 , Andreas Geier 2 ; 1 Laboratory of Metabolic<br />

Liver Diseases, Department of General, Transplant and Liver<br />

Surgery, Medical University of Warsaw, Warsaw, Poland; 2 Division<br />

of Hepatology, Department of Medicine II, University Hospital<br />

Wuerzburg, Wuerzburg, Germany; 3 I. Department of Medicine,<br />

University Medical Center Mainz, Johannes Gutenberg University,<br />

Mainz, Germany; 4 Department of Gastroenterology, Hepatology<br />

and Endocrinology, Hannover Medical School, Hannover,<br />

Germany; 5 Department of Internal Medicine IV, Gastroenterology<br />

and Hepatology, University of Heidelberg, Heidelberg, Germany;<br />

6 Clinic for Gastroenterology and Hepatology, University Hospital<br />

of Cologne, Cologne, Germany; 7 I. Department of Medicine, Hamburg<br />

University Medical Center, Hamburg, Germany; 8 Department<br />

of Medicine II, University Hospital Freiburg, Freiburg, Germany;<br />

9 Department of Medicine II, Saarland University Medical Center,<br />

Homburg, Germany<br />

Introduction: The PNPLA3 (adiponutrin) mutation p.I148M<br />

represents the common genetic risk factor for non-alcoholic<br />

fatty liver disease (NAFLD) and progressive liver fibrosis. Lately<br />

a second prosteatotic variant, TM6SF2 p.E167K, has been<br />

detected (Nat Genet 2014). It remains unclear if this mutation,<br />

alike the PNPLA3 p.I148M variant, increases the risk of<br />

liver fibrosis. In the current study we analyzed the combined<br />

effects of these variants on NAFLD severity in a large cohort<br />

of patients recruited at eight German tertiary referral centres.<br />

Patients and methods: After exclusion of acute or chronic liver<br />

diseases other than NAFLD, 514 patients (age 16 – 88 years,<br />

239 men) were included. In 309 patients liver biopsies were<br />

performed, which were examined by pathologists blinded to<br />

genotyping results. PCR-based assays were used to genotype<br />

the PNPLA3 (rs738409) and TM6SF2 (rs58542926) variants.<br />

Results: The genotype frequencies of the PNPLA3 p.I148M<br />

([CC] = 215, [CG] = 223, [GG] = 76) and the TM6SF2<br />

p.E167K ([CC] = 411, [CT] = 94, [TT] = 9) mutations did not<br />

deviate from Hardy-Weinberg equilibrium. One copy of the<br />

prosteatotic PNPLA3 and TM6SF2 alleles was detected in 58%<br />

and 20% of NAFLD patients, respectively. Patients carrying<br />

the PNPLA3 p.I148M or TM6SF2 p.E167K risk alleles had<br />

significantly (both P < 0.01) increased serum ALT and AST<br />

activities. The PNPLA3 risk genotype [GG] (OR = 2.15; 95%<br />

CI 1.21 - 3.87, P = 0.007), but not the TM6SF2 genotype was<br />

more frequent in individuals scheduled for liver biopsy. Among<br />

biopsied individuals, a total of 149 presented with hepatic steatosis<br />

grades S2 or S3, and fibrosis stage > 1 was detected in<br />

77 patients. The TM6SF2 variant increased the risk of developing<br />

steatosis grade S2 - S3 (OR = 1.52, P = 0.04) but did not<br />

affect fibrosis stage. The PNPLA3 genotype was, in turn, associated<br />

with both steatosis grades S2 - S3 (OR = 1.94, P < 0.001)<br />

and increased hepatic fibrosis (OR = 2.24, P < 0.001). In<br />

patients with the PNPLA3 genotype [GG], the presence of variant<br />

TM6SF2 further increased serum aminotransferase activities<br />

(both P < 0.05) and tended to aggravate hepatic steatosis (P<br />

= 0.09). Conclusions: Our results demonstrate that the PNPLA3<br />

and TM6SF2 variants are associated with increased liver injury<br />

as reflected by serum surrogate and histopathological markers.<br />

The TM6SF2 variant seems to modulate predominantly hepatic<br />

fat accumulation, while the PNPLA3 mutation confers risk of<br />

increased steatosis and fibrosis.<br />

Disclosures:<br />

The following authors have nothing to disclose: Marcin Krawczyk, Monika Rau,<br />

Jörn Schattenberg, Heike Bantel, Anita Pathil, Münevver Demir, Johannes Kluwe,<br />

Tobias Boettler, Frank Lammert, Andreas Geier<br />

2196<br />

Weight Loss Results in Significant Improvements in<br />

Quality of Life for Patients with Nonalcoholic Fatty Liver<br />

Disease<br />

Elliot B. Tapper, Michelle Lai; Gastroenterology, Beth Israel Deaconess<br />

Medical Center, Boston, MA<br />

Background: Nonalcoholic Fatty Liver Disease (NAFLD) is<br />

highly prevalent and associated with decreased quality of life<br />

(QOL). Longitudinal QOL data are lacking. Methods: We prospective<br />

enrolled patients with NAFLD from 2009-2014. All<br />

patients received a liver biopsy, transient elastography, lifestyle<br />

assessment, blood tests and QOL tools including the Chronic<br />

Liver Disease Questionnaire (CLDQ), a validated health-related<br />

quality of life measurement. All patients were followed<br />

for 6 months at which time the patients were re-examined and<br />

re-administered the CLDQ. Our primary outcome was overall<br />

QOL and our principle exposure variable was at least a 5%<br />

decrease in weight. Results: 151 patients were followed for 6<br />

months with complete biochemical and QOL data. The cohort<br />

included 91 (60%) men, aged 51.5 + 12.6 years, 46 (30%)<br />

of whom were diabetic. 30 (21%) had advanced fibrosis or<br />

cirrhosis on biopsy and 67 (47%) had nonalcoholic steatohepatitis<br />

(NASH), as defined by a NAFLD activity score (NAS) ><br />

4. Overall, 47 (31%) patients achieved at least a 5% reduction<br />

in weight. Neither advanced fibrosis (20% vs 22%) nor NASH<br />

(44% vs 54%, p = 0.28) were predictive of weight loss. QOL<br />

improved significantly for patients who achieved at least a<br />

5% reduction in weight, especially in nondiabetic patients,<br />

those with NASH, and without significant fibrosis. The cohort’s<br />

baseline total CLDQ value was 5.6 (4.8 – 6.2). Those who<br />

achieved at least a 5% reduction in weight had a 0.45 (95%<br />

CI:0.24 – 0.66, p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1279A<br />

coholic Fatty Liver Disease (NAFLD), where the average BMI<br />

is higher than in Europe and Asia, is limited. We examined<br />

predictors of liver stiffness measurement (LSM) change and<br />

the effect of a combined testing strategy employing NAFLD<br />

Fibrosis Score (NFS) and VCTE. Methods: A prospective cohort<br />

of 161 biopsy-proven NAFLD patients evaluated with VCTE<br />

using an M probe and NFS at baseline and a repeat VCTE<br />

at 6 months. Unreliable VCTE readings were defined as a<br />

failure to obtain 10 valid measurements or an interquartile<br />

range (IQR) > 30%. Results: 121 (75.2%) patients had reliable<br />

VCTE. The mean body mass index (BMI) of the total cohort<br />

was 33.5 km/m 2 ; 32.2 and 37.2 in those with reliable and<br />

unreliable VCTE (p < 0.0001). The optimal BMI cutoff for an<br />

unreliable exam is BMI 37 kg/m 2 (odds ratio 6.76 95% CI<br />

(3.06 – 15.4). The area under the receiver operating curve<br />

(AUROC) for the detection of F3-F4 fibrosis by VCTE is 0.92<br />

(95% CI: 0.85–0.96) with an optimal LSM cutoff of 9.9 kPa<br />

(sensitivity 95%, specificity 77%). Conversely, the AUROC for<br />

the association between NFS and advanced fibrosis for the<br />

whole cohort was 0.77 (95% CI:0.63–0.97). In the detection<br />

of F3-F4 fibrosis in patients with reliable VCTE, the AUROC for<br />

VCTE is superior to that of NFS (0.92 vs. 0.77, p=0.01). VCTE<br />

led to a lower F3-F4 fibrosis misclassification rate (13% versus<br />

17%) A low risk VCTE result was obtained for 63 patients with<br />

F0-F2 fibrosis. However, 13 of these patients had indeterminate<br />

NFS; a combined strategy correctly classifies fewer (50)<br />

low-risk patients. The median LSM improvement after 6 months<br />

was 11.1% (IQR= -13.0% - 37.8%) and 37 patients (45%)<br />

experienced a 20% improvement. A 5% weight loss was associated<br />

with improved LSM (p = 0.009). Conclusion: Reliable<br />

VCTE exams are useful to rule out advanced fibrosis in the U.S.<br />

NAFLD patients. However, 25% have unreliable <strong>studies</strong> using<br />

an M probe. Weight loss improves liver stiffness in the short<br />

term, likely by addressing steatosis and not fibrosis. Further<br />

data using the XL probe and adjustment for hepatic steatosis<br />

will be important.<br />

Disclosures:<br />

The following authors have nothing to disclose: Elliot B. Tapper, Michelle Lai<br />

2198<br />

Functional inhibition of spleen tyrosine kinase ameliorates<br />

different stages of alcoholic liver disease in mice<br />

Terence N. Bukong, Arvin Iracheta-Vellve, Aditya Ambade, Karen<br />

Kodys, Gyongyi Szabo; Medicine, University of Massachusetts<br />

Medical School, Worcester, MA<br />

Background: The spectrum of alcoholic liver disease (ALD)<br />

includes steatosis triggered by acute alcohol binge drinking,<br />

chronic alcoholic hepatitis and cirrhosis, the latter a leading<br />

cause of mortality with limited therapies available. Because<br />

alcohol targets different signaling pathways in the liver, identification<br />

of a master regulatory target capable of modulating<br />

multiple signaling processes is attractive. The aim of this study<br />

was to evaluate the role of spleen tyrosine kinase (SYK), a<br />

non-receptor tyrosine kinase, given its central modulatory role<br />

of multiple pro-inflammatory signaling pathways previously<br />

identified in the pathomechanism of ALD. Methods: Three different<br />

models of ALD were assessed in mice: 3-day binge,<br />

10-day acute-on-chronic and 5 weeks chronic alcohol Lieber-DeCarli<br />

diet. Some mice received daily dose of SYK inhibitor<br />

(R406) [5-10mg/kg]. Liver and serum samples were assessed<br />

by western blot, ELISA, EMSA, qPCR and histology. Results:<br />

Liver Syk protein expression was increased by acute-on-chronic<br />

and chronic but not acute alcohol administration. However,<br />

we found a significant elevation of activated Syk and its downstream<br />

targets, phosphorylated(p)-SYK Y525/526 , p-ERK1/2 and<br />

p-p38, in livers of alcohol-fed mice compared to controls in<br />

all 3 models. In all models, alcohol administration led to liver<br />

injury and inflammation demonstrated by significant increases<br />

in ALT, p-ERK1/2, NF-κB binding, TNF-α, MCP1, and IL-1β protein<br />

expression compared to controls. In vivo administration of<br />

a functional SYK inhibitor attenuated alcohol-induced hepatic<br />

injury and inflammation indicated by significantly reduced ALT,<br />

p-ERK1/2, p-p38, NF-κB binding, TNF-α, MCP1, E-Selectin,<br />

IL-1β expression compared to vehicle treated ethanol-fed mice.<br />

The protective effect of SYK inhibition on liver inflammation<br />

was present in all 3 models; specifically, neutrophil activation<br />

markers, Ly6G, MPO and E-Selectin, were decreased in the<br />

acute-on-chronic model and macrophage activation (F4/80<br />

and CD68) was attenuated in chronic alcohol feeding. Furthermore,<br />

SYK inhibition resulted in a significant decrease in<br />

hepatic steatosis indicated by Oil Red-O staining, triglyceride<br />

levels and increases in UCP1 and PRDM16 expression in the<br />

liver suggesting that SYK-dependent signaling is involved in<br />

alcoholic steatosis at different levels, including lipogenesis and<br />

lipid metabolism. Conclusions: Our data demonstrate a novel<br />

functional role of SYK in modulating liver inflammation and<br />

steatosis at different stages of alcoholic liver damage. This<br />

study highlights SYK as a potential multifunctional therapeutic<br />

target to suppress inflammation and steatosis in alcoholic liver<br />

disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: Terence N. Bukong, Arvin Iracheta-Vellve,<br />

Aditya Ambade, Karen Kodys, Gyongyi Szabo<br />

2199<br />

The glycosylation marker M2BPGi predicts the development<br />

of hepatocellular carcinoma in nonalcoholic<br />

steatohepatitis<br />

Miwa Kawanaka 1 , Hideyuki Hyogo 2 , Masahiko Koda 3 , Toshihide<br />

Shima 5 , Yuichi Hara 6 , Hiroshi Tobita 4 , Ken Nishino 1 , Akira Hiramatsu<br />

2 , Shuichi Sato 4 , Kazuaki Chayama 2 , Hirofumi Kawamoto 1 ,<br />

Takeshi Okanoue 5 , Keisuke Hino 6 ; 1 General Intrenal medicine2,<br />

Kawasaki-Hostital,Kawasaki Medical school, Okayama-city,<br />

Japan; 2 Medicine and Molecular Sciences,Graduate School of<br />

Biomedical Sciences, Hiroshima University, Hiroshima, Japan;<br />

3 Department of Multidisciplinary Internal Medicine, Tottori University<br />

School of Medicine, Yonago, Japan; 4 Gastroenterology<br />

and Hepatology, Shimane University, School of Medicine, Izumo,<br />

Japan; 5 Saiseikai Suita Hospital, Suita city, Japan; 6 Hepatology<br />

and Pancreatology, Kawasaki Medical School, Kurashiki city,<br />

Japan<br />

[Objectives] The number of non-B non-C hepatocellular carcinoma<br />

(HCC) cases has increased rapidly in recent years,<br />

and the increasing in the incidence of nonalcoholic steatohepatitis<br />

(NASH) has been an associated factor. Therefore,<br />

non-invasive methods are important to detect NASH patients<br />

with a high carcinogenesis risk.The present study examined<br />

the utility of M2BPGi as an indicator of liver carcinogenesis<br />

in NASH patients.[Subjects and methods] In this multicenter<br />

study, 63 patients diagnosed as NASH on the basis of hepatic<br />

histological findings and detected to have HCC (NASH-HCC)<br />

were compared with 280 control patients who, on the basis of<br />

liver biopsy findings, had not developed HCC for 9.8 years<br />

after a NASH diagnosis (NASHnon-HCC) in terms of age,<br />

sex, body mass index, ALT, AST, GTP, AST/ALT ratio, platelet<br />

count, homeostatic assessment model of insulin resistance,<br />

iron, ferritin, and high-sensitivity C-reactive protein, hyaluronic<br />

acid, type IV collagen 7S, and P-III-P, glycosylation markers<br />

(M2BPGi), and fibrosis-4 (FIB4) index, which contribute to HCC<br />

development.[Results] Patients in the NASH-HCC group were


1280A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

older, more likely to be male, and had a higher incidence of<br />

diabetes than those in the NASH non-HCC group. The NASH-<br />

HCC group had more patients with advanced liver fibrosis. The<br />

NASH-HCC group had a lower ALT level, platelet count, and<br />

albumin level and significantly higher fibrosis marker levels,<br />

ALT/AST ratio, M2BPGi, and FIB4 index than the NASH non-<br />

HCC group. M2BPGi was 1.1 in the NASH non-HCC group<br />

and 3.4 in the NASH HCC group (P < 0.0001). M2BPGi<br />

levels were examined at fibrosis stage 3 (1.1 vs. 2.1, P <<br />

0.005) and 4 (2.6 vs. 4.6, P < 0.02); HCC developed in<br />

patients with higher M2BPGi levels. In multivariate analysis,<br />

age, sex, fibrosis stage, and M2BPGi were independent factors<br />

contributing to NASH-HCC. An age of ≥67 years (odds<br />

ratio: 10.5; 95% CI [4.5–26.2]), male sex (odds ratio: 6.935;<br />

95% CI [2.8–17.2]), fibrosis stages 2–4 (stage 2: odds ratio:<br />

2.4 [0.7–10.4]; stage 3: odds ratio: 2.6 [0.7–10.8]; stage<br />

4: odds ratio: 18.3 (4.9–82), P < 0.0001) versus stages 0-1,<br />

and M2BPGi levels of 1.25–2.66 (odds ratio: 2.0 [0.8–5])<br />

and ≥2.67 (odds ratio 6.3 [1.7–24.2]) were associated with<br />

a higher risk of carcinogenesis. Thus, the risk was markedly<br />

elevated in patients with M2BPGi levels of ≥2.67. The rate of<br />

carcinogenesis at 10 years was 60% in patients with M2BPGi<br />

levels ≥2.67 and 15% in those with M2BPGi levels ≤2.67.<br />

[Conclusions] HCC risk was higher in those with non-alcoholic<br />

fatty liver disease who were men, aged ≥67 years, or had<br />

M2BPGi levels ≥2.67. Rourine follow-up using imaging tests is<br />

essential in these patients.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Asuka, Bayer, Daiichi Sankyo, Dainippon Sumitomo, Eisai, Janssen,<br />

Kowa, Mitsubishi Tanabe, MSD, Eli Lily, Nippon Kayaku, Nippon Shinyaku,<br />

Otsuka, Roche, Takeda, Toray, Torii, Tsumura, Zeria; Speaking and Teaching:<br />

Eisai, Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Asuka, Bayer, BMS,<br />

Chugai, Daiichi Sankyo, Dainippon Sumitomo, J&J, Jimro, Miyarisan, MSD,<br />

Nihon Kayaku, Olympus<br />

The following authors have nothing to disclose: Miwa Kawanaka, Hideyuki<br />

Hyogo, Masahiko Koda, Toshihide Shima, Yuichi Hara, Hiroshi Tobita, Ken<br />

Nishino, Akira Hiramatsu, Shuichi Sato, Hirofumi Kawamoto, Takeshi Okanoue,<br />

Keisuke Hino<br />

2200<br />

Improvement in Hepatic Steatosis in Nonalcoholic Steatohepatitis<br />

in Response to Synbiotic Supplementation<br />

Silvia M. Ferolla, Claudia A. Couto, Geyza N. Armiliato, Luciana<br />

C. Silva, Erika C. Lima, Quelson C. Lisboa, Flaviano S. Martins,<br />

Maria de Lourdes A. Ferrari, Aloisio S. Cunha, Cristiano A.<br />

Pereira, Teresa C. Ferrari; Federal University of Minas Gerais, Belo<br />

Horizonte, Brazil<br />

Background: Nonalcoholic fatty liver disease (NAFLD) is one<br />

of the most prevalent chronic liver diseases in the world, which<br />

can progress to a severe form of nonalcoholic steatohepatitis<br />

(NASH), cirrhosis and even hepatocellular carcinoma. Oral supplementation<br />

with a synbiotic has been proposed as an effective<br />

treatment of NAFLD as these agents influence the gut-liver<br />

axis improving inflammatory parameters. Objective: We aim<br />

to evaluate the effects of supplementation with a synbiotic on<br />

the grade of hepatic steatosis, intestinal bacterial overgrowth,<br />

and serum levels of lipopolysaccharide (LPS) and adiponectin<br />

in NASH patients. Design: In a randomized, controlled clinical<br />

trial, 50 patients with biopsy-proven NASH were supplemented<br />

twice daily for three months with a synbiotic or not receive any<br />

supplementation. Both groups were advised to follow an energy-balanced<br />

diet and physical activity recommendations. The<br />

anthropometric and metabolic parameters, grade of steatosis<br />

(evaluated by magnetic resonance [MR] techniques), glucose<br />

hydrogen breath test, and levels of LPS and adiponectin were<br />

evaluated before and after the treatment. Results: At the end of<br />

the study, the following significant differences were observed in<br />

the synbiotic group when compared to the control individuals:<br />

decrease in the hepatic proton density fat fraction measured<br />

by MR (-3.4%; p=0.027); loss of body weight (-1.3kg; 1.5%;<br />

p=0.06); reduction in BMI (-0.4kg/m 2 ; 1.2%; p=0.005) and<br />

reduction in the waist circumference measurement (-1.9cm;<br />

1.8%;p=0.001). Food consumption (total daily calorie intake)<br />

was not different between groups. Conclusions: Synbiotic supplementation<br />

in addition to nutritional counseling seems superior<br />

to nutritional counseling alone for the treatment of NAFLD,<br />

as the intervention group presented attenuation of steatosis,<br />

and reduction of body weight, BMI and waist circumference.<br />

Whether these effects will be sustained with longer treatment<br />

durations remains to be determined.<br />

Disclosures:<br />

The following authors have nothing to disclose: Silvia M. Ferolla, Claudia A.<br />

Couto, Geyza N. Armiliato, Luciana C. Silva, Erika C. Lima, Quelson C. Lisboa,<br />

Flaviano S. Martins, Maria de Lourdes A. Ferrari, Aloisio S. Cunha, Cristiano A.<br />

Pereira, Teresa C. Ferrari<br />

2201<br />

Cross-sectional and longitudinal analysis of advanced<br />

MRI-derived measures of liver fat and volume in the<br />

treatment of nonalcoholic steatohepatitis: A secondary<br />

analysis of an RCT<br />

Steven C. Lin 1 , Elhamy Heba 2 , Ricki Bettencourt 1,3 , Claude B.<br />

Sirlin 2 , Rohit Loomba 1,3 ; 1 NAFLD Translational Research Unit, Division<br />

of Gastroenterology, University of California, San Diego, La<br />

Jolla, CA; 2 Liver Imaging Group, Department of Radiology, University<br />

of California, San Diego, La Jolla, CA; 3 Division of Epidemiology,<br />

Department of Family and Preventive Medicine, University of<br />

California, San Diego, La Jolla, CA<br />

Background: Magnetic resonance imaging (MRI)-derived measures<br />

of liver fat and volume are emerging as accurate, non-invasive,<br />

and validated imaging biomarkers in patients with<br />

biopsy-proven nonalcoholic steatohepatitis (NASH). However,<br />

little is known about the changes in these measures longitudinally,<br />

and their association with each other in a NASH clinical<br />

trial. Aim: This study aims to examine the cross-sectional<br />

relationships between MRI-derived proton density fat fraction<br />

(PDFF, %), total liver volume (TLV, mL), and total liver fat index<br />

(TLFI, % mL) at two time points in a NASH randomized clinical<br />

trial. Methods: This is a secondary analysis of a randomized,<br />

double-blind, placebo-controlled trial (MOZART) in which 50<br />

patients with biopsy-proven NASH were randomized to either<br />

oral ezetimibe 10mg daily (n = 25) vs. placebo (n = 25) over<br />

24 weeks for the treatment of NASH (Loomba et al. Hepatology<br />

2014). Baseline and post-treatment anthropometrics, biochemical<br />

profiling, MRI, and liver biopsies were obtained. PDFF, TLV,<br />

and TLFI were obtained and calculated from MRI scans captured<br />

at these two time points. Results: At baseline, mean PDFF<br />

correlated strongly with TLFI (Spearman’s ρ = 0.94, n = 45,<br />

P


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1281A<br />

trial, PDFF and TLV were both strongly correlated with TLFI. An<br />

increase in hepatic steatosis (as measured by PDFF) is associated<br />

with an increase in hepatomegaly, as measured by TLV.<br />

Our findings indicate that MRI-derived measures of liver fat and<br />

volume can be used both cross-sectionally and longitudinally as<br />

imaging biomarkers in a NASH trial. The varied strengths of<br />

correlations between these MRI-derived measures suggest that<br />

the use of all three measures provides a more comprehensive<br />

assessment of changes in liver fat burden and size, rather than<br />

the use of only one measure in isolation. (Clinicaltrials.gov<br />

NCT01766713)<br />

Disclosures:<br />

Claude B. Sirlin - Grant/Research Support: GE, Pfizer, Bayer, Guerbet, Siemens<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

The following authors have nothing to disclose: Steven C. Lin, Elhamy Heba,<br />

Ricki Bettencourt<br />

2202<br />

Importance of reduction in serum alanine aminotransferase<br />

and the effect of PNPLA3 rs738409<br />

polymorphism for preventing the histological disease<br />

progression of nonalcoholic steatohepatitis in Japanese<br />

patients: a multi-center, retrospective study<br />

Yuya Seko 1 , Yoshio Sumida 1 , Yuichiro Eguchi 2 , Hideyuki Hyogo 3 ,<br />

Saiyu Tanaka 4 , Kohjiroh Mori 4 , Keizo Anzai 2 , Kazuaki Chayama<br />

5 , Yoshito Itoh 1 ; 1 Kyoto Prefectural University of Medicine,<br />

Kyoto, Japan; 2 Division of Hepatology and Internal medicine,<br />

Saga Univercity, Saga, Japan; 3 JA Hiroshima General Hospital,<br />

Hiroshima, Japan; 4 Center for Digestive and Liver Diseases,, Nara<br />

City Hospital, Nara, Japan; 5 Department of Gastroenterology and<br />

Metabolisim,, Hiroshima University Hospital, Hiroshima, Japan<br />

BACKGROUND & AIMS: Some environmental factors and<br />

genetic variation were reported to associate with histological<br />

progression in patients with nonalcoholic fatty liver disease<br />

(NAFLD). Our aim is to determine predictors, including PNPLA3<br />

rs738409 (encoding the I148M variant), of histological progression<br />

in Japanese patients with biopsy-proven NASH.<br />

METHODS: This multi-center retrospective cohort study enrolled<br />

140 patients with NASH who underwent serial liver biopsies.<br />

Histological evaluation included NASH activity score (NAS)<br />

and liver fibrosis. The median interval between initial and second<br />

liver biopsies was 900 days. ALT response was defined<br />

as a decrease to ≤40 U/L or by ≥30% of baseline. RESULTS:<br />

Of 140 patients, NAS was ameliorated in 50.7%, deteriorated<br />

in 20.0%, and remained unchanged in 29.3%. Liver fibrosis<br />

was improved in 22.9% of patients, progressed in 22.9%, and<br />

remained stable in 54.3%. The PNPLA3 rs738409 genotype<br />

frequencies were CC 0.29, CG 0.42, and GG 0.28. Multivariate<br />

analysis identified woman (hazard ratio [HR], 5.36;<br />

p=0.021) and ALT non-response (HR, 6.62; p=0.008) as predictors<br />

of deterioration of NAS. ALT non-response (HR, 2.60;<br />

p=0.043) was also selected as a predictor of progression of<br />

liver fibrosis. The average annual rate of fibrosis was -0.06<br />

stages/year in ALT responders, increasing to 0.13 stages/year<br />

in ALT non-responders. Carriage of G allele was associated<br />

with deterioration of NAS (p=0.045), but not associated with<br />

fibrosis progression. ALT responders were found in 78.1% of<br />

CC homozygotes, 66.7% of CG and 53.3% of GG. The ALT<br />

responders frequencies in PNPLA3 CC was 78.1%, in CG was<br />

66.7%, and in GG was 53.3%. Carriage of G allele was tend<br />

to associated with ALT non-responders (p=0.070). CONCLU-<br />

SIONS: A lack of reduction in serum ALT level was a predictor<br />

for histological progression in patients with NASH. Serum ALT<br />

should be strictly controlled to prevent liver histological progression<br />

in patients with NASH, and ALT response was associated<br />

with PNPLA3 rs738409 G allele.<br />

Disclosures:<br />

Kazuaki Chayama - Consulting: AbbVie; Grant/Research Support: Ajinomoto,<br />

Astellas, Torii, Tsumura, Aska, Bayer, Zeria, Daiichi Sankyo, Dainippon Sumitomo,<br />

Eisai, Eli Lily, Janssen, Kowa, Mitsubishi Tanabe, MSD, Nippon Kayaku,<br />

Nippon Shinyaku, Otsuka, Roche, Takeda, Toray; Speaking and Teaching:<br />

Ajinomoto, AbbVie, Abott, Astellas, AstraZeneca, Aska, Bayer, BMS, Chugai,<br />

Daiichi Sankyo, Dainippon Sumitomo, Eisai, J & J, Jimro, Miyarisan, MSD, Nihon<br />

Kayaku, Olympus<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

The following authors have nothing to disclose: Yuya Seko, Yoshio Sumida,<br />

Yuichiro Eguchi, Hideyuki Hyogo, Saiyu Tanaka, Kohjiroh Mori, Keizo Anzai<br />

2203<br />

First degree relatives of subjects with type 2 Diabetes<br />

Mellitus get early and severe NASH related cirrhosis<br />

Ajeet S. Bhadoria 1 , Chandan K. Kedarisetty 2 , Ankit Bhardwaj 1 ,<br />

Guresh Kumar 1 , Tanmay S. Vyas 2 , Jaya Benjamin 3 , Varsha<br />

Shasthry 3 , Manoj Kumar 2 , Shiv K. Sarin 2 ; 1 Clinical Research, Institute<br />

of Liver and Biliary Sciences, New Delhi, India; 2 Hepatology,<br />

Institute of Liver and Biliary Sciences, New Delhi, India; 3 Clinical<br />

Nutrition, Institute of Liver and Biliary Sciences, New Delhi, India<br />

Background and Aims: Familial aggregation of type 2 diabetes<br />

mellitus (T2DM) and NAFLD is well documented. However,<br />

there is scarcity of data on such association with NASH related<br />

cirrhosis. We investigated the influence of family history (FH) of<br />

T2DM on age at diagnosis and severity at presentation among<br />

NASH related cirrhotics. Patients and Methods: In a cross-sectional<br />

study, all NASH related cirrhotics presenting to Hepatology<br />

Department of ILBS, New Delhi from Jan 2014-May<br />

2015 were included. Other etiologies, like alcoholic, viral,<br />

autoimmune, cholestatitc liver diseases and inherited metabolic<br />

disorders were excluded. FH, demographic characteristics,<br />

medical history data, anthropometrical measurements and laboratory<br />

data were recorded. FH for T2DM was defined as at<br />

least one first degree relative having T2DM. Results: Out of<br />

1,033 NASH related cirrhotics (71.8% males) with mean age<br />

54.7 (±11.1) years, 549 (53.1%) had FH for T2DM (


1282A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

MTs had HCC (7.0% Vs 3.1%, p=0.004) at presentation.<br />

Conclusions: A significant number (52.7%) of NASH related<br />

cirrhotics had FH for T2DM. Importantly, FH of T2DM constitutes<br />

a high risk group for early and severe cirrhosis. Community<br />

based research to further substantiate this hypothesis and<br />

health education strategies to disseminate relevant awareness,<br />

is a prime requisite.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ajeet S. Bhadoria, Chandan K.<br />

Kedarisetty, Ankit Bhardwaj, Guresh Kumar, Tanmay S. Vyas, Jaya Benjamin,<br />

Varsha Shasthry, Manoj Kumar, Shiv K. Sarin<br />

2204<br />

Non-alcoholic Fatty Liver Disease (NAFLD) Is Associated<br />

with Impairment of Health Related Quality of Life<br />

(HRQoL): Data from US Population<br />

Munkhzul Otgonsuren 2 , Rebecca Cable 2 , Sean C. Felix 2 , Patrick<br />

W. Austin 2 , Yun Fang 2 , Zobair M. Younossi 2,1 ; 1 Center For Liver<br />

Disease, Department of Medicine, Inova Fairfax Hospital, Falls<br />

Church, VA; 2 Betty and Guy Beatty Center for Integrated Research,<br />

Inova Health System, Falls Church, VA<br />

Background and Aim: NAFLD is an important cause of chronic<br />

liver disease (CLD). Our aim was to assess the impact of NAFLD<br />

on patients’ HRQoL in the US adult population. Methods:<br />

National Health and Nutrition Examination Survey (NHANES<br />

2001-2011) was used to identify adults with NAFLD [Fatty Liver<br />

Index (FLI)>60 in the absence of other causes of liver disease,<br />

chronic diseases and excessive alcohol use]. Subjects without<br />

any of these conditions were considered as healthy controls.<br />

Patients with HCV RNA(+) were considered as HCV-controls.<br />

All patients completed HRQoL-4 questionnaire: (1) Component<br />

1 (C1) rate health status as poor, fair, good, very good, or<br />

excellent; (2) C2: thinking about your physical health, which<br />

includes physical illness and injury, for how many days during<br />

the past 30 days was your physical health not good; (3) C3:<br />

thinking about your mental health, which includes stress,<br />

depression, and problems with emotions, for how many days<br />

during the past 30 days was your mental health not good; and<br />

(4) C4: about how many days did poor physical/mental health<br />

keep him/her from doing your usual activities, such as selfcare,<br />

work, or recreation. Linear regression models adjusted<br />

for age, gender, race, and BMI was used to assess associations.<br />

Results: 9,661 NHANES participants (3,333 NAFLD,<br />

346 HCV+ and 5,982 controls). The proportion of subjects<br />

rating their health as “fair” or “poor” was higher in NAFLD<br />

(20%) than healthy controls (10%). However, 30% of HCV<br />

reported poor to fair health (all p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1283A<br />

= 4.80 (2.88-9.69)] are independent predictors of mortality<br />

(all p


1284A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

ciated with HCC recurrence in ALD-HCC. Conclusion: Although<br />

NAFLD-HCC and ALD-HCC differ in several characteristics, the<br />

clinical course of these two diseases is similar. Both have a<br />

high recurrence rate as with other types of HCC. Expression of<br />

the tumor markers DCP and AFP differed between NAFLD-HCC<br />

and ALD-HCC patients and might be useful for predicting HCC<br />

recurrence.<br />

Disclosures:<br />

The following authors have nothing to disclose: Tomomi Kogiso, Etsuko<br />

Hashimoto, Hideyuki Hyogo, Tomoaki Nakajima, Masafumi Ono, Takumi Kawaguchi,<br />

Koichi Honda, Yuichiro Eguchi, Yuichi Nozaki, Miwa Kawanaka, Kento<br />

Imajo, Yoshio Sumida, Yoshihiro Kamada, Hideki Fujii, Yasuaki Suzuki, Katsutoshi<br />

Tokushige<br />

2208<br />

Diacylglycerol acyltransferase 1 inhibition, as a metabolic<br />

regulator: clinical benefits of pradigastat in<br />

patients with non-alcoholic fatty liver disease<br />

Kenneth Cusi 1 , Arun J. Sanyal 2 , Samir Patel 3 , Melanie Wright 4 ,<br />

Changlu Liu 3 , Deborah L. Keefe 3 ; 1 Division of Endocrinology,<br />

Diabetes and Metabolism, University of Florida, Gainesville, FL;<br />

2 Division of Gastroenterology, Hepatology and Nutrition, Virginia<br />

Commonwealth University, Richmond, VA; 3 Novartis Pharmaceuticals<br />

Corporation, East Hanover, NJ; 4 Novartis Pharma AG, Basel,<br />

Switzerland<br />

Purpose: Non-alcoholic fatty liver disease (NAFLD) affects up<br />

to 80% of obese and diabetic patients. Pradigastat, a potent<br />

selective diacylglycerol acyltransferase 1 (DGAT1) inhibitor,<br />

reduces post-prandial triglycerides, glucose and insulin<br />

in overweight or obese patients. This study assessed the efficacy<br />

and safety of pradigastat on multiple metabolic parameters<br />

in NAFLD patients. Methods: This was a multicenter,<br />

randomized, double-blind, placebo-controlled 24-week study<br />

(NCT01811472). NAFLD patients with liver fat ≥10% were<br />

randomized (2:1:2) to placebo, pradigastat 5/10 mg or<br />

10/20 mg and up-titrated from pradigastat 5 to 10 mg (or<br />

10 to 20 mg) after 2 weeks. Due to the pilot nature of the<br />

study, the significance level was set at the 1-sided 5% (2-sided<br />

10%) level, and not fully powered for many of the metabolic<br />

parameters listed in Table 1. Results: A total of 52 patients<br />

were randomized, of whom 45 (87%) completed the study.<br />

Mean baseline BMI and waist circumference were 33.4 kg/<br />

m 2 and 110.0 cm, respectively, with 80% of patients being<br />

obese. There was a modest decrease in body weight and waist<br />

circumference with pradigastat compared to placebo, with the<br />

differences reaching statistical significance at Weeks 4 to 8<br />

(5/10 mg) and up to Week 20 (10/20 mg) for weight, and<br />

at week 12 for waist circumference (10/20 mg) (Table 1). A<br />

modest decrease was also observed in abdominal subcutaneous<br />

and visceral fat, as well as in some metabolic parameters<br />

(Table 1). Pradigastat was overall well tolerated except for GI<br />

symptoms. Diarrhea affected 35.0%, 54.5%, and 85.7% of<br />

patients on placebo, pradigastat 5/10 and 10/20 mg, respectively.<br />

Conclusion: Pradigastat treatment for up to 24 weeks<br />

led to a modest reduction in body weight (1.3-1.4 kg). Both<br />

pradigastat groups were associated with numerically greater<br />

and sustained decreases in different markers of body fat compared<br />

to placebo. The significance of these effects needs to be<br />

confirmed in larger <strong>studies</strong>.<br />

Table 1: Adjusted mean change from baseline of metabolic<br />

parameter<br />

*significant at 2-sided 10% level vs placebo<br />

Disclosures:<br />

Kenneth Cusi - Consulting: Merck, Janssen, Pfizer, Pfizer; Grant/Research Support:<br />

Janssen, Novartis<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Samir Patel - Employment: Novartis Pharmaceutical<br />

Melanie Wright - Employment: Novartis Pharma AG<br />

Changlu Liu - Employment: Novartis Pharmaceuticals<br />

Deborah L. Keefe - Employment: Novartis Pharmaceuticals Corp<br />

2209<br />

Subjects with non alcoholic fatty liver disease (NAFLD)<br />

display increased visceral and pancreatic fat associated<br />

with worse metabolic profile<br />

Chiara Saponaro 1 , Melania Gaggini 1 , Chiara Rosso 2 , Emma<br />

Buzzigoli 1 , Fabrizia Carli 1 , Demetrio Ciociaro 1 , Lavinia Mezzabotta<br />

2 , Francesca Saba 2 , Andrea Marengo 2 , Maria Lorena<br />

Abate 2 , Riccardo Faletti 2 , Maurizio Cassader 2 , Roberto Gambino<br />

2 , Antonina Smedile 2 , Mario Rizzetto 2 , Elisabetta Bugianesi 2 ,<br />

Amalia Gastaldelli 1 ; 1 CNR, Pisa, Institute of Clinical Physiology,<br />

Pisa, Italy; 2 Dept. of Medical Sciences, University of Turin, Division<br />

of Gastroenterology and Hepatology and Lab. of Diabetology,<br />

Turin, Italy<br />

Together with insulin resistance (IR), lipotoxicity and inflammation,<br />

patients with non-alcoholic fatty liver diseases (NAFLD)<br />

often show increased visceral fat (VF) that correlates with<br />

increased insulin and hepatic fat (HF). If VF is playing a key<br />

role in the progression of liver disease is not clear. Beyond<br />

VF and HF, fat can also accumulate in the pancreas (PF) leading<br />

to impaired insulin secretion and increasing the risk of<br />

diabetes (T2DM). The aim of this work was to measure by<br />

magnetic resonance (MR) the separate impact of VF, HF and<br />

PF on histological liver damage, lipid profile and metabolic<br />

alterations in non diabetic patients with biopsy proven NAFLD.<br />

Methods: We studied 34 non diabetic subjects with NAFLD<br />

(biopsy scored according to Kleiner) and 8 healthy controls<br />

(CT). We measured plasma concentrations of triglycerides<br />

(TG), free fatty acids (FFA), insulin, C-peptide, monocyte chemoattractant<br />

protein-1(MCP-1) and adiponectin, FFAs composition<br />

by chromatography mass spectrometry (GCMS), visceral,<br />

hepatic and pancreatic fat by MR. In addition, using tracers,<br />

we evaluated lipolysis and endogenous glucose production<br />

(EGP). We calculated indexes of IR (i.e., HOMA, Adipo-IR<br />

=Lipolysis x Insulin and Hep-IR= EGP x Insulin); the ratios palmitic/linoleic<br />

acid (16:0/18:2, an indirect index of de novo<br />

lipogenesis, DNL) and the saturated to unsaturated fat ratio<br />

(SFA/PUFA, associated with impaired lipid metabolism and<br />

oxidative stress). Results: NAFLD patients with fibrosis F1-4<br />

(n=24) had a worse metabolic profile compared to NAFLD<br />

with F0 (n=10) and CT, showing increased parameters of


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1285A<br />

insulin resistance, inflammation (MCP-1) and lipotoxicity (i.e.<br />

increased DNL index and SFA/PUFA). Visceral and hepatic<br />

fat were both significantly increased in NAFLD, especially<br />

in F1-4 vs F0 vs CT (VF=2.9±1.1 vs 2.1±0.6 kg vs 0.7±0.4<br />

kg p


1286A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2212<br />

A Meta-analytic Assessment of Worldwide Prevalence<br />

of Non Alcoholic Fatty Liver Disease (NAFLD) and Associated<br />

Co-morbidities<br />

Aaron B. Koenig 1 , Dinan Abdelatif 1 , Yousef Fazel 1 , Mark Wymer 1 ,<br />

Linda Henry 1 , Zobair M. Younossi 2 ; 1 Betty and Guy Beatty Center<br />

for Integrated Research, Inova Health System, Falls Church, VA;<br />

2 Center For Liver Disease, Department of Medicine, Inova Fairfax<br />

Hospital, Falls Church, VA<br />

Background: NAFLD has become a major public health problem<br />

worldwide. Our aim was to determine the prevalence of<br />

NAFLD worldwide, its associated comorbidities, and prevalence<br />

of non-alcoholic steatohepatitis (NASH). Methods: A systematic<br />

literature review using PubMed, Ovid Medline and<br />

Web of Science was done with appropriate search terms.<br />

Exclusion criteria: any study for special population (morbid<br />

obesity) and alcohol consumption>10 g/day. Studies were<br />

reviewed by 3 investigators. Reporting followed the Preferred<br />

Reporting Items for Systematic reviews and Meta-Analyses<br />

(PRISMA) statement. Random-effects model univariate meta-regression<br />

explored high heterogeneity sources (i.e. age, gender,<br />

country, year of study, and study design). Results: 700+ articles<br />

were reviewed. After exclusion criteria, 95 <strong>studies</strong> examined,<br />

53 analyzed [Asia n=13, Europe n=13, Middle East<br />

(ME) n=3, North America (NA) n=23 and South American (SA)<br />

n=1]. Number of <strong>studies</strong> for prevalence of comorbidities varied<br />

[obesity n=19; diabetes (DM) n=35, hyperlipidemia (HL)<br />

n=14, hypertriglyceridemia (HTG) n=8, metabolic syndrome<br />

(MS) n=18, hypertension (HTN) n=26, NASH n=4]. Worldwide<br />

NAFLD prevalence was 21.3% (17.9%-30.5%); By continent,<br />

Asia- 24%(19.6%-30.5%); Europe-21%(12.7%-31.7%);<br />

Middle East- 31.8% (13.5%-58.2%); North America- 18.5%<br />

(14.3%-23.6%); SA-35.3% (27.8%-43.5%)(Figure). Rate of<br />

obesity was 47.4% (34.2%-60.9%); DM: 18% (13.6%-23.5%);<br />

HL-62.2% (45.3%-76.5%); HTG- 37.3% (25.4%-51.0%); MS-<br />

43.3% (29.4%-58.3%); HTN- 38.6% (32.1%-45.7%) and<br />

NASH 26.2% (17.3%-37.7%). Conclusion: The prevalence of<br />

NAFLD and NASH are high with some variation throughout<br />

the world. HL and MS appeared to be the most prevalent risk<br />

factors associated with NAFLD.<br />

Disclosures:<br />

Zobair M. Younossi - Advisory Committees or Review Panels: Salix, Janssen,<br />

Vertex; Consulting: Gilead, Enterome, Coneatus<br />

The following authors have nothing to disclose: Aaron B. Koenig, Dinan Abdelatif,<br />

Yousef Fazel, Mark Wymer, Linda Henry<br />

2213<br />

Risk factors for progression of liver disease status in<br />

Japanese patients with non-alcoholic steatohepatitis<br />

Ayako Urabe 1 , Naoki Hiramatsu 1 , Tsugiko Oze 1 , Takayuki<br />

Yakushijin 1 , Ryoko Yamada 1 , Naoki Morishita 1 , Yuki Tahata 1 ,<br />

Yuichi Yoshida 1 , Tomohide Tatsumi 1 , Masahide Oshita 2 , Akira<br />

Kaneko 3 , Eiji Mita 4 , Masahiko Tsujii 5 , Toshihiko Nagase 6 ,<br />

Hiroyuki Fukui 8 , Kunio Suzuki 7 , Atsuo Inoue 9 , Yasuharu Imai 10 ,<br />

Tetsuo Takehara 1 ; 1 Gastroenterology and Hepatology, Osaka University<br />

graduate School of Medicine, Suita, Japan; 2 Osaka Police<br />

Hospital, Osaka, Japan; 3 NTT West Osaka Hospital, Osaka,<br />

Japan; 4 National Hospital Organization Osaka National Hospital,<br />

Osaka, Japan; 5 Osaka Rosai Hospital, Sakai, Japan; 6 Suita<br />

Municipal Hospital, Suita, Japan; 7 Saiseikai Senri Hospital, Suita,<br />

Japan; 8 Yao Municipal Hospital, Yao, Japan; 9 Osaka General<br />

Medical Center, Osaka, Japan; 10 Ikeda Municipal Hospital, Ikeda,<br />

Japan<br />

Background & Aim: The incidence of hepatocellular carcinoma<br />

(HCC) of chronic liver disease with hepatitis C virus infection<br />

is decreasing by the advance of antivirus therapy, while the<br />

increase of HCC from non-alcoholic steatohepatitis (NASH)<br />

is becoming a problem, recently. However, the risk factors<br />

for progression of NASH are still unknown. We performed<br />

a longitudinal study to investigate factors associated with the<br />

progression of histologic fibrosis and long-term outcomes of<br />

NASH. Patients & Methods: We did a retrospective analysis<br />

of 150 Japanese patients histologically diagnosed with NASH<br />

from March 1998 to March 2009 at the Osaka University<br />

Hospital and other institutions participating in the Osaka Liver<br />

Forum. All patients were Japanese. The patient characteristics<br />

at baseline were: mean age, 55.7±15.5 y.o.; male/female,<br />

86/64; BMI>25kg/m 2 , 104 (69%); ALT level, 113.6±82.8<br />

U/L; platelet count, 20.5±7.8 /μL. Concurrent diseases were:<br />

dyslipidemia, 77 (51%); hypertension, 56 (37%); diabetes, 49<br />

(33%). Histological appearance according to Brunt classification<br />

was: grade 1/2/3, 57(38%)/74(49%)/16(11%); stage<br />

1/2/3/4, 49(33%)/42(28%)/39(26%)/17(11%). Risk factors<br />

associated with the incidence of decompensated cirrhosis<br />

(varix, ascites, hepatic encephalopathy), HCC and liver-related<br />

death were examined for 134 patients who were followed up.<br />

The mean observation period was 72.9 months±41.2 months.<br />

The Kaplan-Meiyer method and Cox proportional-hazards<br />

model were used. Results: Significant factors associated with<br />

advanced fibrosis at baseline were diabetes (p=0.016), Brunt<br />

grade (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1287A<br />

Disclosures:<br />

Tetsuo Takehara - Grant/Research Support: Chugai Pharmaceutical Co., MSD<br />

K.K., Bristol-Meyer Squibb, Mitsubishi Tanabe Pharma Corparation, Toray Industories<br />

Inc. ; Speaking and Teaching: MSD K.K., Bristol-Meyer Squibb, Janssen<br />

Pharmaceutical Companies<br />

The following authors have nothing to disclose: Ayako Urabe, Naoki Hiramatsu,<br />

Tsugiko Oze, Takayuki Yakushijin, Ryoko Yamada, Naoki Morishita, Yuki<br />

Tahata, Yuichi Yoshida, Tomohide Tatsumi, Masahide Oshita, Akira Kaneko, Eiji<br />

Mita, Masahiko Tsujii, Toshihiko Nagase, Hiroyuki Fukui, Kunio Suzuki, Atsuo<br />

Inoue, Yasuharu Imai<br />

2214<br />

Concomitant Non-Alcoholic Fatty Liver Disease Drives<br />

Progression of Overall Liver Disease More Than Chronic<br />

Hepatitis B Alone in Chinese-Americans<br />

Rouchelle D. dela Cruz 1 , Lan Sun Wang 1 , Anthony Ng 3 , Rene S.<br />

Eng 3 , Lauren M. Eng 3 , Neil D. Theise 1,2 ; 1 Division of Digestive<br />

Diseases, Mount Sinai Beth Israel Medical Center, New York, NY;<br />

2 Department of Pathology, Mount Sinai Beth Israel Medical Center,<br />

New York, NY; 3 Ng Medical Group, New York, NY<br />

Introduction: Despite the proven efficacy of current antiviral<br />

therapy for chronic hepatitis B (CHB), up to 20% of patients on<br />

long term oral antiviral treatment who have complete HBV suppression<br />

have persistently elevated serum alanine transferase<br />

(ALT) levels. It has been previously shown that non-alcoholic<br />

fatty liver disease (NAFLD) is a likely culprit. This phenomenon<br />

will contribute to an increasingly large cohort of patients who<br />

will develop NAFLD-related cirrhosis, decompensated liver disease,<br />

and hepatocellular carcinoma. Asians may be especially<br />

affected given endemic CHB and significant prevalence of metabolic<br />

syndrome and NAFLD despite lower body mass index<br />

and rate of obesity compared to other ethnicities. Methods:<br />

Consecutive Chinese American patients from a single non-transplant<br />

urban referral center (Mount Sinai Beth Israel Medical<br />

Center, New York City) who underwent biopsies for CHB<br />

between January 1, 2008 and Dec 31, 2012 were included in<br />

this retrospective study. The biopsy specimens were reviewed,<br />

regraded and restaged for NAFLD according to the scoring<br />

system designed and validated by Kleiner, Brunt and the<br />

Pathology Committee of the NASH Clinical Research Network.<br />

Concomitant HBV grading and staging were also evaluated<br />

using Scheuer’s classification for chronic hepatitis. Corresponding<br />

clinical data was gathered. The patients were divided into<br />

those with NAFLD (+) and without NAFLD (-) on pathology.<br />

Pathology and clinical features were compared between the<br />

two groups using the Wilcoxon rank-sum test. Results: Of the<br />

148 patients, 41 were NAFLD+: 38 with steatosis, 18 with<br />

steatohepatitis, 13 with steatofibrosis. Two of the patients with<br />

steatofibrosis had no NAFLD activity but had significant fibrosis<br />

attributable to NASH. There is a significantly higher combined<br />

necroinflammatory activity (P=.00001) and overall combined<br />

staging (P =.0015) in patients in the NAFLD+ group as compared<br />

to the NAFLD- group. Clinically, alanine transaminase<br />

(ALT, P=0.0084), body mass index (BMI, P


1288A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2216<br />

The dual and opposite role of the TM6SF2-rs58542926<br />

Variant in Protecting against Cardiovascular Disease<br />

and Conferring Risk for Non-alcoholic fatty liver: A<br />

meta-analysis<br />

Carlos J. Pirola 2 , Silvia Sookoian 1 ; 1 Clinical and Molecular Hepatology,<br />

IDIM-CONICET, Ciudad Autonoma de Buenos Aires,<br />

Argentina; 2 Molecular Genetics and Biology of Complex Diseases,<br />

IDIM CONICET, Buenos Aires, Argentina<br />

Background: The nonsynonymous p.Glu167Lys (rs58542926<br />

C/T, E167K) variant, located in TM6SF2 (transmembrane 6<br />

superfamily member 2) gene, was associated not only with<br />

blood lipid levels, including serum total cholesterol (TC),<br />

low-density lipoprotein cholesterol (LDL-C) and triglycerides<br />

(TG), but also myocardial infarction risk and NAFLD susceptibility.<br />

This variant, which has a relatively low frequency not<br />

only in Caucasian population (0.09) but globally (0.07), presents<br />

a clinical paradox, as the C (Glu167) allele was shown<br />

to be associated with increased CVD risk, while the T allele<br />

(Lys167) was associated with NAFLD. Objective: to examine<br />

the evidence provided in the available literature to estimate the<br />

strength of the effect of rs58542926 on both circulating lipid<br />

traits and NAFLD across different populations. In addition, we<br />

assessed whether associations are consistent across <strong>studies</strong><br />

in magnitude and direction for all explored traits. Methods/<br />

design: We performed a systematic review by a meta-analysis;<br />

literature searches identified eleven <strong>studies</strong>. Results: The<br />

rs58542926 exerts a significant role in modulating lipid traits,<br />

including TC, LDL-C, TG, and NAFLD. However, this influence<br />

on lipids and NAFLD is opposite between genotypes in the<br />

dominant model of inheritance. Pooled estimates of random<br />

effects in 193,931 individuals showed that, compared with<br />

carriers of the minor K allele (EK+KK individuals), subjects<br />

homozygous for the ancestral C allele (EE genotype) are at risk<br />

of having higher levels of TC, LDL-C and TG; the differences<br />

in mean±SE (mg/dL) are 7.4±2.3, 3.7±0.9 and 9.4±2.1,<br />

respectively. The rs58542926 was not associated with HDL-C<br />

in a large sample (n =91,937). Of note, the EE genotype is<br />

associated with a relatively large effect on blood levels of TC<br />

(about 2-3% increase on an upper limit of 200 mg/dL), LDL-C<br />

(about 1-2% increase on an upper limit of 150 mg/dL), and<br />

TG (about 6.4% increase on an upper limit of 150 mg/dL) for<br />

a single variant influencing a polygenic trait. In contrast, homozygous<br />

EE subjects appear to be protected against NAFLD<br />

(OR 0.469, 95% CI 0.300-0.734, p = 0.0009, n = 3273),<br />

while carriers of the K allele show about~ 2.2% higher lipid<br />

fat content when compared with homozygous EE (n = 3,413),<br />

and have 2.13-fold higher risk of developing NAFLD. Conclusion:<br />

The rs58542926 appears to be an important modifier of<br />

blood lipid traits in different populations. As a challenge for<br />

the personalized medicine, the “major” C-allele, which has a<br />

frequency as high as 93%, is associated with CVD risk, while<br />

the “minor”-low frequency T allele confers risk for NAFLD; in<br />

turn, CVD and NAFLD are strongly related outcomes.<br />

Disclosures:<br />

The following authors have nothing to disclose: Carlos J. Pirola, Silvia Sookoian<br />

2217<br />

Mitochondrial Mutator-Phenotype May Be Related<br />

to Pathogenesis of Nonalcoholic Fatty Liver Disease:<br />

Insights from Deep Sequencing of Liver Mitochondrial<br />

Genomes<br />

Carlos J. Pirola 1 , Romina Scian 1,2 , Cristian O. Rohr 3 , Hernán<br />

Dopazo 3 , Gustavo O. Castaño 4 , Silvia Sookoian 2 ; 1 Molecular<br />

Genetics and Biology of Complex Diseases, IDIM-CONICET,<br />

Buenos Aires, Argentina; 2 Clinical and Molecular Hepatology,<br />

IDIM-CONICET, Buenos Aires, Argentina; 3 Biomedical Genomics<br />

and Evolution Laboratory, CONICET, Buenos Aires, Argentina;<br />

4 Medicine and Surgery Department, Liver Unit, Hospital Abel Zubizarreta,<br />

Buenos Aires, Argentina<br />

Background: Mitochondrial (mt) dysfunction is involved in the<br />

development of NAFLD; normal activity of mitochondria critically<br />

determines fatty acid beta-oxidation, OXPHOS and insulin<br />

signaling. In addition, liver mitochondrial biogenesis is reduced<br />

in NAFLD. The mt-genome is highly polymorphic and variants<br />

in mtDNA affect mt function. A small proportion of mtDNA<br />

belongs to the control region involved in mtDNA duplication<br />

(D-loop). To understand the clinical implications of mtDNA-variation<br />

in the pathogenesis of NAFLD, we sequenced whole liver<br />

mtDNA-genomes of 28 individuals, including patients with<br />

NAFLD (n = 20) and age and sex-matched controls (n = 8). In<br />

addition, we sequenced the entire nuclear POLG and POLG2<br />

genes, which are involved in mtDNA-replication. Methods:<br />

Liver mtDNA was first amplified by long-range PCR; deep next<br />

generation sequencing was further performed. Results: We<br />

achieved an average read depth >800 per individual; mtDNA<br />

sequencing revealed 689 variants, 525 (76%) of them were<br />

observed in NAFLD showing an enrichment of 1.28-fold mutation<br />

fraction compared to controls (p = 0.0056). Ten of 16<br />

base positions containing heteroplasmic variants were highly<br />

polymorphic (>0.1) and were observed in NAFLD. The mutation<br />

fraction of liver mtDNA-genomes in controls compared with<br />

that in patients with simple steatosis (NAFL) shows no significant<br />

differences. Remarkably, patients with NASH compared<br />

with controls harbored a significantly higher number of mutations<br />

in mitochondrially encoded ATP synthase 6 (p = 0.033),<br />

cytochrome b (MT-CYB) (p = 0.011), and members of the<br />

NADH-dehydrogenase complex (p = 4.5 E -5 ). The comparison<br />

of liver mtDNA-diversity among patients with NAFL and NASH<br />

showed that the disease severity was associated with increased<br />

number of variants in the NADH-complex (p = 7.4 E -3 ). For variants<br />

predicted to be deleterious, an allelic association analysis<br />

was further conducted; we found a nonsynonymous variant in<br />

MT-CYB (m.15326) and two mutations in D-loop (m.146 and<br />

m.16298) that were significantly associated with the disease<br />

severity. Overall, mutations located in the NADH-complex were<br />

significantly associated with liver-related phenotypes and arterial<br />

hypertension. The missense p.Gln1236His variant in POLG<br />

was associated with liver mtDNA-copy number (p = 0.01) and<br />

the POLG2- rs7223078 was associated with the degree of histological<br />

steatosis (p = 0.036). Conclusion: The burden of mutations<br />

in liver mt-genomes may contribute to the pathogenesis<br />

of NAFLD and metabolic syndrome-associated comorbidities<br />

explaining part of the “missing heritability”. NASH development<br />

may be associated with an OXPHOS-negative phenotype.<br />

Disclosures:<br />

The following authors have nothing to disclose: Carlos J. Pirola, Romina Scian,<br />

Cristian O. Rohr, Hernán Dopazo, Gustavo O. Castaño, Silvia Sookoian


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1289A<br />

2218<br />

Reverse correlation with FXR expression of peripheral<br />

mononuclear cell and NPC1L1 expression of terminal<br />

ileum in non-alcoholic fatty liver disease.<br />

Sang Bong Ahn 1 , Dae Won Jun 2 , Yong Kyun Cho 3 , Eun Chul Jang 4 ,<br />

Seung Min Lee 5 , Kiseok Jang 6 ; 1 Department of Internal Medicine,<br />

Eulji University School of Medicine, Seoul, Korea (the Republic of);<br />

2 Department of Internal Medicine, Hanyang University College of<br />

Medicine, Seoul, Korea (the Republic of); 3 Department of Internal<br />

Medicine, Kangbuk Samsung Hospital, Sungkyunkwan University,<br />

School of Medicine, Seoul, Korea (the Republic of); 4 Department<br />

of Occupational and Environmental Medicine, Soonchunhyang<br />

University Cheonan Hospital, Cheonan, Korea (the Republic of);<br />

5 Department of Food and Nutrition, Sungshin Women’s University,<br />

Seoul, Korea (the Republic of); 6 Department of Internal Medicine,<br />

and Pathology, Hanyang University School of Medicine, Seoul,<br />

Korea (the Republic of)<br />

Background: Niemann-Pick C1-Like 1 (NPC1L1) functions as<br />

sterol transporter to mediate intestinal cholesterol absorption.<br />

Although the role of Liver X receptor (LXR) and Farnesoid X<br />

receptor (FXR) in hepatic steatosis is well known, its correlation<br />

with LXR/FXR and intestinal cholesterol circulation has not been<br />

thoroughly studied. We investigated the associations among<br />

LXR, FXR and NPC1L1 as well as its correlation with intestinal<br />

cholesterol circulation in patients with non-alcoholic fatty liver<br />

disease (NAFLD). Methods: We evaluated clinical characteristics<br />

from 25 NAFLD and 28 control patients. We conducted<br />

upper endoscopy, colonoscopy and non-enhanced CT. Fatty<br />

liver diagnosed with decreased liver/spleen HU. We calculated<br />

the expression level of LXR and FXR in peripheral mononuclear<br />

cell by qPCR. Immunohistochemical analysis (ABCA1,<br />

ABCG5/8, NPC1L1, SREBP, LXR, FXR, CD36) was carried<br />

out on tissue samples of duodenum and ileum from NAFLD<br />

patients. Results: The expression of LXR (p=0.01) and FXR<br />

(p=0.03) in mononuclear cells increased in NAFLD group compared<br />

to the control group. Expression of LXR, FXR, NPC1L1,<br />

ABCA1, ABCG5/8, and SREBP in ileum was decreased in<br />

fatty liver disease. The expression of LXR in mononuclear cell<br />

showed a negative correlation with the expression of LXR,<br />

FXR, ABCA1, ABCG8, SREBP, NPC1L1 in ileum. Expression<br />

of NPC1L1 (p=0.15), FXR (p=0.30) in ileum was decreased<br />

in patients with hypercholesterolemia. Expression of NPC1L1<br />

(p=0.03), LXR (duodenum) and LXR (ileum) was decreased as<br />

triglyceride increased. Conclusion: We can predict the expression<br />

of NPC1L1 in terminal ileum by measuring the expression<br />

of LXR and FXR in peripheral mononuclear cell. Keywords: Liver<br />

X receptor, Farnesoid X receptor, NPC1L1, non-alcoholic fatty<br />

liver disease<br />

Disclosures:<br />

The following authors have nothing to disclose: Sang Bong Ahn, Dae Won Jun,<br />

Yong Kyun Cho, Eun Chul Jang, Seung Min Lee, Kiseok Jang<br />

2219<br />

Association of Nonalcoholic Fatty Liver Damage and<br />

the Degree of Nocturnal Hypoxia in Obstructive Sleep<br />

Apnea<br />

Erol Cakmak 1 , Faysal Duksal 2 , Engin Altinkaya 3 , Fettah Acibucu 5 ,<br />

Omer T. Dogan 4 , Ozlem Yonem 1 , Abdulkerim Yilmaz 1 ; 1 Department<br />

of Gastroenterology, Cumhuriyet University School of Medicine,<br />

Sivas, Turkey; 2 Department of Chest Diseases, Numune<br />

Hospital, Sivas, Turkey; 3 Department of Gastroenterology, Kayseri<br />

Training and Research Hospital, Kayseri, Turkey; 4 Department of<br />

Chest Diseases, Cumhuriyet University School of Medicine, Sivas,<br />

Turkey; 5 Department of Endocrinology, Numune Hospital, Sivas,<br />

Turkey<br />

Background: Nonalcoholic fatty liver disease (NAFLD) has a<br />

large spectrum of findings ranging from a simple fatty liver to<br />

advanced fibrosis, cirrhosis and hepatocellular cancer. Obstructive<br />

sleep apnea (OSA) can lead to the development and progression<br />

of NAFLD because of repeated nocturnal hypoxia.<br />

The aim of this study was to evaluate the effect of the nocturnal<br />

hypoxia severity on hepatic steatosis. Methods: The parameters<br />

evaluated in this retrospective study in 137 subjects were body<br />

mass index, biochemical tests, degree of steatosis according<br />

to the findings of the liver ultrasound, and polysomnographic<br />

parameters demonstrating the severity of OSA (apnea-hypopnea<br />

index [AHI], oxygen desaturation index [ODI], mean<br />

nocturnal SpO2, lowest desaturation value [LaSO2]). According<br />

to the AHI score, patients with sleep apnea with a score of<br />

30 were categorized as control, mild,<br />

intermediate, and severe, respectively. Results: The total number<br />

of subjects included in this study was 137 (76 females and<br />

61 males) with a mean age of 56±10 years. Among the 118<br />

patients who were diagnosed with OSA, 19 had mild OSA,<br />

39 had intermediate OSA, 60 had severe OSA, and there<br />

were 19 cases in the control group. AHI and ODI values were<br />

significantly higher in the intermediate NAFLD (p


1290A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2220<br />

Transient Elastography is Feasible with High Success<br />

Rate for Evaluation of Non-Alcoholic Fatty Liver Disease<br />

(NAFLD) in a Multicenter Setting<br />

Raj Vuppalanchi 1 , Mohammad S. Siddiqui 2 , Erin K. Hallinan<br />

3 , Manal F. Abdelmalek 4 , Brent A. Neuschwander-Tetri 5 ,<br />

Rohit Loomba 6 , Kris V. Kowdley 7 , Norah Terrault 8 , Arthur J.<br />

McCullough 9 , James Tonascia 3 , Elizabeth M. Brunt 10 , David E.<br />

Kleiner 11 , Edward Doo 12 , Naga P. Chalasani 1 , Arun J. Sanyal 2 ;<br />

1 Indiana University, Indianapolis, IN; 2 Virginia Commonwealth,<br />

Richmond, VA; 3 Johns Hopkins Bloomberg School of Public Health,<br />

Baltimore, MD; 4 Duke, Durham, NC; 5 St.Louis University, St.Louis,<br />

MO; 6 University of California at San Diego, San Diego, CA;<br />

7 Swedish Medical Center, Seattle, WA; 8 University of California at<br />

San Francisco, San Francisco, CA; 9 Cleveland Clinic, Cleveland,<br />

OH; 10 Washington University, St.Louis, MO; 11 NCI, Bethesda,<br />

MD; 12 NIH, Bethesda, MD<br />

Background: Vibration-controlled transient elastography using<br />

FibroScan to obtain the liver stiffness measurement (LSM) and<br />

controlled attenuation parameter (CAP) is a promising non-invasive<br />

tool for assessment of hepatic fibrosis and steatosis<br />

respectively. However, prior <strong>studies</strong> using older versions of<br />

FibroScan reported high failure rates in nonalcoholic fatty liver<br />

disease (NAFLD). Aim: The aim of the current study was to<br />

examine the feasibility and success of the FibroScan 502 Touch<br />

in patients with NAFLD who were participating the NAFLD<br />

Database Study conducted by the NASH Clinical Research<br />

Network (NASH CRN). Methods: A total of 511 NAFLD<br />

patients across eight clinical centers were eligible and provided<br />

informed consent. The FibroScan 502 Touch was used in<br />

all patients with either the medium (M+) or large (XL+) probes.<br />

Both LSM and CAP were measured after an overnight fast<br />

using a standardized protocol. The scanning was performed<br />

twice (FS1 and FS2) during the same study visit. “Failure” was<br />

defined as the inability to obtain a valid exam whereas “unreliable”<br />

was defined as those exams with 30%. Results: The cohort<br />

included 65% women with a mean (± SD) age of 56 (± 12)<br />

years, BMI 33.6 (± 6.8) kg/m 2 and waist circumference 107<br />

(± 14) cm. Based on the recommendations from the automatic<br />

probe selection tool, XL+ probe was required for the study in<br />

57% of patients. Five patients (1.0%) refused the procedure<br />

after providing informed consent and five patients (1.0%) had<br />

a skin-to-capsule distance greater than 3.5 cm and the exam<br />

could not be completed. Among remaining 501 patients, the<br />

results were unreliable in 8 (1.6%). Using logistic regression<br />

analysis, only female gender [OR: 0.29, 95%CI: 0.10, 0.83,<br />

p=0.02) and waist circumference [OR: 1.04 per cm, 95%CI:<br />

1.01, 1.08, p=0.004) were associated with a failed or unreliable<br />

exam. There was excellent agreement between FS1 and<br />

FS2 for LSM (r=0.96, 95% CI: 0.95, 0.97) and CAPs (r=0.81,<br />

95% CI: 0.77, 0.84). In a subset of patients who had a liver<br />

biopsy within 12 months of the FibroScan (n=74), log-linear<br />

regression analysis showed statistically significant relationships<br />

between the geometric mean (GM) LSM and advanced (stage<br />

3 or 4) vs. no or less advanced fibrosis [GM ratio: 2.04 (95%<br />

CI: 1.60, 2.60), p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1291A<br />

Disclosures:<br />

The following authors have nothing to disclose: Claudia P. Oliveira, José Tadeu<br />

Stefano, Roberto M. Ribeiro, Sebastião M. Duarte, Livia Rodrigues, Priscila B.<br />

Campos, Fernando G. Costa, Daniel F. Mazo, Flair J. Carrilho, Ester C. Sabino<br />

2222<br />

Outcome of bariatric surgery, in highly selected morbid<br />

obese with compensated cirrhosis<br />

Guillaume Lassailly, Robert Caiazzo, Charlotte Vanveuren,<br />

gnemmi vivianne, Alexandre Louvet, Emmanuelle Leteurtre, Florent<br />

Artru, Sebastien Dharancy, Valerie Canva, Francois Pattou,<br />

Philippe Mathurin; CHRU lille, Lille, France<br />

Introduction: Morbid obesity contributes to liver function impairment<br />

in cirrhotic patients and especially in NASH. Weight loss<br />

strategies could improve liver function, but bariatric surgery<br />

is not considered due to an increased risk of mortality after<br />

surgery. However its benefit and feasibility in highly selected<br />

patients is still unclear. The aim of this study was to describe<br />

the outcome of highly selected cirrhotic patients after bariatric<br />

surgery. Methods: Among the 2024 morbid obese patients of<br />

the Lille Bariatric prospective cohort (1994-2015), 28 compensated<br />

cirrhotic patients underwent bariatric surgery. When cirrhosis<br />

was identified before surgery, each case was discussed<br />

to validate the procedure. Only Child-Pugh A5 patients without<br />

portal hypertension (without varice and hepatic vein gradient<br />


1292A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

The following authors have nothing to disclose: Leonardo A. Martinez Rodriguez,<br />

Aldo Torre, David Kershenobich<br />

2224<br />

The association of ALT levels with myocardial perfusion<br />

imaging and cardiovascular morbidity<br />

David Yardeni 2 , Ohad Etzion 1,3 , Ronen Toledano 4 , Victor<br />

Novack 4 , Aryeh Shalev 5 , Arik Wolak 5 , Yaron Rotman 1 ; 1 Liver Disease<br />

Branch, NIH, Rockville, MD; 2 Internal Medicine Division,<br />

Soroka University Medical Center, Beer-Sheva, Israel; 3 Gastroenterology<br />

and Liver disease, Soroka University Medical Center,<br />

beer-Sheva, Israel; 4 Clinical Research Center, soroka University<br />

Medical Center, Beer-Sheva, Israel; 5 Department of Cardiology,<br />

Soroka University Medical Center, beer-Sheva, Israel<br />

Background and aims: There is a growing body of evidence<br />

suggesting that non-alcoholic fatty liver disease (NAFLD) is<br />

associated with an independent increased risk of cardiovascular<br />

disease (CVD) beyond that is conferred by the metabolic<br />

syndrome. We utilized a large cohort of patients undergoing<br />

myocardial perfusion imaging (MPI) with single photon emission<br />

computed tomography (SPECT) to determine the association<br />

between alanine aminotransferase (ALT) as a surrogate<br />

marker for NAFLD,and the presence of myocardial ischemia as<br />

well as the effects of ALT on mortality in patients with and without<br />

ischemia. Methods: We retrospectively assessed SPECT-MPI<br />

and laboratory test results, prescriptions and clinical diagnoses,<br />

extracted from the electronic medical records of all individuals<br />

who underwent SPECT MPI at Soroka University Medical<br />

Center between 1997 and 2008. We excluded patients with<br />

viral hepatitis, positive autoimmune markers, ALT values 340 u/L, and absent liver tests within 1 year prior or<br />

following SPECT MPI. Cases with elevated ALT were classified<br />

as presumed NAFLD. The primary outcome was abnormal myocardial<br />

perfusion, defined as >1% perfusion defect. Secondary<br />

outcomes were a composite cardiac outcome of cardiac death<br />

or acute myocardial infarction 3 and 5 years following SPECT-<br />

MPI, and all-cause mortality Results: Of 26,028 patients who<br />

underwent SPECT-MPI, 11,582 met inclusion criteria and were<br />

included in the final analysis. 1,704 (14.7%) patients had elevated<br />

ALT within a year of SPECT-MPI. Patients with abnormal<br />

ALT were younger and included significantly more males and<br />

smokers (p < 0.001 for all). Dyslipidemia, diabetes mellitus,<br />

obesity and congestive heart failure were also more common<br />

in the elevated ALT group. SPECT-MPI results did not differ<br />

between subjects with elevated ALT and controls. Elevated ALT<br />

was associated with increased risk for the composite outcome<br />

of cardiac death or acute myocardial infarction at 3-year follow-up,<br />

HR, 1.46 (95% CI 1.07–1.99) and in all cause mortality<br />

(HR. 1.69, CI (1.28–2.23) but only in patients with normal<br />

SPECT-MPI. At 5-years follow-up ALT levels were not associated<br />

with the composite outcome or all-cause mortality. Conclusions:.<br />

The long-term mortality of patients with abnormal SPECT-MPI is<br />

not modulated by ALT, likely reflecting an already high risk and<br />

established advanced atherosclerotic process. On the other<br />

hand, patients with normal SPECT-MPI are at increased risk for<br />

a future cardiac event if they have an elevated ALT level, suggesting<br />

an important role for NAFLD in earlier stages of CVD.<br />

Our results highlight the complex interaction between NAFLD<br />

and ischemic heart disease.<br />

Disclosures:<br />

The following authors have nothing to disclose: David Yardeni, Ohad Etzion,<br />

Ronen Toledano, Victor Novack, Aryeh Shalev, Arik Wolak, Yaron Rotman<br />

2225<br />

Non-Alcoholic Fatty Liver Disease (NALFD) and Hepatocellular<br />

Carcinoma (HCC), clinical characteristics and<br />

outcome: A single centre experience.<br />

Sum Team Lo, David W. Orr, Adam Barlett, Edward J. Gane; New<br />

Zealand Liver Transplant Unit, Auckland, New Zealand<br />

Aims: To compare patient survival between NAFLD-related<br />

HCC and chronic viral hepatitis-related HCC and to examine<br />

the risk factors and characteristics of NAFLD patients with<br />

HCC. Methods: A retrospective analysis was undertaken at<br />

a national hepatocellular carcinoma (HCC) multidisciplinary<br />

service of all cases of NALFD-related HCCs and matched cases<br />

with chronic viral hepatitis-related HCCs diagnosed between<br />

2001 and 2014. Results: A total of 93 patients with NAFLD-<br />

HCC were diagnosed. These were compared to case-matched<br />

patients with chronic viral hepatitis; HBV or HCV-related HCC<br />

(4:1; n=366). The mean age in NAFLD-HCC was 70.8 and<br />

viral-HCC 58.1 years respectively (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1293A<br />

is a common finding which can be associated with chronic<br />

liver disease but referral practices vary. Aim: to investigate<br />

referral practices, prevalance of liver disease and three, five<br />

and 10 year mortality of minimally deranged ALT. Methods All<br />

LFT results requested by primary care in Newcastle between<br />

1 st January and 31 st December 2003 were collected. Patients<br />

with at least one ALT in the range 41-90IU/L were flagged<br />

and the patients subsequently referred to Gastroenterology or<br />

Hepatology were identified and case notes reviewed. Mortality<br />

was investigated by matching ALT41-90 patients by age,<br />

gender and GP practice to patients with normal range ALT.<br />

Up to date mortality data was collected and three, five and<br />

10 year mortality rates for both groups were calculated and<br />

compared using chi-square tests. Results In 2003, 56,131 LFTs<br />

were checked by primary care in Newcastle. Of these, 5,489<br />

unique patients (9.8%) had an ALT in the range 41-90, while in<br />

986 (1.8%) ALT was >90. Of the ALT41-90 patients, only 331<br />

(6%) were referred to Gastroenterology or Hepatology within a<br />

year. Casenotes were available for review in 309 (93%). 179<br />

(58%) had evidence of significant liver disease and 22 (7%)<br />

had established cirrhosis. Mortality: 1,195 (21.7%) ALT41-<br />

90 patients were matched to normal ALT controls. Mean ALT<br />

was 55.6 (SD 15.9) and 22.1 (SD 7.7) in the ALT41-90 and<br />

normal groups respectively. Mean age was 50.2 (SD 14.1) in<br />

both groups. There was no statistically significant difference<br />

in the three, five or 10 year mortality between the ALT41-90<br />

patients and the normal ALT patients (see table). Conclusions<br />

Minimally deranged ALT is associated with a high risk of liver<br />

pathology in asymptomatic patients but many of these patients<br />

are not referred to a specialist. We did not find that a minimally<br />

deranged ALT was associated with significantly higher<br />

mortality rates at three, five or 10 years but it is possible that<br />

this would be observed after longer follow-up. It would be valuable<br />

to study the relative benefits of other methods of screening<br />

for asymptomatic liver disease, in particular the combination of<br />

aspartate aminotransferase (AST) with ALT in the AST:ALT ratio.<br />

key pathological processes in NAFLD. We sought to determine<br />

whether changes in VLDL profiles predict NAFLD disease progression.<br />

Methods: We evaluated VLDL profiles of127 patients<br />

from a single center NAFLD registry, and examined VLDL size<br />

and subclass particle concentrations in relation with biopsy-determined<br />

NASH and liver fibrosis. Results: NASH was associated<br />

with larger VLDL size, but not higher concentration. In<br />

a multivariate model, every nm increase in mean VLDL size<br />

was associated with 2% increase in the relative risk of NASH<br />

(95% CI 1.01–1.03, p = 0.002). A decrease in small VLDL<br />

particle concentration was associated with more advanced<br />

liver fibrosis. Every nmol/L decrease in the concentration of<br />

small VLDL particles was associated with 2% increase (95% CI<br />

1.003–1.04, p = 0.03) in the relative risk of stage 2 or above<br />

fibrosis. Mean VLDL size and small VLDL particle concentration<br />

alone performed equally well as cytokeratin-18 (CK18) and<br />

NAFLD fibrosis score (NFS) in predicting both NASH and significant<br />

liver fibrosis (Figure). The incorporation of VLDL measurements<br />

improves the predictive value of existing NASH and<br />

fibrosis indicators. Conclusions: The development of NASH is<br />

associated with an increase in mean VLDL size, whereas the<br />

progression of liver fibrosis is associated with a decrease in<br />

the concentration of small VLDL particles. Profiles of circulating<br />

VLDL may be used as biomarkers for NAFLD.<br />

A. Comparison of CK18, VLDL size and the combination of both<br />

in predicting NASH (NAS score ≥ 4). B. Comparison of NFS,<br />

small VLDL concentration and the combination of both in predicting<br />

significant liver fibrosis (stage 2 or above).<br />

Mortality of normal ALT vs ALT41-90<br />

Disclosures:<br />

The following authors have nothing to disclose: James G. Orr, Richard C.<br />

Thomas, David Jones, Mark Hudson<br />

2227<br />

Steatohepatitis and liver fibrosis are predicted by the<br />

size and subclass concentration of very low density lipoprotein<br />

in nonalcoholic fatty liver disease<br />

Zhenghui G. Jiang 1 , Elliot B. Tapper 1 , Margery A. Connelly 2 , Carolina<br />

F. Pimentel 1 , Linda Feldbrügge 1 , Misung Kim 3 , Sarah Krawczyk<br />

3 , Nezam H. Afdhal 1 , Simon C. Robson 1 , Mark A. Herman 3 ,<br />

James Otvos 2 , Kenneth Mukamal 4 , Michelle Lai 1 ; 1 Gastroenterology<br />

and Hepatology, Beth Israel Deaconess Medical Center,<br />

Harvard Medical School, Boston, MA; 2 LabCorp, Raleigh, NC;<br />

3 Endocrinology, Beth Israel Deaconess Medical Center, Harvard<br />

Medical School, Boston, MA; 4 General Internal Medicine, Beth<br />

Israel Deaconess Medical Center, Harvard Medical School, Boston,<br />

MA<br />

Background: A major challenge in the management of nonalcoholic<br />

fatty liver disease (NAFLD) is to identify patients at<br />

risk for disease progression characterized by steatohepatitis<br />

(NASH) and early liver fibrosis. Very low density lipoprotein<br />

(VLDL) is produced exclusively from the liver, and influenced by<br />

Disclosures:<br />

Margery A. Connelly - Employment: LabCorp<br />

Nezam H. Afdhal - Advisory Committees or Review Panels: Trio Helath Care;<br />

Board Membership: Journal Viral hepatitis; Consulting: Merck, EchoSens, BMS,<br />

Achillion, GlaxoSmithKline, Springbank, Gilead, AbbVie; Grant/Research Support:<br />

Gilead; Stock Shareholder: Springbank<br />

Simon C. Robson - Grant/Research Support: Pfizer, NIH, Dainippon; Independent<br />

Contractor: Biolegend, EMD Millipore, Mersana; Management Position:<br />

eBioscience; Speaking and Teaching: ACP, Elsevier, ATC; Stock Shareholder:<br />

Nanopharma, Puretech<br />

Mark A. Herman - Speaking and Teaching: Pfizer<br />

The following authors have nothing to disclose: Zhenghui G. Jiang, Elliot B.<br />

Tapper, Carolina F. Pimentel, Linda Feldbrügge, Misung Kim, Sarah Krawczyk,<br />

James Otvos, Kenneth Mukamal, Michelle Lai


1294A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2228<br />

Relationships between very low density lipoprotein<br />

profiles and pathological processes in nonacoholic fatty<br />

liver disease: insights from the Cardiovascular Health<br />

Study<br />

Zhenghui G. Jiang 1 , Ian H. de Boer 2 , Rachel H. Mackey 3 , Majken<br />

K. Jensen 4 , Michelle Lai 1 , Russel Tracy 3 , Simon C. Robson 1 , Lewis<br />

Kuller 5 , Kenneth Mukamal 6 ; 1 Gastroenterology and Hepatology,<br />

Beth Israel Deaconess Medical Center, Harvard Medical School,<br />

Boston, MA; 2 Nephrology, University of Washington, Seattle, WA;<br />

3 Pathology, University of Vermont College of Medicine, Burlington,<br />

VT; 4 Nutrition, Harvard School of Public Health, Boston, MA;<br />

5 Epidemiology, University of Pittsburgh Graduate School of Public<br />

Health, Pittsburgh, PA; 6 General Internal Medicine, Beth Israel<br />

Deaconess Medical Center, Harvard Medical School, Boston, MA<br />

Background: Nonalcoholic fatty liver disease (NALFD) significantly<br />

impacts the production of very low density lipoprotein<br />

(VLDL) from the liver. As a result circulating VLDL profile<br />

may serve as a marker for NAFLD disease severity. Herein,<br />

we aim to determine the impact on VLDL profiles from three<br />

key pathological processes in NAFLD: insulin resistance, liver<br />

synthetic function and inflammation. Methods: We examined<br />

cross-sectional associations of markers for insulin sensitivity,<br />

inflammation, and liver synthetic function with VLDL particle<br />

(VLDL-P) subclass concentration and size among 1,850 participants<br />

in the Cardiovascular Health Study. Results: Insulin resistance<br />

(lower Matsuda index and 1/HOMA-IR) was associated<br />

with larger mean VLDL size and higher concentrations of large<br />

VLDL-P. Markers of liver synthetic function (albumin, fibrinogen<br />

and factor VII) were positively associated with concentrations<br />

of total and subclass VLDL-P. CRP and IL-6, markers for inflammation,<br />

demonstrated a nonlinear relationship with both total<br />

VLDL-P concentration and VLDL size. When mutually adjusted,<br />

one standard deviation (SD) increment in Matsuda index and<br />

CRP were associated with -4.9 nmol/L (-8.2 - -1.5, p = 0.005)<br />

and -6.3 nmol/L (-11.0 - -1.6, p = 0.009) lower VLDL-P concentration<br />

respectively. In contrast, one-SD increment in factor VII<br />

was associated with 16.9 nmol/L (12.6 – 21.2, p < 0.001)<br />

higher VLDL-P concentration. In addition, a one-SD increment in<br />

the Matsuda index was associated with -1.1 nm (-2.0 - -0.3, p<br />

= 0.006) smaller mean VLDL size, whereas CRP and factor VII<br />

were not associated with VLDL size. Conclusion: Insulin resistance<br />

is associated with higher VLDL-P concentration and larger<br />

mean VLDL size, whereas both inflammation and impaired liver<br />

synthetic function are associated with lower concentrations of<br />

VLDL-P, but have no relationship with VLDL size. These differential<br />

impacts on VLDL profile set the foundation for the utility of<br />

VLDL in disease staging of NAFLD.<br />

Adjusted regression coefficients for VLDL profiles from standardized<br />

Matsuda index, CRP and factor VII<br />

Regression models calculated using the Z-scores of three primary<br />

exposures (Mastuda index, CRP, factor VII) in units of standard<br />

deviation (SD), and covariates including age, gender, ethnicity,<br />

BMI, alcohol, smoking history and the use of lipid lowering medications.<br />

The following authors have nothing to disclose: Zhenghui G. Jiang, Ian H. de<br />

Boer, Rachel H. Mackey, Majken K. Jensen, Michelle Lai, Russel Tracy, Lewis<br />

Kuller, Kenneth Mukamal<br />

2229<br />

WITHDRAWN<br />

2230<br />

The development of hepatocellular carcinoma in Japanese<br />

patients with biopsy-proven nonalcoholic fatty liver<br />

disease : the relation between carriage of the PNPLA3<br />

rs738409 C >G polymorphism and hepatocarcinogenesis<br />

Yuya Seko 1 , Yoshio Sumida 1 , Saiyu Tanaka 2 , Hiroyoshi Taketani 1 ,<br />

Kazuyuki Kanemasa 2 , Hiroshi Ishiba 1 , Akira Okajima 1 , Tasuku<br />

Hara 1 , Kanji Yamaguchi 1 , Michihisa Moriguchi 1 , Kohichiroh<br />

Yasui 1 , Hironori Mitsuyoshi 1 , Yoshito Itoh 1 ; 1 Kyoto Prefectural University<br />

of Medicine, Kyoto, Japan; 2 Center for Digestive and Liver<br />

Diseases, Nara City Hospital, Nara, Japan<br />

BACKGROUND: Some patients with nonalcoholic fatty liver disease<br />

(NAFLD) develop hepatocellular carcinoma (HCC). Carriage<br />

of the PNPLA3 rs738409 (encoding the I148M variant)<br />

has been reported to be associated with advanced fibrosis<br />

and HCC related to NAFLD. Our aim was to determine the<br />

incidence of and risk factors including PNPLA3 rs738409 polymorphism<br />

for HCC in Japanese patients with biopsy-proven<br />

NAFLD. METHODS: This retrospective cohort study analyzed<br />

hepatocarcinogenesis in 341 patients with biopsy-proven<br />

NAFLD. PNPLA3 rs738409 genotype was determined by allelic<br />

discrimination in 133 patients. RESULTS: Of 341 patients, 205<br />

(60.1%) were diagnosed with nonalcoholic steatohepatitis. The<br />

PNPLA3 rs738409 genotype frequencies were CC 0.14, CG<br />

0.46, GG 0.40. During a median follow-up period of 5.0<br />

years, 9 patients (2.6%) developed HCC. The cumulative rate<br />

of HCC was 1.5% at the end of the 5th year and 7.1% at the<br />

end of the 10th year. Multivariate analysis identified PNPLA3<br />

genotype (GG; hazard ratio [HR] 14.2, p=0.035), and platelet<br />

count (^15x10/μl; HR 11.6, p=0.036) as predictors for the<br />

development of HCC. PNPLA3 genotype was analysed in 9<br />

patients with HCC, 7 were PNPLA3 GG and 2 were CG. In<br />

patients with GG allele, the cumulative rate of HCC was 4.9%<br />

at the end of the 5th year and 33.5% at the end of the 10th<br />

year. The incidence of HCC was significantly higher in patients<br />

with PNPLA3 GG allele than in those with CC and CG allele<br />

(p=0.027). CONCLUSIONS: Severe fibrosis was a predictor<br />

for HCC in NAFLD patients. And PNPLA3 polymorphism genotyping<br />

was also the risk to the development of HCC. These<br />

findings might help us HCC surveillance in NAFLD patients.<br />

Disclosures:<br />

Yoshito Itoh - Grant/Research Support: MSD KK, Bristol-Meyers Squibb, Dainippon<br />

Sumitomo Pharm. Co., Ltd., Otsuka Pharmaceutical Co., Chugai Pharm<br />

Co., Ltd, Mitsubish iTanabe Pharm. Co.,Ltd., Daiichi Sankyo Pharm. Co.,Ltd.,<br />

Takeda Pharm. Co.,Ltd., AstraZeneca K.K.:, Eisai Co.,Pharm.Ltd, FUJIFILM Medical<br />

Co.,Ltd., Gelaed Sciences Co., GlaxoSmithKline<br />

The following authors have nothing to disclose: Yuya Seko, Yoshio Sumida, Saiyu<br />

Tanaka, Hiroyoshi Taketani, Kazuyuki Kanemasa, Hiroshi Ishiba, Akira Okajima,<br />

Tasuku Hara, Kanji Yamaguchi, Michihisa Moriguchi, Kohichiroh Yasui, Hironori<br />

Mitsuyoshi<br />

Disclosures:<br />

Simon C. Robson - Grant/Research Support: Pfizer, NIH, Dainippon; Independent<br />

Contractor: Biolegend, EMD Millipore, Mersana; Management Position:<br />

eBioscience; Speaking and Teaching: ACP, Elsevier, ATC; Stock Shareholder:<br />

Nanopharma, Puretech


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1295A<br />

2231<br />

High prevalence of undiagnosed liver cirrhosis and<br />

advanced fibrosis in type 2 diabetic outpatients<br />

Francisco Barrera, Juan P. Arab, Carlos E. Benítez, Arnoldo<br />

Riquelme, Marco Arrese; Department of Gastroenterology, Escuela<br />

de Medicina, Pontificia Universidad Catolica de Chile, Santiago,<br />

Chile<br />

BACKGROUND: Patients with Type 2 Diabetes Mellitus (T2DM)<br />

are at risk for developing end-stage liver disease due to nonalcoholic<br />

steatohepatitis (NASH). Non-invasive methods have<br />

been validated for assessing the severity of nonalcoholic fatty<br />

liver disease (NAFLD). Data on prevalence of advanced fibrosis<br />

among T2DM patients evaluated by these methods is scarce.<br />

AIM: To evaluate prevalence of steatosis, advanced fibrosis<br />

and cirrhosis by non-invasive methods in T2DM patients.<br />

METHODS: We conducted a cross-sectional study in 145 consecutive<br />

T2DM patients (>55 years-old). The presence of cirrhosis<br />

and advanced fibrosis was evaluated by liver morphology<br />

assessed by magnetic resonance imaging (MRI) and NAFLD<br />

fibrosis score (NFS) respectively. Exclusion criteria included<br />

significant alcohol consumption, viral hepatitis or other liver<br />

diseases and exposure to hepatotoxic agents. RESULTS: 52.6%<br />

were women, average age was 60 years (57-64), BMI was<br />

29.6±4.7 kg/m 2 and diabetes duration was 7.6±6.9 years.<br />

A high prevalence of liver steatosis (62.4%) and steatosis<br />

and abnormal ALT (37.6%) was found. The prevalence of<br />

advanced fibrosis using the NFS was 12.8 and evidence of<br />

liver cirrhosis on MRI was 5.3%. In a multivariate analysis GGT<br />

>82 IU/L (P=0.004) and no alcohol intake (P=0.032) were<br />

independently associated to advanced fibrosis. CONCLUSION:<br />

A high frequency of undiagnosed advanced fibrosis and cirrhosis<br />

was observed in otherwise unselected T2DM patients older<br />

than 55 y/o. These patients are at high risk of developing liver-related<br />

complications such as portal hypertension and hepatocellular<br />

carcinoma. Routine screening for liver disease should<br />

be considered in this population (Grant Support: FONDECYT<br />

1150327 to M.A.and 1150311 to F.B) .<br />

Disclosures:<br />

The following authors have nothing to disclose: Francisco Barrera, Juan P. Arab,<br />

Carlos E. Benítez, Arnoldo Riquelme, Marco Arrese<br />

2232<br />

Discovery of plasma biomarkers for NAFLD using high<br />

resolution metabolomics<br />

Miriam B. Vos, Ran Jin, Sophia Banton, Shuzhao Li, Dean P. Jones;<br />

Emory University, Atlanta, GA<br />

Purpose: Past attempts to develop diagnostic tests for NAFLD<br />

using single metabolite biomarkers have failed in part because<br />

of the heterogeneity of NAFLD. High throughput, high resolution<br />

metabolomics is rapid and relatively inexpensive, making it a<br />

promising platform for disease identification and characterization.<br />

Methods: We performed an exploratory, unbiased metabolomics<br />

pilot study to develop a biomarker panel for detecting<br />

NAFLD. Fasting plasma samples of 39 Hispanic-American adolescents<br />

with either MRS-documented NAFLD (intrahepatic fat ><br />

5%) or age and BMI matched control individuals (intrahepatic<br />

fat < 5%) were analyzed using LC-MS high-resolution metabolomics.<br />

A total of 9,583 metabolite features were yielded<br />

initially. The top discriminatory and identifiable metabolites<br />

were selected. Multivariate receiver operating characteristic<br />

(ROC) curves were generated for individual metabolites and<br />

combinations of metabolites, respectively. Results: Univariate<br />

ROC curves >0.80 were observed with 7 of the top metabolites,<br />

which included metabolites from de novo lipogenesis,<br />

ketogenic amino acids and hormones. Multivariate ROC analysis<br />

with cross validation using balanced subsampling demonstrated<br />

that combinations of the top 7, 10, and 16 selected<br />

metabolites had higher prediction for NAFLD (AUC>0.9 for all<br />

panels, Figure 1) compared to single metabolites. Conclusions:<br />

This pilot “omics”-based biomarker development study shows<br />

feasibility in using a metabolomics based biomarker panel to<br />

identify NAFLD with high sensitivity and specificity. Independent<br />

validation <strong>studies</strong> are warranted to confirm clinical utility.<br />

Disclosures:<br />

The following authors have nothing to disclose: Miriam B. Vos, Ran Jin, Sophia<br />

Banton, Shuzhao Li, Dean P. Jones<br />

2233<br />

Polychlorinated biphenyl exposure in Anniston, Alabama<br />

is associated with elevated prevalence of liver<br />

disease<br />

Heather B. Clair 1 , Keith C. Falkner 1 , Russell A. Prough 2 , Christina<br />

Pinkston 1 , Guy N. Brock 1 , Marian Pavuk 3 , N. Dutton 3 , Matthew<br />

C. Cave 1,4 ; 1 Department of Medicine/GI, University of Louisville,<br />

Louisville, KY; 2 Biochemistry, University of Louisville, Louisville, KY;<br />

3 CDC, Atlanta, GA; 4 Robley Rex VAMC, Louisville, KY<br />

Purpose: Elevated serum concentrations of Polychlorinated<br />

Biphenyls (PCBs) are associated with increased liver injury<br />

in the general US population. Residents of Anniston, AL are<br />

highly exposed to PCBs, resulting in a mean serum PCB concentrations<br />

3-4 times the mean of the general US population.<br />

An adult cohort of Anniston residents (Anniston Community<br />

Health Survey – ACHS) was assembled to study health effects<br />

and PCB exposure. We report elevated liver injury biomarkers<br />

characteristic of Toxicant-Associated Steatohepatitis (TASH),<br />

cytokine and adipokine abnormalities. A resampling of this<br />

population provided longitudinal data related to liver disease,<br />

metabolic abnormalities and total PCB load, as well as an<br />

expanded analysis of the dioxin-like PCB congener concentration<br />

in Anniston residents. Methods: Subjects were recruited<br />

from residences within the city limits of Anniston, Alabama.<br />

Serum samples from 738 of the individuals recruited in 2005-<br />

2007 (ACHS1), and 352 individuals re-sampled in 2014<br />

(ACHS2) were included in our analysis. In both groups, whole<br />

(CK18 M65) and caspase-cleaved cytokeratin 18 (CK18 M30)<br />

were measured using the PEVIVA ELISAs. Serum cytokine/


1296A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

adipokines were measured using the Luminex cytokine bead<br />

array. We also analyzed ACHS2 samples for clinical indicators<br />

of liver injury. Results: ACHS1 was stratified into three<br />

groups based on CK18: no liver disease (None, M30


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1297A<br />

0.838±0.035 and 0.816±0.036 respectively, FIB4 (cut-off<br />

1.4, Se=77%, Sp=79.8%, and DA=78.6%) and FM-NAFLD<br />

(cut-off 0.6; Se=68.9%, Sp=85.7%, and DA=78.6%) outperformed<br />

other models in predicting moderate (F2) hepatic<br />

fibrosis (P15 e ≤15. Results:<br />

The sample included 37severe obese patients. The mean age<br />

was 38.1(SD=9.8) and 51.4% were women. Steatosis was<br />

observed in 43.2% (16) of them and NASH with and without<br />

fibrosis in 57% (21). Comparing the 2 groups, patients<br />

with NASH presented higher levels of ROS [Md 1<br />

: 11621<br />

(IIQ:954-30033) versus Md 2<br />

: 2485 (IIQ:1206-5564) photons,<br />

p


1298A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2238<br />

Non-alcoholic fatty liver disease: assessment of intestinal<br />

microbiota and metabolites<br />

Hannah Da Silva 1,2 , Anastasia Teterina 1 , Bianca M. Arendt 1 ,<br />

Marialena Mouzaki 3 , Sandra E. Fischer 4 , Scott Fung 5 , Johane P.<br />

Allard 5 ; 1 Gastroenterology, Toronto General Hospital, Toronto,<br />

ON, Canada; 2 Department of Nutritional Sciences, University of<br />

Toronto, Toronto, ON, Canada; 3 GI/Hepatology/Nutrition, Hospital<br />

for Sick Children, Toronto, ON, Canada; 4 Pathology, University<br />

Health Network, Toronto, ON, Canada; 5 Medicine, Toronto<br />

General Hospital, Toronto, ON, Canada<br />

Background: Intestinal microbiota (IM) may contribute to non-alcoholic<br />

fatty liver disease (NAFLD) through products of bacterial<br />

metabolism. This study aimed to characterize the bacterial<br />

products present in serum and stool of NAFLD patients and<br />

HC and to examine the relationship between these products<br />

and IM. Methods: This was a prospective, cross-sectional<br />

study. Fecal and serum metabolite profiles of 39 patients with<br />

biopsy-proven NAFLD and 28 living liver donors as HC were<br />

examined using nuclear magnetic resonance spectroscopy.<br />

Metabolic profiles included 11 fecal and 9 serum bacterial<br />

products. In a subset of 29 NAFLD and 16 HC, IM composition<br />

was assessed. Clinical, biochemical, and anthropometric<br />

data were also collected. Quantitative real-time polymerase<br />

chain reaction was used to measure total bacterial counts,<br />

Firmicutes:Bacteroidetes ratio, Bacteroidetes, C. leptum, C.<br />

coccoides, Bifidobacteria, E. coli and Archaea in stool. Continuous<br />

variables were non-normal and described as median<br />

(Q1, Q3). Groups were compared using Wilcoxon rank-sum<br />

tests. Spearman correlation coefficients were used to examine<br />

association between IM and bacterial products. Results:<br />

NAFLD patients were older than HC with higher BMI, liver<br />

enzymes (AST, ALT), insulin resistance (HOMA-IR), hemoglobin<br />

A1c, and blood triglycerides (p


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1299A<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

Rohit Loomba - Advisory Committees or Review Panels: Galmed Inc, Tobira Inc,<br />

Arrowhead Research Inc; Consulting: Gilead Inc, Corgenix Inc, Janssen and<br />

Janssen Inc, Zafgen Inc, Celgene Inc, Alnylam Inc, Inanta Inc, Deutrx Inc; Grant/<br />

Research Support: Daiichi Sankyo Inc, AGA, Merck Inc, Promedior Inc, Kinemed<br />

Inc, Immuron Inc, Adheron Inc<br />

Stephen H. Caldwell - Advisory Committees or Review Panels: Vital Therapy;<br />

Grant/Research Support: Genfit, Gilead Sciences, Immuron, Hyperion, Immuron,<br />

NGM<br />

Reem H. Ghalib - Grant/Research Support: Bristol Myers Squibb Pharmaceuticals,<br />

Vertex Pharmaceuticals, Janssen, Merck, Genentech, Idenix, Zymogenetics,<br />

Pharmasset, Anadys, Duke Clinical Research Institute, Achillion, Boehringer Ingelheim,<br />

Gilead Pharmaceuticals, Virochem Pharmaceuticals, Abbott, Medtronic<br />

Inc, Novartitis, Roche, Schering Plough, Salix, tibotec, Inhibitex, Takeda, Abbvie<br />

Dawn M. Torres - Advisory Committees or Review Panels: Genetech; Speaking<br />

and Teaching: Merck<br />

Andrew J. Muir - Advisory Committees or Review Panels: BMS, Gilead, Janssen,<br />

Merck; Consulting: Theravance; Grant/Research Support: Abbvie, Abbvie, BMS,<br />

Gilead, Janssen, Merck, Achillion, Lumena<br />

Robert P. Myers - Employment: Gilead Sciences, Inc.; Stock Shareholder: Gilead<br />

Sciences, Inc.<br />

Raul E. Aguilar Schall - Employment: Gilead Sciences, Inc.<br />

Mani Subramanian - Employment: Gilead Sciences<br />

John G. McHutchison - Employment: Gilead Sciences; Stock Shareholder: Gilead<br />

Sciences<br />

Jaime Bosch - Consulting: Falk, Gilead Science, Intercept Therapeutics, Conatus<br />

Pharmaceuticals, Exalenz, Almirall, Chiasma<br />

Stephen A. Harrison - Advisory Committees or Review Panels: Merck, Nimbus<br />

Discovery, Fibrogen, RuiYi, CLDF; Consulting: NGM Biopharmaceuticals; Speaking<br />

and Teaching: Gilead, Abbvie, Janssen, CLDF<br />

Manal F. Abdelmalek - Consulting: Islet Sciences; Grant/Research Support:<br />

Tobira, Gilead Sciences, NIH/NIDDK, Synageva, Genfit Pharmaceuticals,<br />

Immuron, Galmed, TaiwanJ Pharma, Intercept, NGM Pharmaceuticals<br />

The following authors have nothing to disclose: Arthur J. McCullough<br />

2240<br />

The effectiveness of SGLT-2 inhibitors as second-line<br />

treatments for NAFLD patients with type 2 diabetes<br />

mellitus who failed to incretin based therapy including<br />

GLP-1 analogues and DPP-4 inhibitors<br />

Takamasa Ohki 1 , Isogawa Akihiro 2 , Nobuo Toda 1 ; 1 Gastroenterology,<br />

Mitsui Memorial Hospital, Tokyo, Japan; 2 Diabetes and<br />

Metabolism, Mitsui Memorial Hospital, Tokyo, Japan<br />

Background: We already reported that incretin based medicine,<br />

such as GLP-1 analogues or DPP-4 inhibitors, improved<br />

glycaemic control and liver inflammation in non-alcoholic fatty<br />

liver disease (NAFLD) patients with type 2 diabetes mellitus<br />

(T2DM). However, the effect on ALT normalization was still<br />

limited. Aims: The aim of this study is to elucidate the effectiveness<br />

of sodium-glucose co-transporter 2 (SGLT-2) inhibitors<br />

as second-line treatments for NAFLD patients with T2DM who<br />

failed to incretin based therapy. Methods: We retrospectively<br />

enrolled consecutive 130 Japanese NAFLD patients with T2DM<br />

who were treated with GLP-1 analogues or DPP-4 inhibitors.<br />

Among them, 70 (53.8%) patients achieved normal ALT level<br />

and the rest 60 (46.2%) patients did not. We finally obtained<br />

informed consent from 24 (40.0%) patients out of 60 patients<br />

and administered SGLT-2 inhibitors in addition to GLP-1 analogues<br />

or DPP-4 inhibitors. We compared the changes of laboratory<br />

data including ALT level and body weight at the end of<br />

follow-up. Results: Thirteen patients were used in combination<br />

with DPP-4 inhibitors and rest 11 patients were used in combination<br />

with GLP-1 analogues. The median dosing period was<br />

320 days. At the end of follow-up, body weight significantly<br />

decreased (84.8 kg to 81.7 kg, P < 0.01) with amelioration<br />

of HbA1c level (8.4% to 7.6%, P < 0.01). Serum ALT level<br />

also significantly decreased (62 IU/l to 49 IU/l, P < 0.01)<br />

with improvement of FIB-4 index (1.75 to 1.39, P = 0.04).<br />

Univariate analysis indicated that 3% reduction of body weight<br />

(OR: 12.0, 95%CI; 1.18-122, P = 0.04) as a factor which<br />

contributed to normalization of serum ALT level. Conclusions:<br />

Administration of SGLT-2 inhibitors led not only good glycaemic<br />

control but also reduction of body weight, normalization of<br />

ALT level, and alteration of FIB-4 index, even in patients who<br />

failed to incretin based therapy.<br />

Disclosures:<br />

Takamasa Ohki - Speaking and Teaching: Otsuka Pharmaceutical Co. Ltd.<br />

The following authors have nothing to disclose: Isogawa Akihiro, Nobuo Toda<br />

2241<br />

The impact of alcohol consumption for hepatocarcinogenesis<br />

in Japanese fatty liver disease patients<br />

Yusuke Kawamura 1,2 , Yasuji Arase 1,3 , Kenji Ikeda 1 , Yushi Sorin 1 ,<br />

Hideo Kunimoto 1 , Hitomi Sezaki 1 , Tetsuya Hosaka 1 , Norio Akuta 1 ,<br />

Masahiro Kobayashi 1 , Satoshi Saitoh 1 , Fumitaka Suzuki 1 , Yoshiyuki<br />

Suzuki 1 , Mie Inao 2 , Satoshi Mochida 2 , Hiromitsu Kumada 1 ;<br />

1 Hepatology, Toranomon Hospital, Tokyo, Japan; 2 Gastroenterology<br />

and Hepatology, Saitama Medical University, Saitama,<br />

Japan; 3 Health Management Center, Toranomon Hospital, Tokyo,<br />

Japan<br />

Background & Aims: The effect of ethanol consumption for<br />

hepatocarcinogeneis of fatty liver is not clear. The aims of this<br />

study were to determine the influence of alcohol consumption<br />

on hepatocarcinogenesis and the risk factors for hepatocellular<br />

carcinoma (HCC) in a large population of Japanese fatty liver<br />

patients without viral hepatitis. Methods: This study was retrospective<br />

cohort study and setting was specialized center for<br />

hepatology. The study subjects were 9,959 patients with fatty<br />

liver disease (FLD) without viral hepatitis diagnosed by ultrasonography.<br />

In this study, the levels of daily ethanol consumption<br />

divided into following four category;


1300A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

Disclosures:<br />

Kenji Ikeda - Speaking and Teaching: Eisai company, Dainippon Sumitomo<br />

Pharmaceutical company<br />

Norio Akuta - Patent Held/Filed: SRL. Inc.; Speaking and Teaching: Mitsubishi<br />

Tanabe Pharma, Janssen Pharmaceutical K.K., Bristol-Myers Squibb<br />

Fumitaka Suzuki - Speaking and Teaching: BMS<br />

Yoshiyuki Suzuki - Speaking and Teaching: Bristol-Myers Squibb<br />

Satoshi Mochida - Grant/Research Support: Chugai, MSD, Tioray Medical,<br />

BMS; Speaking and Teaching: MSD, Toray Medical, BMS, Tanabe Mitsubishi<br />

Hiromitsu Kumada - Patent Held/Filed: SRL; Speaking and Teaching: Bristol-Myers<br />

Squibb,Pharma International, MSD, Janssen Pharma., Glaxosmithkline<br />

The following authors have nothing to disclose: Yusuke Kawamura, Yasuji<br />

Arase, Yushi Sorin, Hideo Kunimoto, Hitomi Sezaki, Tetsuya Hosaka, Masahiro<br />

Kobayashi, Satoshi Saitoh, Mie Inao<br />

2242<br />

WITHDRAWN<br />

2243<br />

Differences of lifestyle behaviors between non-obese<br />

and obese non-alcoholic fatty liver disease; beyond visceral<br />

obesity and metabolic syndrome<br />

Joo Hee Kwak 1 , Dae Won Jun 1 , Seung Min Lee 2 , Yong Kyun<br />

Cho 3 , Eun Chul Jang 4 ; 1 Department of Internal Medicine, Hanyang<br />

University College of Medicine, Seoul, Korea (the Republic of);<br />

2 Department of Food and Nutrition, Sungshin Women’s University,<br />

Seoul, Korea (the Republic of); 3 Department of Internal Medicine,<br />

Kangbuk Samsung Hospital, Sungkyunkwan University, School of<br />

Medicine, Seoul, Korea (the Republic of); 4 Department of Occupational<br />

and Environmental Medicine, Soonchunhyang University<br />

Cheonan Hospital, Cheonan, Korea (the Republic of)<br />

Background and Aims: Best way to treat non-alcoholic fatty<br />

liver disease (NAFLD) is weight reduction. bu However, strategy<br />

of lifestyle modification in non-obese NAFLD patient is<br />

unclear. The aim of study was to investigate differences of<br />

lifestyle behaviors between non-obese and obese non-alcoholic<br />

fatty liver disease. Methods: This study has been performed<br />

with 209 patients who wanted to participate in nutrition education<br />

program because of their abnormal liver enzyme or<br />

fatty liver. All participants undertook computed tomography.<br />

NAFLD was diagnosed when liver/spleen Hounsfield Unit (HU)<br />

was less than 1.1. Five day food record and physical activity<br />

was measured by two dietitians. Results: Prevalence of NAFLD<br />

was significantly higher as 66.1% in obese subjects (BMI >25)<br />

than 42.4% in normal weight (BMI≤25) (p=0.001). Non-obese<br />

NAFLD group showed higher visceral fat area, aminotransferase<br />

activity, and fasting glucose compare to normal liver/<br />

spleen HU subjects. In univariate analysis, non-obese NAFLD<br />

group showed higher carbohydrate intake (281 g/day vs. 246<br />

g/day), and did lower moderate level exercise compare to<br />

control group (p=0.013). But total calorie intake (1,955 Kcal<br />

vs. 1,837 Kcal) and walking time/week were not difference.<br />

In multivariate analysis, the amount of carbohydrate (p=0.018)<br />

and more than 2 hours of exercise a week (p=0.010) were<br />

risk factors for NAFLD independent with visceral fat area and<br />

total calorie intake. Meanwhile fasting glucose (p25). Conclusions: The moderate level exercise of 2 hours a<br />

week and the amount of carbohydrate intake were independent<br />

risk factor in non-obese NAFLD patients.<br />

Disclosures:<br />

The following authors have nothing to disclose: Joo Hee Kwak, Dae Won Jun,<br />

Seung Min Lee, Yong Kyun Cho, Eun Chul Jang<br />

2244<br />

Subtyping nonalcoholic steatohepatitis for the development<br />

of effective treatments<br />

Ibon Martinez-Arranz 1 , Cristina Alonso 1 , David Fernández 2 ,<br />

Rebeca Mayo 1 , Marta Varela 2 , Mazen Noureddin 3 , Maria L. Martinez-Chantar<br />

2 , Shelly C. Lu 3 , Azucena Castro 1 , Jose M. Mato 2 ;<br />

1 OWL, Derio, Spain; 2 CIC bioGUNE, Ciberehd, Derio, Spain;<br />

3 Division of Gastroenterology, Cedars-Sinai Medical Center, Los<br />

Angeles, CA<br />

NASH (nonalcoholic steatohepatitis) is a histopathological<br />

definition that groups together defects in diverse biochemical<br />

processes that cause hepatic fat accumulation, inflammation<br />

and necrosis as it were a single disease. The identification<br />

of the types of mechanisms leading to NASH, together with<br />

the development of noninvasive biomarkers of these different<br />

NASH subtypes, is central for the development of precision<br />

treatments, since the pathways that should be targeted will<br />

depend upon the specific subtype of NASH. For tissues whose<br />

primary physiological role is metabolic, such as the liver,<br />

metabolites are unquestionably the best indicators of organ<br />

function. Therefore, the detailed analysis of serum metabolome<br />

may reveal metabolic profiles that may be used as disease<br />

biomarkers. We previously found that hepatic methionine adenosyltransferase<br />

(MAT, the enzyme that synthesizes S-adenosylmethionine<br />

or SAMe) deletion (Mat1a KO) in mice led to the<br />

spontaneous development of NASH and liver cancer. NASH<br />

patients often show reduced expression of MAT1A, indicating<br />

that the Mat1a KO mouse model is relevant to study human<br />

NASH. Here we have analyzed the serum lipidome (over 400<br />

different molecular species) in 10-11 month old Mat1a KO<br />

with histological diagnose of NASH and wild type mice with<br />

normal liver (twelve animals per group), and selected the top<br />

fifty serum lipids that more significantly differentiated between<br />

both genotypes. Then, we observed that this serum biomarker<br />

panel robustly classified a cohort of 387 patients with biopsy<br />

proven NAFLD into two well-defined clusters, based on optimum<br />

average silhouette width, with around 66% of the patients<br />

showing a MAT1A metabolic subtype (M-subtype). Since there<br />

is considerable interest in the utility of SAMe to ameliorate<br />

NAFLD, these results suggest that NAFLD patients with a serum<br />

M-subtype are best suited to be selected to study the efficacy<br />

of SAMe treatment.<br />

Disclosures:<br />

The following authors have nothing to disclose: Ibon Martinez-Arranz, Cristina<br />

Alonso, David Fernández, Rebeca Mayo, Marta Varela, Mazen Noureddin,<br />

Maria L. Martinez-Chantar, Shelly C. Lu, Azucena Castro, Jose M. Mato


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1301A<br />

2245<br />

Non-alcoholic fatty liver is not associated with incident<br />

chronic kidney disease: a large histology based case<br />

controlled study<br />

Neeraj Saraf, Narendra S. Choudhary, Sanjiv Saigal, Naveen<br />

Kumar, Rahul Rai, Dheeraj Gautam, Arvinder Soin; Hepatology,<br />

Medanta Medicity, Gurgaon, India<br />

Background: Non-alcoholic steatohepatitis or fibrosis is associated<br />

with the increased prevalence of impaired kidney function.<br />

It is not known whether non alcoholic fatty liver (NAFL)<br />

which is steatosis without inflammation or fibrosis is associated<br />

with renal impairment as these subjects are not candidates for<br />

liver biopsy. Materials and methods: The study group included<br />

all liver donors who underwent donor hepatectomy at our centre<br />

and had a preoperative liver biopsy for various reasons<br />

like, high body mass index, dyslipidemia, presence of metabolic<br />

syndrome or evidence of liver steatosis on imaging. All<br />

the subjects had negative viral markers. Non-alcoholic fatty<br />

liver (NAFL) was defined as >5% hepatocytes having steatosis<br />

and no changes of steatohepatitis and/or fibrosis. Subjects<br />

with NAFL were compared with subjects with normal liver histology.<br />

Estimated glomerular filtration rate was calculated with<br />

Modification of Diet in Renal Disease (MDRD) and CKD-EPI<br />

method. All subjects had assessment for proteinuria and serum<br />

creatinine. Results: The study group aged 35±10.3 years and<br />

mean BMI was 26±3.2 kg/m 2 . One hundred and eighty seven<br />

adults having NAFL (80 males) were compared to 186 (88<br />

males) subjects with normal liver histology (controls). Subjects<br />

with steatosis had significantly higher body mass index<br />

(26.8±3.5 versus 25.5±3.8 kg/m 2 , P


1302A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

CVC and PGZ exposures were slightly lower when both drugs<br />

were co-administered, with a modest decrease in steady-state<br />

C max<br />

and AUC 0–tau<br />

for both drugs, whereas C min<br />

remained<br />

unchanged; GMRs for both drugs were ≥0.80 (Table). Effects<br />

of co-administration on PGZ M-III and M-IV metabolites were<br />

less pronounced, with 90% CI for systemic exposure ratios<br />

within the 0.80–1.25 ‘no effect’ range for all 3 PK parameters<br />

(data not shown). Combination treatment was well tolerated,<br />

with no serious AEs or AEs leading to discontinuation. All AEs<br />

were of mild severity and the 2 most commonly reported were<br />

headache and fatigue. Conclusions: Co-administration of CVC<br />

and PGZ was well tolerated and resulted in a modest interaction<br />

that was not considered clinically significant, which suggests<br />

that dose adjustment is not required when both drugs are<br />

used in combination.<br />

(62.1%) male, with a mean age of 57.7 years. 54 (47%) had<br />

not ultrasonography signals of NAFLD. Between those with<br />

NAFLD (n=61; 53.1%), 34 (29,6%) had NAFLD grade I, 19<br />

(16,5%) grade II and 8 (7%) grade III. 26 (22.4%) patients<br />

had no signals of CAD on their coronary angiography. In 50<br />

(43.1%), serious injury has found. NAFLD showed correlation<br />

with CAD in 57 (64.04%) patients (p 25kg/m 2 ), and 15 patients with<br />

alcoholic (ASH) cirrhosis. Hemostatic status was assessed by<br />

basal and agonist-induced platelet activation using flow cytometry,<br />

thrombin generation testing, thromboelastography (TEG),<br />

a plasma-based clot lysis assay, and an analysis of fibrin pore<br />

structure by permeation analyses. Results: Basal and agonist-induced<br />

platelet activation were comparable between patients<br />

with NAFLD and controls, but agonist-induced platelet activation<br />

was decreased in patients with cirrhosis. Thrombomodulin-modified<br />

thrombin generation was comparable between<br />

patients and controls, although patients with cirrhosis had a<br />

reduced anticoagulant response to thrombomodulin. TEG test<br />

results were comparable between controls and non-cirrhotic<br />

NAFLD patients, but revealed moderate hypocoagulability<br />

in cirrhosis. Plasma fibrinolytic potential was decreased in<br />

NAFLD, but accelerated fibrinolysis was observed in ASH cirrhosis.<br />

Clot permeability was decreased in overweight controls<br />

and patients with NAFLD. Platelet activity, thrombin generation,<br />

and clot lysis test results showed a higher variability in<br />

patients than in controls, suggesting individual patients may<br />

have a more thrombogenic hemostatic profile. Conclusions: The


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1303A<br />

combined results of this study show that the overall hemostatic<br />

status is comparable between patients with NAFLD and controls,<br />

which contrasts with previously published results in similar<br />

populations. However, our data show some pro-thrombotic<br />

features in patients with NAFLD, including hypofibrinolysis and<br />

a pro-thrombotic structure of the fibrin clot, the latter of which<br />

appears driven by obesity rather than NAFLD. Our study suggests<br />

that the role for hyperactive hemostasis in the increased<br />

risk of thrombosis in patients with NAFLD is probably limited.<br />

Disclosures:<br />

Arun J. Sanyal - Advisory Committees or Review Panels: Bristol Myers, Gilead,<br />

Genfit, Abbott, Ikaria, Exhalenz; Consulting: Salix, Immuron, Exhalenz, Nimbus,<br />

Genentech, Echosens, Takeda, Merck, Enanta, Zafgen, JD Pharma, Islet<br />

Sciences; Grant/Research Support: Salix, Genentech, Intercept, Ikaria, Takeda,<br />

GalMed, Novartis, Gilead, Tobira; Independent Contractor: UpToDate, Elsevier<br />

The following authors have nothing to disclose: Wilma Potze, Mohammad S.<br />

Siddiqui, Sherry L. Boyett, Jelle Adelmeijer, Kalyani Daita, Ton Lisman<br />

2250<br />

Characteristics and Clinical Outcomes of A Diverse<br />

Cohort of Patients with Biopsy-Proven Non-Alcoholic<br />

Fatty Liver Disease (NAFLD)<br />

Benjamin Yip 1 , Brittany E. Yee 2,4 , Ailinh Do 3,4 , Nghia H.<br />

Nguyen 2,4 , Joseph K. Hoang 4 , Mindie H. Nguyen 4 ; 1 Department<br />

of Internal Medicine, University of California, Irvine, Orange, CA;<br />

2 Department of Medicine, University of California, San Diego, San<br />

Diego, CA; 3 School of Medicine, Texas Tech University, El Paso,<br />

TX; 4 Division of Gastroenterology and Hepatology, Stanford University<br />

Medical Center, Palo Alto, CA<br />

PURPOSE: NAFLD is the most common cause of chronic liver<br />

disease in the U.S. and is associated with increased rates of<br />

all-cause mortality with known ethnic disparities, though little is<br />

known about NAFLD in Asian populations. This study aims to<br />

evaluate presentation and outcomes of patients with NAFLD in<br />

a cohort of ethnically diverse U.S. cohort. METHODS: We identified<br />

a cohort of 696 consecutive Asian (n=71) and non-Asian<br />

(n=625: 41% White, 16% Hispanic, 3% Black, and 30% other)<br />

patients who underwent liver biopsy showing at least mild steatosis<br />

(greater than or equal to 5%) who were evaluated at a<br />

university hospital. Patients with significant alcohol use (> 2<br />

drinks a day or > 14 drinks a week) were excluded. RESULTS:<br />

The majority (67%) were female with a median age of 49<br />

(18-82). The average alcohol intake was 0.29 ± 0.44 drinks<br />

a day. Concurrent liver diseases included chronic viral hepatitis<br />

in 12% and other liver disease (including autoimmune,<br />

iatrogenic, and hereditary liver disease) in 5%. Hypertension<br />

was common (64%) as well as dyslipidemia (49%) and diabetes<br />

mellitus (23%). About half underwent liver biopsy during<br />

another surgical intervention (43% were for weight reduction<br />

and 10% others) with most of the remaining for work up of<br />

liver disease (43%). Asians had similar NAFLD activity scores<br />

(NAS) on liver biopsy as non-Asians (2.08 ± 1.13 vs. 2.00 ±<br />

1.02, p=0.54) with similar rates of non-alcoholic steatohepatitis<br />

(NASH, 24% vs. 21%, respectively, p=0.61). However,<br />

the Asian cohort had higher rates of subsequent cirrhosis (30%<br />

vs. 17%, p=0.009), HCC (23% vs. 5%, p < 0.0001), and<br />

orthotopic liver transplantation (13% vs. 4%, p=0.002). On<br />

multivariate analysis, Asian ethnicity was a significant predictor<br />

for HCC (OR=3.20, CI 1.31-7.82, p=0.01) but not for<br />

developing cirrhosis (OR=1.56, CI 0.84-2.91, p=0.16), after<br />

adjusting for age, gender, daily alcohol use, NAS, the presence<br />

of NASH, cirrhosis, and other liver diseases. The presence<br />

of another liver disease was also a significant predictor<br />

for cirrhosis (OR 3.03, CI 1.83-5.01, p < 0.0001) and HCC<br />

(OR 13.02, CI 5.49-30.94, p < 0.0001). CONCLUSION: The<br />

majority of patients with NAFLD in this study underwent incidental<br />

liver biopsy during surgery. Asian patients with NAFLD<br />

have significantly higher risk for HCC compared to other ethnicities,<br />

though they presented with similar NAS and presence<br />

of NASH, cirrhosis, and secondary liver disease.<br />

Disclosures:<br />

Mindie H. Nguyen - Advisory Committees or Review Panels: Bristol-Myers<br />

Squibb, Bayer AG, Gilead, Novartis, Onyx; Consulting: Gilead Sciences, Inc.;<br />

Grant/Research Support: Gilead Sciences, Inc., Bristol-Myers Squibb, Novartis<br />

Pharmaceuticals, Roche Pharma AG, Idenix, Hologic, ISIS<br />

The following authors have nothing to disclose: Benjamin Yip, Brittany E. Yee,<br />

Ailinh Do, Nghia H. Nguyen, Joseph K. Hoang<br />

2251<br />

Normal body mass index as a risk factor for osteopenia<br />

in patients with non-alcoholic fatty liver disease patients<br />

Maria Farrugia 2 , Martina Gerada 2 , John Bonello 2 , James Mario<br />

Gauci 2 , Richard Pullicino 1 , Jurgen Gerada 2 ; 1 Radiology, Mater<br />

Dei Hospital, Msida, Malta; 2 Gastroenterology, Mater Dei Hospital,<br />

Msida, Malta<br />

Background/Aim: The effect of non-alcoholic fatty liver disease<br />

(NAFLD) on bone mineral density (BMD) is poorly understood.<br />

NAFLD is more prevalent with higher body mass indexes (BMI).<br />

We aimed to evaluate the effect of different classes of BMI on<br />

BMD in adult NAFLD patients. Methods: Adult patients diagnosed<br />

with NAFLD on liver imaging in 2013 were enrolled<br />

in the study. Patients with concomitant liver pathologies were<br />

excluded. Patient demographics were obtained from the medical<br />

notes and the BMI calculated and classified using WHO<br />

classification (normal (18.5-24.9), overweight (25-29.9) and<br />

obese (≥30)). BMDs of the femur and lumbar spine were measured<br />

in all patients by dual energy X-ray absorptiometry.<br />

Age and gender-matched Z scores and T scores were calculated<br />

and analysed against the different BMI classes using the<br />

ANOVA model. Results: 197 NAFLD patients were enrolled in<br />

the study (178 females, 90.3%; mean age 63.5, range 35-82;<br />

mean BMI 32.8, range 22-48.6). Obese patients had higher<br />

BMD T scores of the whole femur than overweight patients and<br />

normal BMI patients (p 0.001)(Table 1). This was also true<br />

for the femoral neck (p 0.007) and lumbar spine (p 0.017).<br />

Likewise, obese patients had higher BMD Z scores of the whole<br />

femur (p 0.002), femoral neck (p 0.005) and lumbar spine<br />

(p 0.008) compared to overweight patients and normal BMI<br />

patients. On subgroup analysis of obese patients, obese class<br />

3 patients (BMI >40) had higher BMD T scores of the whole<br />

femur (p 0.002) and lumbar spine (p 0.003) but not of the<br />

femoral neck (p 0.321) when compared to obese class 2 (BMI<br />

35-39.9) and obese class 1 (BMI 30-34.9) patients. Conclusions:<br />

In NAFLD patients, obesity was associated with stronger<br />

bone mineralisation, demonstrating a linear relationship<br />

with increasing BMIs. NAFLD patients with normal BMI should<br />

undergo BMD to screen for osteopenia. Weight loss in NAFLD<br />

obese patients might have deleterious effects on bone health.<br />

Table 1. Comparison between BMI and bone mineral density in<br />

NAFLD patients<br />

Disclosures:<br />

The following authors have nothing to disclose: Maria Farrugia, Martina Gerada,<br />

John Bonello, James Mario Gauci, Richard Pullicino, Jurgen Gerada


1304A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

2252<br />

Testosterone Is Associated with Biochemical Markers of<br />

Non Alcoholic Fatty Liver Disease (NAFLD) in Women<br />

With Polycystic Ovarian Syndrome (PCOS)<br />

Monika Sarkar, Chia-Ning Kao, Norah Terrault, Phyllis Tien, Ruth<br />

Greenblatt, Marcelle I. Cedars, Heather Huddleston; University of<br />

California, San Francisco, San Francisco, CA<br />

Background: Non-alcoholic fatty liver disease (NAFLD) affects<br />

at least half of all women with PCOS. This may relate to the<br />

high prevalence of insulin resistance (IR) and obesity in PCOS.<br />

PCOS is also characterized by hyperandrogenism although<br />

it is unclear whether this resting high androgen state further<br />

promotes worsening NAFLD in PCOS. Methods: This is a cross<br />

sectional study of women without heavy alcohol use and with<br />

confirmed PCOS (n=149) based on established Rotterdam Criteria,<br />

evaluated in a multidisciplinary PCOS clinic between<br />

2006-2014. We aimed to determine if total testosterone is<br />

associated with elevated ALT as a marker of NAFLD, after<br />

adjusting for age, body mass index (BMI) and Homeostasis<br />

Model Assessment of Insulin Resistance (HOMA-IR). Linear<br />

regression was used to evaluate the association between total<br />

testosterone and ALT levels, with point estimates reflecting beta<br />

coefficients. Results: The mean age was 28.4 (+/-6.6), median<br />

BMI was 28.5 (IQR 24.1-36.0), and median HOMA-IR was<br />

1.9, (IQR 1.0-4.7). Median ALT was 21 (IQR 14-32) and half<br />

of women had ALT > 1.5 times upper limits (>29 U/L). Median<br />

total testosterone was 54ng/dL (IQR 40-66). On unadjusted<br />

analysis, total testosterone was positively associated with ALT<br />

(beta coefficient 0.1047, 95% CI 0.0023-0.207, p=0.045)<br />

with stronger association after adjusting for age, BMI, and<br />

HOMA-IR (beta coefficient 0.1152, 95% CI 0.016-0.21,<br />

p=0.02). Increasing BMI (beta coefficient 0.576, 95% CI 0.18-<br />

0.005, p=0.005), but not HOMA-IR (beta coefficient 0.0227,<br />

95% CI -0.122-0.575, p=0.2) was also independently associated<br />

with elevated ALT. Conclusion: Elevated ALT levels, a biochemical<br />

marker of NAFLD, is common in women with PCOS.<br />

Total testosterone was associated with increasing ALT, independent<br />

of BMI and insulin resistance. These findings support an<br />

important role of androgens as a potential mediator of risk for<br />

NAFLD in PCOS, which may offer potential targets for NAFLD<br />

treatment in this high-risk population.<br />

Disclosures:<br />

Norah Terrault - Advisory Committees or Review Panels: Eisai, Biotest; Consulting:<br />

BMS, Merck, Achillion; Grant/Research Support: Eisai, Biotest, Vertex,<br />

Gilead, AbbVie, Novartis, Merck<br />

The following authors have nothing to disclose: Monika Sarkar, Chia-Ning Kao,<br />

Phyllis Tien, Ruth Greenblatt, Marcelle I. Cedars, Heather Huddleston<br />

2253<br />

Prevalence of hepatic steatosis in healthy health care<br />

workers as quantified by controlled attenuation parameter<br />

(CAP)<br />

Anita Arslanow, Frank Lammert, Caroline S. Stokes; Department<br />

of Medicine II, Saarland University Medical Center, Homburg,<br />

Germany<br />

Background: Hepatic fat accumulation is the hallmark of non-alcoholic<br />

fatty liver (NAFL), for which prevalence estimates range<br />

widely from 5 to 50% in the population due to differences<br />

of screening and detection strategies. Since up to 25% of<br />

patients with NAFLD might develop advanced liver fibrosis<br />

and cirrhosis, which increase the risk of liver-associated mortality,<br />

better NAFL screening strategies are needed. Our aim<br />

was to assess the frequency of hepatic steatosis in participants<br />

without a history of chronic liver disease who were recruited<br />

from the Occupational Health Department at Saarland University<br />

Medical Center. Methods: For this prospective study, 81<br />

participants volunteered (80% women, median age 29 years,<br />

range 21 – 63 years). Hepatic steatosis was assessed using<br />

controlled attenuation parameter (CAP), which quantifies the<br />

degree of ultrasound attenuation based on vibration- controlled<br />

transient elastography (VCTE; Fibroscan). Body composition<br />

was determined using an eight electrode bioelectrical impedance<br />

analyzer (BIA). Liver function tests (LFT) and serum lipid<br />

concentrations were measured with standard clinical assays.<br />

Results: The median CAP score was 226 dB/m (range 100<br />

– 400 dB/m). In total, 33 participants (41%) presented with<br />

hepatic steatosis, when using 238 dB/m as cut-off for liver fat<br />

accumulation (Sasso Ultrasound Med Biol 2010). The median<br />

body mass index was significantly (P < 0.001) higher in participants<br />

with than without liver steatosis (26 vs. 22 kg/m 2 ,<br />

respectively) and accordingly, the majority of NAFL patients<br />

were overweight (42%) or obese (15%). In contrast, only 15%<br />

of participants without hepatic steatosis were overweight, and<br />

none was obese (P < 0.0001). In line with these associations,<br />

body fat mass as determined by BIA was elevated in 48% of<br />

participants with hepatic steatosis, as compared to only 6% in<br />

those with low liver fat contents. In contrast, LFT and serum lipid<br />

parameters did not differ significantly between the two groups.<br />

Conclusions: Presence of hepatic steatosis was documented<br />

in almost half of this cohort of perceived healthy individuals.<br />

Whereas serum surrogate markers did not differentiate the individuals,<br />

body composition reflected the differences in liver fat<br />

contents among the two groups. CAP (and BIA) represent rapid<br />

non-invasive methods that facilitate the identification of fatty<br />

liver. Broad NAFL screening in the workplace might be encouraged,<br />

since it may allow timely detection of at risk individuals<br />

and early implementation of lifestyle changes.<br />

Disclosures:<br />

The following authors have nothing to disclose: Anita Arslanow, Frank Lammert,<br />

Caroline S. Stokes<br />

2254<br />

Assessment of parameters related to choline metabolism<br />

in non-alcoholic fatty liver disease: a comparison with<br />

healthy controls<br />

Hannah Da Silva 1,2 , Bianca M. Arendt 1 , Anastasia Teterina 1 , Sandra<br />

E. Fischer 3 , Scott Fung 4 , Johane P. Allard 4 ; 1 Gastroenterology,<br />

Toronto General Hospital, Toronto, ON, Canada; 2 Department of<br />

Nutritional Sciences, University of Toronto, Toronto, ON, Canada;<br />

3 Pathology, Toronto General Hospital, Toronto, ON, Canada;<br />

4 Medicine, Toronto General Hospital, Toronto, ON, Canada<br />

Background/Objectives: Choline deficient diets promote steatosis<br />

similar to non-alcoholic fatty liver disease (NAFLD). Thus,<br />

one of the proposed mechanisms for NAFLD pathogenesis is<br />

reduced choline bioavailability resulting from the metabolism<br />

of dietary choline to trimethylamine (TMA) by gut micriobiota,<br />

causing subsequent reduction of phosphatidylcholine in the<br />

liver and decrease in VLDL export of triglycerides. The validity<br />

of this proposed mechanism was tested in animal models,<br />

but data on choline metabolism in general NAFLD population<br />

and its relation to intestinal microbiota is lacking. Our objective<br />

was to assess parameters related to choline metabolism<br />

and determine whether these parameters are associated with<br />

NAFLD. Methods: We studied serum choline and fecal choline<br />

and TMA in 39 biopsy-confirmed patients (15 simple steatosis<br />

(SS), 24 non-alcoholic steatohepatitis (NASH)), and 28 living<br />

liver donors as healthy controls (HC) by nuclear magnetic resonance<br />

spectroscopy. Clinical, biochemical, and anthropometric<br />

data were also collected. Dietary intake was assessed using<br />

a 7-day food record in a subset of participants (10 SS, 20


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1305A<br />

NASH, 25 HC). The data are presented as median (Q1, Q3)<br />

for non-normal and mean (SD) for normal continuous variables.<br />

Groups were compared using Wilcoxon rank-sum tests and<br />

t-tests. Spearman correlation coefficients were used to examine<br />

association between variables. Results: NAFLD patients were<br />

older than HC and had significantly higher BMI, liver enzymes<br />

(AST, ALT), insulin resistance (HOMA-IR), hemoglobin A1c,<br />

and blood triglycerides (p


1306A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

(5%) pt had uncontrolled diarrhea requiring hospitalization.<br />

Other minor reported AE’s were fatigue (20%), anemia (15%),<br />

rash/itching (10%) and nausea (5%). Conclusion In summary,<br />

HCV in ESRD is considered a difficult-to-treat group and in<br />

whom SVR is associated with overall clinical improvement and<br />

outcomes after kidney transplantation. Half-dose SOF based<br />

regimens provides an attractive option. The regimen appears<br />

to be a safe, well-tolerated and efficacious resulting in higher<br />

rates of sustained virologic response. Additional SVR 12 will<br />

be available for presentation.<br />

Demographics<br />

Grade 2 event (body aches in a patient receiving 100 mg EDP-<br />

239). There were no serious adverse events. The mean terminal<br />

half-life for EDP-239 was approximately 24 hours; inter-subject<br />

variability as measured by CV% was


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1307A<br />

treatment of chronic Hepatitis C Virus infection. TT-034 uses the<br />

process of DNA directed RNAi interference (ddRNAi) which<br />

triggers the cell’s own transcriptional machinery to continuously<br />

produce steady state levels three independent short hairpin<br />

RNA (shRNA) to target 3 independent regions of the HCV<br />

genome. Designed to be administered as a single treatment<br />

delivered intravenously, TT-034 uses an Adeno-Associated<br />

Virus type 8 (AAV8) capsid to deliver a recombinant genome<br />

preferentially into hepatocytes. Biodistribution analyses presented<br />

here demonstrate that 90% of the vector distributes<br />

into liver tissues, with close to 100% transduction of the liver<br />

hepatocytes, as assessed by in situ hybridization. Furthermore,<br />

the durability to expression following a single injection, as<br />

assessed by qPCR, demonstrated that shRNA expression persisted<br />

for the duration of the 180 day experiment. Because<br />

previous reports have suggested that high expression of shRNA<br />

may cause global dysregulation of endogenous miRNA processes<br />

within cells, the impact of long term expression of TT-034<br />

on endogenous miRNA levels was interrogated. Analyses were<br />

performed using RNA isolated from liver biopsies at Day 15,<br />

and from liver and heart tissues collected 60 or 180 days<br />

post TT-034 administration. This miRNA profiling demonstrated<br />

reliable detection of 260 microRNAs in the primate heart samples,<br />

as compared to 266 in liver and 269 in the liver biopsy<br />

samples. The analyses of heart tissues demonstrated that there<br />

was no statistical difference across the groups treated with<br />

TT-034 (ANOVA Benjamini Hochberg (BH) pvalue < 0.05).<br />

Although the liver biopsy samples showed significant effects in<br />

27 microRNA, analyses of the day 60 and day 180 liver samples<br />

showed no statistical differences from the control group.<br />

Collectively, these data suggest that TT-034 is not likely to have<br />

any adverse effects in miRNA processes in primate cells.<br />

Disclosures:<br />

David Suhy - Advisory Committees or Review Panels: Regen BioPharma; Management<br />

Position: Benitec Biopharma<br />

Tin Mao - Employment: Benitec Biopharma<br />

Shih-Chu Kao - Employment: Benitec Biopharma<br />

Per Lindell - Consulting: Benitec Ltd, Tacere Inc<br />

2259<br />

Pharmacokinetics, Safety, and Tolerability of EDP-239,<br />

a Novel Hepatitis C NS5A inhibitor, in Healthy Volunteers<br />

Lijuan Jiang 2 , Xuemin Jiang 1 , Kiran Dole 1 , Kimberley Caliri 2 , Kenneth<br />

Tack 2 , Richard Colvin 1 , Yat Sun Or 2 ; 1 NIBR, Cambridge, MA;<br />

2 Enanta Pharmaceuticals, Watertown, MA<br />

Background: EDP-239, a novel NS5A inhibitor, demonstrates<br />

potent anti-HCV activity in vitro, with mean EC 50<br />

values of 31<br />

and 10 pM against genotypes 1a and 1b, respectively. Methods:<br />

Pharmacokinetics, safety, and tolerability of EDP239 were<br />

assessed in a randomized, double-blind, placebo-controlled<br />

study. 94 healthy volunteers received EDP239/placebo either<br />

as single oral doses from 3 mg to 300 mg or 7 days of QD<br />

dosing of 10 mg to 200 mg. Results: EDP239 plasma exposure<br />

(C max<br />

and AUC) increased with dose in a slightly more<br />

than dose-proportional manner following single doses or 7-day<br />

dosing. The half-life was approximately 24 hours. Steady-state<br />

was achieved by Day 5 with a 2-fold increase in exposure. On<br />

Day 7, inter-subject variation for AUC 0-24<br />

was less than 22%<br />

from 10 mg through 200 mg; mean AUC 0-24<br />

was 73,500<br />

ng*hr/mL at the 200 mg dose. EDP239 was well tolerated.<br />

Rates of adverse events (AEs) were similar in EDP239 and<br />

placebo-treated subjects. There were no serious adverse events<br />

and no discontinuations due to AEs. No clinically significant<br />

laboratory abnormalities were observed. Conclusions: EDP239<br />

was well absorbed, achieving therapeutic levels at achievable<br />

doses. No safety or tolerability issues were observed. Further<br />

development of EDP239 as part of a combination regimen for<br />

HCV infection is warranted.<br />

EDP239 plasma PK parameters following QD oral dosing x 7d in<br />

healthy volunteers<br />

Disclosures:<br />

Lijuan Jiang - Employment: Enanta Pharmaceuticals, Inc<br />

Xuemin Jiang - Employment: Novartis<br />

Kenneth Tack - Consulting: Enanta Pharmaceuticals, Inc<br />

Richard Colvin - Employment: Novartis Institute for Biomedical Research<br />

Yat Sun Or - Employment: Enanta Pharmaceuticals, Inc, Enanta Pharmaceuticals,<br />

Inc<br />

The following authors have nothing to disclose: Kiran Dole, Kimberley Caliri<br />

2260<br />

Pharmacokinetics, Safety and Tolerability of EDP-239, a<br />

Novel Hepatitis C (HCV) NS5A inhibitor, in Patients with<br />

HCV Infection<br />

Lijuan Jiang 1 , Xuemin Jiang 2 , Kiran Dole 2 , Eric Lawitz 3 , Fred Poordad<br />

3 , Stefan Zeuzem 4 , Kimberley Caliri 1 , Kenneth Tack 1 , Richard<br />

Colvin 2 , Yat Sun Or 1 ; 1 Enanta Pharmaceuticals, Watertown,<br />

MA; 2 Novartis Institutes for Biomedial Research, Cambridge, MA;<br />

3 Texas Liver Institute, San Antonio, TX; 4 Johan Wolfgang Goethe<br />

Universitat, Frankfurt, Germany<br />

Background: EDP-239, a novel NS5A inhibitor, demonstrates<br />

potent anti-HCV activity in vitro. Studies in normal volunteers<br />

have shown good absorption, safety, and tolerability at doses<br />

up to 200 mg daily x 7 days. Methods: Pharmacokinetics,<br />

safety, and tolerability of EDP-239 were assessed in a randomized,<br />

double-blind, placebo-controlled study. 28 patients with<br />

chronic HCV (GT 1a/b) infection received EDP-239 or placebo<br />

as single oral doses from 10 mg to 200 mg. Results: EDP-<br />

239 was rapidly absorbed following a single oral dose with<br />

T max<br />

of 2 – 3 hrs. Plasma exposure (C max<br />

and AUC) increased<br />

with dose in a slightly more than dose-proportional manner<br />

following single doses from 10 mg to 200 mg. The half-life<br />

was approximately 24 hours. EDP-239 was well tolerated.<br />

Rates of adverse events (AEs) were similar in EDP239- and<br />

placebo-treated patients. Almost all AEs were mild in intensity.<br />

There were no serious adverse events. No clinically significant<br />

laboratory abnormalities were observed. Conclusions: EDP-<br />

239 was well absorbed, achieving supratherapeutic levels in<br />

HCV-infected patients. Mean plasma concentration 24 hrs after<br />

a single oral dose of 100 mg was > 10,000 X the in vitro<br />

replicon EC 50<br />

values of genotype 1 virus. EDP-239 was well<br />

tolerated and no safety issues were observed. Further development<br />

of EDP-239 as part of a combination regimen for HCV<br />

infection is warranted.


1308A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

EDP-239 plasma PK parameters following oral dosing in HCV<br />

patients<br />

Disclosures:<br />

Lijuan Jiang - Employment: Enanta Pharmaceuticals, Inc<br />

Xuemin Jiang - Employment: Novartis<br />

Eric Lawitz - Advisory Committees or Review Panels: AbbVie, Achillion Pharmaceuticals,<br />

Regulus, Theravance, Enanta, Idenix Pharmaceuticals, Janssen, Merck<br />

& Co, Novartis, Gilead; Grant/Research Support: AbbVie, Achillion Pharmaceuticals,<br />

Boehringer Ingelheim, Bristol-Myers Squibb, Gilead Sciences, GlaxoSmith-<br />

Kline, Idenix Pharmaceuticals, Intercept Pharmaceuticals, Janssen, Merck & Co,<br />

Novartis, Nitto Denko, Theravance, Salix, Enanta; Speaking and Teaching: Gilead,<br />

Janssen, AbbVie, Bristol Meyers Squibb<br />

Fred Poordad - Advisory Committees or Review Panels: Abbott/Abbvie, Achillion,<br />

BMS, Inhibitex, Boeheringer Ingelheim, Pfizer, Genentech, Idenix, Gilead,<br />

Merck, Vertex, Salix, Janssen, Novartis; Grant/Research Support: Abbvie,<br />

Anadys, Achillion, BMS, Boehringer Ingelheim, Genentech, Idenix, Gilead,<br />

Merck, Pharmassett, Vertex, Salix, Tibotec/Janssen, Novartis<br />

Stefan Zeuzem - Consulting: Abbvie, Bristol-Myers Squibb Co., Gilead, Merck<br />

& Co., Janssen<br />

Kenneth Tack - Consulting: Enanta Pharmaceuticals, Inc<br />

Richard Colvin - Employment: Novartis Institute for Biomedical Research<br />

Yat Sun Or - Employment: Enanta Pharmaceuticals, Inc, Enanta Pharmaceuticals,<br />

Inc<br />

The following authors have nothing to disclose: Kiran Dole, Kimberley Caliri<br />

2261<br />

Short-term safety and pharmacokinetics of co-administration<br />

of EDP239, an HCV NS5A inhibitor, and alisporivir<br />

(ALV), a cyclophilin inhibitor, in healthy subjects<br />

Xuemin Jiang 1,2 , Richard Colvin 1,2 , June Ke 1,2 , Sachin Desai 1,2 ,<br />

Surendra Machineni 1,2 , Jie Zhang 1,2 , Wei Zhou 1,2 , Sun Haiying<br />

1,2 , Steven J. Kovacs 1,2 ; 1 Novartis, East Hanover, NJ; 2 Institutes<br />

for Biomedical Research, East Hanover, NJ<br />

Introduction: Cyclophilin A plays essential roles in HCV replication<br />

and assembly, including modulating the function of NS5A.<br />

Combined cyclophilin and NS5A inhibition is a clinically<br />

untested therapeutic approach despite in vitro <strong>studies</strong> suggesting<br />

additive-to-synergistic effects of the two mechanisms, acting<br />

at functionally dependent sites of the HCV replication complex.<br />

An EDP239 C 24h<br />

of 317 ng/mL reduced viral load by >3<br />

log10 after a single 100 mg dose in GT1-infected patients. ALV<br />

is a cyclophilin inhibitor that has pan-genotypic anti-HCV activity<br />

with demonstrated clinical efficacy. It is a time-dependent<br />

CYP3A4 inhibitor and inhibitor of multiple uptake and efflux<br />

transporters, including P-gp. EDP239 is a substrate of CYP3A4<br />

and P-gp in vitro. Methods: Open-label; fixed sequence design<br />

(n=26). Subjects received multiple doses of EDP239 30 mg BID<br />

for 21 days with co-administration of ALV 400 mg BID days 8<br />

through 21. Both drugs administered with food. Blood samples<br />

were collected up to 12 hours after dose administration on<br />

days 7, 15 and 21 to measure plasma concentrations by LC/<br />

MS/MS. Non-compartmental PK analysis using WinNonlin;<br />

ANOVA for statistical comparisons; routine safety assessments<br />

with safety data descriptively summarized. Results: 22 subjects<br />

completed the study. Study treatments were well tolerated<br />

with no safety-related discontinuations. Adverse effects were<br />

consistent with the known effects of either EDP239 (gastrointestinal<br />

complaints) or ALV (gastrointestinal complaints, headache,<br />

laboratory abnormalities, increases in blood pressure).<br />

Median EDP239 T max<br />

was 3, 4.5, and 4 h on days 7, 14,<br />

and 21, respectively. The EDP239 C 12h<br />

was 2180 ng/mL<br />

when administered with ALV. EDP 100 mg was comparable<br />

to C12h of EDP 30 mg when combined with a alisporivir.<br />

EDP239 had no apparent effect on ALV exposure (compared<br />

with historical data). Conclusion Short term co-administration<br />

of EDP239 and ALV was safe and well tolerated. The data<br />

suggest that therapeutic combination of ALV and EDP239 is<br />

feasible. The observed EDP239 exposures can be expected<br />

to have clinical activity against GT1 and GT3 HCV. The time<br />

course of exposure to EDP239 was relatively flat with minimal<br />

difference between peak and trough. Exploiting the interaction<br />

may allow EDP239 to treat patients with G3 HCV infection<br />

without increasing overall exposure beyond exposures previously<br />

safely investigated in humans.<br />

Disclosures:<br />

Xuemin Jiang - Employment: Novartis<br />

Richard Colvin - Employment: Novartis Institute for Biomedical Research<br />

Surendra Machineni - Employment: Novartis Healthcare Private Limited<br />

Steven J. Kovacs - Employment: Novartis<br />

The following authors have nothing to disclose: June Ke, Sachin Desai, Jie Zhang,<br />

Wei Zhou, Sun Haiying<br />

2262<br />

SB 9200 Shows Antiviral Activity Against HCV-RNA<br />

by Targeting Host Cytosolic Sensor Proteins RIG-I and<br />

NOD-2 to activate IFN signaling pathways<br />

Kumar Visvanathan 2,1 , Rosemary M. Millen 2,1 , Stuart K. Roberts 4,5 ,<br />

Peter W. Angus 6,2 , Wendy Cheng 8 , Nada Farhat 9 , My My Trinh 9 ,<br />

Radhakrishnan P. Iyer 10 , Murray L. Barclay 7 , Alex J. Thompson 3,2 ;<br />

1 Innate Immunity Laboratory, St Vincent’s Hospital Melbourne, Melbourne,<br />

VIC, Australia; 2 Medicine, Univeristy of Melbourne, Melbourne,<br />

VIC, Australia; 3 Gastroenterology, St. Vincents Hospital,<br />

Melbourne, Fitzroy, VIC, Australia; 4 Gastroenterology, Alfred Hospital,<br />

Melbourne, VIC, Australia; 5 Medicine, Monash University,<br />

Melbourne, VIC, Australia; 6 Liver Transplant Unit, Austin Hospital,<br />

Melbourne, VIC, Australia; 7 Gastroenterology & Clinical Pharmacology,<br />

Christchurch Hospital and University of Otago, Christchurch,<br />

New Zealand; 8 Gastroenterology, Royal Perth Hospital,<br />

Perth, WA, Australia; 9 Pharsight Consulting Services, Princeton,<br />

NJ; 10 Spring Bank Pharmaceuticals, Milford, MA<br />

Introduction. SB9200 activates RIG-I and NOD-2 pathways in<br />

virally infected cells and has been shown to maximally reduce<br />

HCV RNA by 1.9 log10 in HCV patients (Thompson EASL<br />

2015). We aimed to evaluate the relationship between the<br />

pharmacokinetics of SB 9200, and the induction of innate<br />

immunity and antiviral response. Methods. HCV patients, GT 1<br />

and 3, received 200-900 mg of SB9200 PO daily for 7 days.<br />

Early antiviral kinetics were measured from baseline through<br />

7 days after the first dose of SB 9200. Patients were divided<br />

into null responders at day 7 (NR, < 0.5 log10 reduction in<br />

HCV RNA, n=11) or responders (R, >0.9 log10, n=6). Clinical<br />

characteristics at baseline for NR vs. R were similar for<br />

median HCV RNA, BMI, ALT, AST, platelets and Fibroscan.<br />

50% of R were IL28b CC compared to 18% of NR. Plasma<br />

interferon alpha (IFNa) was measured on day 1, 7 and 14<br />

by ELISA. Interferon stimulating genes ISG15 and OAS1 were<br />

measured on Day 1, 2, 3, 7 and 14 in peripheral blood via<br />

RT-PCR. GAPDH was used as an internal control. Data are represented<br />

as ratios from baseline for IFNa, ISG15, and OAS1.<br />

For pharmacokinetic analysis, blood samples were collected on


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1309A<br />

Day 7 from pre-dose through 48 hrs post-dose. Results. Plasma<br />

IFNa was significantly increased from baseline in R vs.NR<br />

(Fig 1; P


1310A AASLD ABSTRACTS HEPATOLOGY, October, 2015<br />

HCV replication and the impact on the response to treatment of<br />

HCV Materials and Methods: The replicon cells Huh7/Ava.5<br />

(genotype 1b), Huh7/J6/JFH (genotype 2a) and Huh7.5/<br />

Con1 (genotype 1b) were obtained and the mirVana let-7g<br />

mimic/inhibitor, miR-122 inhibitor and miRNA mimic/inhibitor<br />

negative control were purchased. The 1.0-kb and 0.5-<br />

kb fragments of the let-7g promoter were amplified by PCR.<br />

Site-directed mutagenesis of the AP-1 binding-site presented in<br />

the let-7g promoter region was carried out. WST-1 assay and<br />

Renilla Luciferase Assay were used. Expression levels of let-7g<br />

in each sample were normalized to the corresponding level of<br />

snU6B. Expression levels of let-7g were determined by using<br />

the Quantitative real-time PCR. Anti-phospho-ERK, p38, JNK,<br />

and total-ERK, p38, JNK antibodies, Anti-GAPDH and α-tubulin<br />

antibodies were used for Immunoblot analysis. Results:<br />

Our results demonstrated that overexpression of let-7g reduces<br />

HCV expression in the cell line. The HCV loads were more<br />

decreased by let-7g mimic than miR-122 inhibitor transfected<br />

cell. High levels of lin28 correlate with low levels of let-7g<br />

in HCV-infected cells and the knockdown of lin28 via siRNA<br />

reduces HCV replication. The treatment with a let-7g mimic<br />

alone was shown to induce IFN-induced genes inclusion MxA<br />

and OAS1. Interferon (IFN)/RBV induces let-7g expression,<br />

and let-7g and IFN/ribavirin also elicited an addictive inhibitory<br />

effect on HCV expression. The anti-viral effects of let-7g<br />

mediated IFN/RBV signaling and the regulation of let-7g by<br />

IFN/RBV occurs through p38/AP1 signaling. Conclusions: We<br />

have indicated an important role of let-7g on the replication of<br />

HCV and on the response of HCV to anti-HCV treatment. The<br />

IFN/ribavirin induces let-7g expression through p38/AP-1 signaling<br />

and the let-7g and IFN/RBV have additively inhibitory<br />

effect on HCV replication. The let-7g may serve as a potential<br />

target for HCV therapy.<br />

Disclosures:<br />

Wan-Long Chuang - Advisory Committees or Review Panels: Gilead, Abbvie;<br />

Speaking and Teaching: BMS, Roche, MSD<br />

Ming-Lung Yu - Advisory Committees or Review Panels: ABBOTT, MSD, ABBVIE,<br />

GILEAD, J&J, ROCHE, BMS; Grant/Research Support: ABBOTT, ROCHE, MSD,<br />

ABBVIE, GILEAD; Speaking and Teaching: ABBOTT, ROCHE, MSD, GILEAD,<br />

BMS, GSK<br />

The following authors have nothing to disclose: Chia-Yen Dai, Wen-Wen Chou,<br />

Yi-Shan Tsai, Shu-Chi Wang, Chung-Feng Huang, Ming-Lun Yeh, Jee-Fu Huang<br />

2265<br />

Drug-Drug Interaction Studies between HCV Antivirals<br />

Sofosbuvir and GS-5816 and HIV Antiretrovirals<br />

Erik Mogalian 1 , Luisa M. Stamm 2 , Anu Osinusi 2 , Gong Shen 4 ,<br />

Karim Sajwani 3 , John McNally 2 , Anita Mathias 1 ; 1 Clinical Pharmacology,<br />

Gilead Sciences, Inc., Foster City, CA; 2 Clinical Research,<br />

Gilead Sciences, Inc., Foster City, CA; 3 Clinical Operations, Gilead<br />

Sciences, Inc., Foster City, CA; 4 Biostatistics, Gilead Sciences,<br />

Inc., Foster City, CA<br />

Introduction A once-daily fixed-dose combination tablet composed<br />

of sofosbuvir (SOF; nucleotide analog NS5B inhibitor)<br />

and GS-5816 (pangenotypic NS5A inhibitor) is in clinical<br />

development for the treatment of chronic HCV infection. Phase<br />

1 <strong>studies</strong> were conducted in healthy volunteers to evaluate<br />

potential drug-drug interactions (DDIs) between SOF/GS-5816<br />

and HIV antiretroviral (ARV) regimens without a pharmacokinetic<br />

“booster” (ritonavir or cobicistat) to support their use<br />

together in HIV/HCV co-infected patients. Methods These were<br />

multiple-dose, randomized, cross-over DDI <strong>studies</strong>. Subjects<br />

received SOF/GS-5816 and ARVs (EFV/FTC/TDF [Atripla ® ;<br />

ATR], FTC/RPV/TDF [Complera ® ; CPA], DTG [Tivicay ® ], and<br />

RAL [Isentress®] + FTC/TDF [Truvada®]) alone and in combination.<br />

Steady-state plasma concentrations of SOF, its predominant<br />

circulating nucleoside metabolite GS-331007, GS-5816,<br />

and ARVs were analyzed on the last day of dosing for each<br />

treatment and PK parameters were calculated. Geometric leastsquares<br />

means ratios and 90% confidence intervals (combination<br />

vs. alone) for SOF, GS-331007, GS-5816, and ARV<br />

AUC tau<br />

, C max<br />

and C tau<br />

were estimated and compared against<br />

pre-specified lack of PK alteration boundaries of 70-143% for<br />

all analytes except RAL (50-200%). Safety assessments were<br />

conducted throughout the study. Results Twenty-nine of 30 subjects<br />

in Group 1 and all subjects in Groups 2-4 (N=78 total)<br />

completed the study. The majority of adverse events (AEs) were<br />

Grade 1 and there were no serious AEs. The most frequent AEs<br />

were headache (12%) and constipation (10%). One subject in<br />

Group 1 discontinued due to an AE (Grade 1 urticaria). SOF/<br />

GS-5816 PK was unaffected by RPV, DTG, and RAL regimens.<br />

A decrease (~53%) in GS-5816 exposure with no impact on<br />

overall SOF or GS-331007 exposure was observed following<br />

administration with ATR. No clinically significant changes<br />

in the PK of DTG, EFV, FTC, RAL, and RPV were observed<br />

when administered with SOF/GS-5816. Increased TFV exposure<br />

(ATR: ~1.8-2.2-fold; CPA: ~1.4-1.8-fold; FTC/TDF+RAL:<br />

~1.4-1.7-fold) was observed with SOF/GS-5816; overall,<br />

absolute TFV AUC in the test (combination) treatments were<br />

similar to those achieved when FTC/TDF is administered with<br />

ritonavir-boosted PIs, which do not warrant dose adjustment.<br />

Conclusion Study treatments were generally well tolerated.<br />

Results from this study demonstrate that SOF/GS-5816 may be<br />

administered safely with CPA, RAL and DTG with a backbone<br />

of FTC/TDF. Additional data from SOF/GS-5816 Phase 3 PK/<br />

PD will guide the recommendation for use with ATR.<br />

Disclosures:<br />

Erik Mogalian - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Luisa M. Stamm - Employment: Gilead Sciences<br />

Anu Osinusi - Employment: Gilead Sciences<br />

Karim Sajwani - Employment: Gilead Sciences, Inc.<br />

John McNally - Employment: Gilead Sciences, Inc; Stock Shareholder: Gilead<br />

Sciences, Inc<br />

Anita Mathias - Employment: Gilead Sciences Inc.,<br />

The following authors have nothing to disclose: Gong Shen<br />

2266<br />

Discovery of AT-337 and AT-339, Two Highly Potent<br />

and Selective Nucleotide Prodrug Inhibitors of HCV<br />

Polymerase<br />

Steven S. Good 1 , Adel Moussa 1 , Jean-Christophe Meillon 2 , Jean-<br />

Pierre Sommadossi 1 ; 1 Atea Pharmaceuticals Inc., Boston, MA;<br />

2 Oxeltis, Montpellier, France<br />

Background: Nucleotide analogs are preferred candidates for<br />

treatment of HCV infection since they are potent, pan-genotypic<br />

inhibitors of NS5B polymerase with a high barrier to<br />

drug resistance. Here we report novel nucleotide prodrugs with<br />

both base and sugar modifications that are selective for and<br />

highly active against in vitro HCV replication. Methods: Novel<br />

nucleotide prodrugs were synthesized and tested in Huh-7 cells<br />

bearing an HCV genotype 1b replicon and evaluated for activity<br />

against other viruses. Selectivity was assessed in 14-day<br />

exposure with human BFU-E and CFU-GM cells and in 3-day<br />

exposure with human iPS cell-derived cardiomyocytes. Selected<br />

nucleoside-5’-triphosphate analogs were evaluated against the<br />

NS5B polymerase and human DNA polymerases α, β and<br />

γ. Results: The two most potent nucleotide prodrugs, AT-337<br />

and AT-339, inhibited HCV replication, with EC 50<br />

values as<br />

low as 0.005 mM. Intra-assay comparisons demonstrated that<br />

AT-337 and AT-339 were at least 10-fold more potent than


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AASLD ABSTRACTS 1311A<br />

sofosbuvir, as shown in the inhibition curves below. A unique<br />

feature of this series is the formation of at least two active<br />

metabolites from each candidate that compete against different<br />

naturally occurring ribonucleoside triphosphates, with IC 50<br />

values<br />

against the NS5B polymerase between 1.7 and 0.11 mM.<br />

AT-337 and AT-339 demonstrated excellent selectivity, with<br />

CC 50<br />

values greater than 100 mM in Huh-7 cells, human bone<br />

marrow stem cells and human cardiomyocytes. Furthermore,<br />

no inhibition of human DNA polymerase α, β or γ, no activity<br />

against other RNA or DNA viruses, and no toxicity in all host<br />

cell lines was observed at concentrations up to 100 mM. Conclusions:<br />

New nucleotide prodrugs potently inhibited the HCV<br />

NS5B polymerase by producing multiple active analogs of different<br />

ribonucleoside triphosphates. IND enabling <strong>studies</strong> are<br />

underway, targeting clinical evaluation in early 2016.<br />

Disclosures:<br />

Steven S. Good - Employment: Atea Pharmaceuticals Inc.; Stock Shareholder:<br />

Atea Pharmaceuticals Inc.<br />

Jean-Christophe Meillon - Employment: Oxeltis; Independent Contractor: Atea<br />

Pharmaceuticals, Inc.; Stock Shareholder: Oxeltis<br />

The following authors have nothing to disclose: Adel Moussa, Jean-Pierre Sommadossi<br />

2267<br />

The impact of α-fetoprotein level during interferon-free<br />

treatment of hepatitis C virus<br />

Masataka Seike, Koichi Honda, Junya Oribe, Mizuki Endo, Mie<br />

Yoshihara, Masao Iwao, Masanori Tokoro, Kazunari Murakami;<br />

gastroenterology, Oita University, Yufu-city, Japan<br />

AIM: The prevalence of older patients with hepatitis C virus<br />

(HCV) has been increasing in Japan. Elderly patients are at<br />

higher risk for hepatocellular carcinoma (HCC) even after HCV<br />

eradication. The risk of developing HCC is age-dependent, and<br />

patients aged ≥ 75 years occasionally present with HCC. α-Fetoprotein<br />

(AFP) levels after interferon (IFN) therapy for HCV are<br />

significantly associated with hepatocarcinogenesis. However,<br />

the significance of AFP level after IFN-free therapy in older<br />

patients remains unclear. Thus, we analyzed AFP levels during<br />

IFN-free treatment in patients with HCV. Patients and methods:<br />

This study included 63 patients (31 males and 32 females)<br />

with HCV genotype 1b who received 60-mg daclatasvir once<br />

daily plus 100-mg asunaprevir twice daily. The mean age of<br />

these patients was 71.3 years (range, 49–82 years). AFP level<br />

was assessed at baseline and every month. These cases were<br />

followed up by abdominal ultrasonography (US), contrast-enhanced<br />

computed tomography (CT), or ethoxybenzyl contrast<br />

magnetic resonance imaging (MRI) every 3–6 months. Results:<br />

The mean serum AFP level in all patients before treatment was<br />

26.7 ng/ml. AFP level during treatment tended to decrease<br />

immediately and had decreased to 11.7 ng/ml at the end of<br />

treatment. Serum AFP levels in 36 patients decreased to < 5<br />

ng/ml at the end of treatment (group A). The group A patients<br />

had no HCC on abdominal US, CT, or MRI. However, serum<br />

AFP levels in 27 patients decreased to > 5 ng/ml at the end of<br />

treatment or increased during therapy (group B). Twelve group<br />

B patients (44%) were diagnosed with HCC after treatment.<br />

Conclusion: The AFP trend during treatment was useful for predicting<br />

future HCC risk after all-oral therapy with daclatasvir<br />

plus asunaprevir for HCV. Patients with HCV and an AFP level<br />

> 5 ng/ml at the end of treatment should be examined for HCC<br />

by US, CT, and/or MRI as soon as possible.<br />

Disclosures:<br />

The following authors have nothing to disclose: Masataka Seike, Koichi Honda,<br />

Junya Oribe, Mizuki Endo, Mie Yoshihara, Masao Iwao, Masanori Tokoro,<br />

Kazunari Murakami


1312A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Author Index<br />

All numbers refer to the corresponding abstract.<br />

A. Bhanji, Rahima 308<br />

Aa, Jiye 919, 934<br />

Aagaard, Niels Kristian 768, 2071<br />

Aarslev, Kristian 301<br />

Abate, Maria Lorena 1740, 2209<br />

Abbas, Zaigham 1872<br />

Abbasy, Mohammed 733<br />

Abd, Mortadha 943<br />

Abdallah, Ayat 1890<br />

Abdel-Razek, Wael 733<br />

Abdelatif, Dinan 892, 2173, 2192,<br />

2212<br />

Abdelmalek, Manal F. 105, 161, 162,<br />

239, 737, 739, 889, 1428, 1435,<br />

2145, 2153, 2154, 2169, 2220, 2239<br />

Abdelsameea, Eman 1890<br />

Abe, Hiromi 663, 986, 1150, 1627,<br />

1645, 1649, 1670, 1679, 1681,<br />

1858, 1895, 1944, 2120<br />

Abe, Hiroshi 1097<br />

Abe, Masafumi 551<br />

Abe, Masanori 338, 788, 1293, 2146<br />

Abe, Mitsuhiko 688<br />

Abe, Yuta 336<br />

Abecassis, Michael M. 71<br />

Aberg, Judith 380, 457<br />

Abergel, Armand 1036<br />

Abergel, Armando 95, 219, 777, 1009,<br />

1505<br />

Abiru, Seigo 408, 783, 1600<br />

Aboelsoud, Mohammed M. 169, 1625<br />

Abouelhoda, Mohamed 1821<br />

Abouljoud, Marwan S. 556, 1274<br />

Aboulnasr, Fatma 501, 506, 1096<br />

Abraham-Enachescu, Ioana 1940<br />

Abraldes, Juan 742, 745<br />

Abramowsky, Carlos R. 12<br />

Abrams, Marc 2116<br />

Abrams, Michael 2007<br />

Abreu, Rodrigo M. 1544<br />

Abreu de Carvalho, Luis Filipe 487<br />

Abreu-Gonzalez, Pedro 1342<br />

Abrignani, Sergio 2011<br />

Abt, Peter L. 837, 853, 864, 868, 871<br />

Abu Shanab, Ahmed 1275<br />

Abu-Raddad, Laith 1829<br />

Abunimeh, Manal 722, 1039<br />

Abzug, Mark J. 1737<br />

Acalovschi, Monica 463<br />

Achahboun, Mohamed 480<br />

Achiwa, Koichi 967<br />

Acibucu, Fettah 2219<br />

Ackerman, Peter 716, 726, 728<br />

Acosta Umanzor, Carlos 691<br />

Adam, Rene 449, 487, 1237<br />

Adams, David H. 1296, 1298, 1710<br />

Adams, Leon A. 848, 1264, 2099,<br />

2185<br />

Adams, Paul 2117<br />

Adams, William 1172<br />

Adamson, Gord 424<br />

Addissie, Benyam D. 397<br />

adebajo, corlan 539<br />

Adelmeijer, Jelle 740, 2249<br />

Adem, Fatima 1118<br />

Adeyi, Oyedele 11<br />

Adhoute, Xavier 394, 2082<br />

Adinolfi, Luigi 1837<br />

Adler, Joel T. 1732<br />

Adorini, Luciano 142, 882, 893, 902,<br />

907<br />

Aerts, Raymond 171<br />

Afdhal, Nezam H. 96, 249, 737, 739,<br />

746, 1130, 1428, 1435, 1442, 2227<br />

Afdhal, Sophie K. 1177<br />

Afendy, Mariam 534<br />

Afonso, Marta B. 176, 832, 897, 940,<br />

960<br />

Afredj, Nawel 1621<br />

Africa, Jonathan A. 159<br />

Afsharzadeh, Danial 1416<br />

Agarwal, Amol 536<br />

Agarwal, Banwari 1769, 1780, 1781<br />

Agarwal, Kosh 249, 646, 1084, 1098,<br />

1113, 1117, 1132, 1173, 1588,<br />

1704, 1744, 2053, 2054<br />

Agarwal, Parul D. 434<br />

Agarwal, Rajiv 267<br />

Agarwal, Ritu 1095, 1185<br />

Agarwal, Shaleen 853<br />

Aggarwal, Rakesh 1475, 1817<br />

Aggarwal, Richa 1945<br />

Aghemo, Alessio 1743, 1747<br />

Agiasotelli, Danai 1489, 2096<br />

Agopian, Vatche 369<br />

Agostinelli, Laura 21<br />

Agrawal, Saurabh 1903<br />

Agrawal, Swastik 214<br />

Aguiar, Miguel Osman D. 1484, 1494,<br />

1495<br />

Aguilar, Humberto I. 41, 1065<br />

Aguilar, Maria 237, 1252, 2159<br />

Aguilar Schall, Raul E. 624, 632, 737,<br />

739, 746, 777, 1428, 1435, 1442,<br />

2149, 2239<br />

Aguilar-Urbano, Víctor 2171<br />

Aguilera, Victoria 1230, 1251<br />

Ahammed, Sk Mahiuddin 750<br />

Ahlén, Gustaf 992<br />

Ahlquist, David 169<br />

Ahluwalia, Vishwadeep 52, 1472,<br />

1491<br />

Ahmad, Dina 748, 769<br />

Ahmad, Irfan 745<br />

Ahmad, Jawad 1185, 1922, 1931<br />

Ahmad, Maliha 295, 1181<br />

Ahmed, Aijaz 199, 201, 237, 534,<br />

541, 877, 1079, 1182, 1189, 1252,<br />

1534, 1538, 1745, 1750, 2155, 2159<br />

Ahmed, Emilia 2072, 2076<br />

Ahmed, Emily 1242<br />

Ahmed, Kazi 316<br />

Ahmed, Ola 1821<br />

Ahmed, Osman 542<br />

Ahmed, Shahzad 1279<br />

Ahmed, Zohair 774<br />

Ahmed Mohammed, Hager 381, 415<br />

Ahn, Hongkeun 242, 479, 2016<br />

Ahn, Jem Ma 433, 1504, 1591<br />

Ahn, Joseph 1101<br />

Ahn, Joseph C. 7<br />

Ahn, Sang Bong 924, 941, 951, 2218<br />

Ahn, Sang Hoon 241, 346, 504, 1542,<br />

1558, 1581, 2014, 2025<br />

Ahn, Daegeon 423<br />

Ahrens, William A. 653, 1023<br />

Ahuja, Chirag K. 214<br />

Ahya, Vivek N. 536<br />

Aibe, Yuki 760, 1447, 1464, 1492<br />

Aichelburg, Maximilian C. 561, 1204<br />

Aihara, Arihiro 1962<br />

Aihara, Eitaro 676<br />

Aikata, Hiroshi 472, 986, 1056, 1150,<br />

1627, 1649, 1670, 1681, 1858,<br />

1893, 1895, 1944, 2120, 2158<br />

Ain, Dani 1076<br />

Ait Younes, Sonia 1621<br />

Ait-Oufella, Hafid 2148<br />

Aït-Slimane, Tounsia 196<br />

Aithal, Guruprasad P. 225, 908, 1524,<br />

1930, 1933<br />

Aitken, Celia 1803<br />

Aizawa, Nobuhiro 466, 770, 781,<br />

1533, 1669, 1975<br />

Ajmera, Veeral H. 2161<br />

Akamatsu, Sakura 986, 1649, 1681,<br />

1858, 1944<br />

Akarca, Ulus S. 246, 558, 1247, 1739,<br />

1822<br />

Akat, Kemal M. 1364<br />

Akazawa, Yoshihiro 1956<br />

Akazawa, Yuko 952, 1671<br />

Akbani, Rehan 130<br />

Akbar, Sheikh Mohammad Fazle 2157<br />

Akiba, Jun 465, 1990<br />

Akihiro, Isogawa 2240<br />

Akinc, Akin 36<br />

Akita, Maaya 906, 2156<br />

Akiyama, Takumi 296<br />

Akpadock, Evelyn 197<br />

Akremi, Raoudha 206<br />

Akriviadis, Evangelos 1126<br />

Akuta, Norio 1062, 1191, 2020, 2241<br />

Akyuz, Filiz 478<br />

Akyuz, Umit 478<br />

Al Ikhwan, Mahdi 1431<br />

Al Kaabi, Saad R. 1843<br />

Al Mamari, Said 622<br />

Al Mamun, Ayub 2157<br />

Al Masaoudi, Malika 1924<br />

Al Zoairy, Ramona 1092, 1885<br />

Al-Akkad, Walid 337, 1388<br />

Al-ashgar, Hamad I. 1607<br />

Al-Ejji, Khalid 1188<br />

Al-Judaibi, Bandar 1246<br />

Al-khafaji, Ali 275<br />

Al-Khatib, Safa H. 113<br />

Al-mana, Hadeel 1607<br />

Al-Nujaidi, Danya Y. 413<br />

Al-Osaimi, Abdullah M. 765, 766,<br />

1122<br />

Al-Rubaish, Fatima A. 413<br />

Alahdab, Fares 282<br />

Alain, Sophie 1162<br />

Alam, Altaf 1872<br />

Alam, Seema 1697, 1725<br />

Alao, Hawwa 218, 1498


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1313A<br />

Alaparthi, Lakshmi 892, 1428, 2173,<br />

2192<br />

Alarcon-Galvan, Gabriela 1281<br />

Alarcón-Vila, Cristina 899<br />

Alatrakchi, Nadia 25, 1015<br />

Alavi, Maryam 543, 1868<br />

Albeladi, Khalid 1607<br />

Albert, Dustin 285<br />

Alberti, Alfredo 775, 1089, 1802,<br />

1844<br />

Alberts, Steven 127, 1966<br />

Albillos, Agustin 745, 2171<br />

Albrecht, Jeffrey 31<br />

Albuerne, Marisol 2100<br />

Albuquerque, Miguel 480<br />

Alcântara, Camila M. 2072, 2076<br />

Alcaraz-Quiles, Jose 2087<br />

Aldairy, Wassim 1920<br />

Aldea, A. 596<br />

Alegre, Fernando 1913<br />

Alessandra, Mazzola 1239<br />

Alessio, Loredana 1543<br />

Àlex, Amorós 2069<br />

Alexander, Graeme J. 258<br />

Alexandrino, Paula 150<br />

Alexopoulou, Alexandra 1489, 2019,<br />

2096<br />

Alfadda, Abdulrahman A. 1843<br />

Alghanem, Mansour G. 1246<br />

Alharbi, Asia 1118<br />

Alhassani, Abdulla 1118<br />

Alhasson, Firas 962, 2133, 2137<br />

Ali, Irfan 2173, 2192<br />

Alibrandi, Angela 288, 1898<br />

Alison, Malcolm 101<br />

Aljudaibi, Bandar 204<br />

Aljumah, Abdulrahman A. 1607<br />

Alkaabi, Saad R. 1147, 1188<br />

Alkhatib, Zahraa 344<br />

Allaire, Manon 2082<br />

Allard, Johane P. 2184, 2238, 2254<br />

Allard, Nicole 2060<br />

Allawi, Hussain 892<br />

Allawy, Allawy 330, 334, 518, 651,<br />

667, 668<br />

Allen, Alina M. 6, 2155<br />

Aller, Rocío 2171<br />

Allotey, Loretta K. 1965, 1966<br />

Almagro, Jorge 1295, 1520<br />

Almarzooqi, Saeed 1099<br />

Almazrouei, Sultan 1118<br />

Almeida, Marcio D. 469<br />

Almeida-Dalmet, Swati S. 1527<br />

Almer, Sven 76<br />

Almishri, Wagdi 1290<br />

Aloman, Costica 1266, 1294, 1370,<br />

2124<br />

Alonso, Cristina 953, 2163, 2244<br />

Alonso, Estella M. 133, 1714<br />

Alonso-Guallart, Paula 2129<br />

Aloysius, M. 1037, 1074, 1075<br />

Alpini, Gianfranco 190, 602, 621, 817,<br />

834, 835, 929, 1317, 1349, 1365,<br />

1978<br />

Alqahtani, Saleh 1057<br />

Alric, Laurent 95, 1158, 1803, 2082<br />

Alsaad, Khaled 1607<br />

Alsahhar, Jamil S. 1199<br />

Alsebaey, Ayman 1862, 1879<br />

Alsina, Angel 526, 1784<br />

Alswat, Khalid A. 1607<br />

Alt, Yvonne 1949<br />

Altamirano, José T. 110, 1307, 1327,<br />

2148<br />

Althouse, Sandra K. 436<br />

Altice, Frederick 40<br />

Altinkaya, Engin 2219<br />

Altintas, Mehmet M. 914<br />

Altraif, Ibrahim H. 1607<br />

Aluri, Krishna C. 36<br />

Aluvihare, Varuna 1744<br />

Alvarado, Edilmar A. 734, 1509<br />

Alvarez, Fernando 305<br />

Alvarez, Maria 86, 543<br />

Alvarez-Guaita, Anna 64<br />

Alvaro, Domenico 602, 621, 1654<br />

Alves, Erivaldo 2237<br />

Alves, Venancio Avancini F. 240<br />

Alzaabi, Mohamed 1118<br />

Amaddeo, Giuliana 421, 449, 1232<br />

Amador, Cory 1100<br />

Amadori, Dino 435<br />

Amano, Keisuke 304, 1819<br />

Amaral, Andreia J. 960<br />

Amaral, Joana D. 832<br />

Amarapurkar, Deepak N. 752<br />

Amathieu, Roland 275<br />

Ambade, Aditya 496, 1318, 1330,<br />

1351, 1958, 2127, 2198<br />

Ambroise, Jerome 1734<br />

Ambros, Victor 512<br />

Ameneshoa, Kelly 564<br />

Amer, Aliaa 1147, 1188<br />

Amer, Johnny 660, 1532, 1878<br />

Amin, Janaki 1868<br />

Amin, Saista 1733<br />

Amiot, Bruce 2<br />

Amiot, Xavier 2082<br />

Amitrano, Lucio 742<br />

Amiya, Takeru 181, 2131, 2140<br />

Ammons, Christian E. 1309<br />

Amorosa, Valerianna 1826<br />

Amoroso, Antonio 1<br />

Amory, Catherine 1151<br />

Ampuero, Javier 1134, 1863, 2171<br />

An, Jihyun 446<br />

An, Linjing 354<br />

Anan, Akira 385, 410<br />

Anand, Lovkesh 146, 262, 690, 694,<br />

1340, 1483, 2088<br />

Anand, Ravi 420<br />

Anand, Vijay 1791<br />

Ananthanarayanan, Meena 661<br />

Ananthavarathan, Abbesega 2047<br />

Anastassopoulos, Georgios 775<br />

Anders, Lisanne C. 2141<br />

Andersen, Jesper B. 677<br />

Anderson, Aimee 1687, 1688, 2134<br />

Anderson, James 20<br />

Anderson, Joshua 1479<br />

Anderson, Karl E. 2100, 2111, 2112,<br />

2114<br />

Ando, Satsuki 1474<br />

Andrade, Raul J. 225, 583, 595, 596,<br />

1863, 2171<br />

Andre, Remy 1072<br />

André, Patrice 1652<br />

Andreasson, Anna 158<br />

Andreola, Fausto 908, 1407, 1517<br />

Andreone, Pietro 609, 1086, 1802,<br />

1844, 1888<br />

Andreoni, Massimo 222, 1802, 1844<br />

Andres, Riera 202<br />

Andreu, Victoria 378<br />

Anduze-Faris, Beatrice 217<br />

Angeli, Paolo 148, 266, 268, 292,<br />

767, 775<br />

Angelico, Mario 203, 222<br />

Anguita, Juan 953<br />

Angus, Peter W. 106, 627, 798, 842,<br />

1077, 1225, 1264, 2262<br />

Annable, Tami 190, 835, 929, 1317,<br />

1349<br />

Annamalai, Alagappan A. 1479<br />

Annicchiarico, Brigida E. 432<br />

Annoni, Giorgio 1453<br />

Ansari, Aftab A. 77<br />

Ansari-Gilani, Kianoush 1894<br />

Ansolabehere, Xavier 1052<br />

Anstee, Quentin M. 105, 162, 2145,<br />

2183<br />

Antal-Szalmas, Peter 2090<br />

Anthony, Tracy 919<br />

Antonelli, Barbara B. 1743<br />

Antoniak, Silvio 81<br />

Antonini, Teresa Maria 1237, 1782,<br />

2082<br />

Antonucci, FrancescoPaolo 222<br />

Anty, Rodolphe 95, 145, 264<br />

Anwar, Tarique 327<br />

Anwer, Mohammed S. 582<br />

Anzai, Kazuya 383, 697<br />

Anzai, Keizo 2202<br />

Anzalone, Christopher R. 32, 2022<br />

Aoki, Tomoko 781, 1533, 1975<br />

Aoki, Yoshihiko 476, 794, 1643, 1795,<br />

2158<br />

Aono, Satoshi 26, 669, 1635, 1636,<br />

1661, 1809, 1982, 2130, 2136<br />

Aoudjehane, Lynda 30, 997, 1409<br />

Aouizerat, Brad 2161<br />

Aoyama, Tomonori 917, 1771<br />

Apostolova, Nadezda 1913<br />

Appenrodt, Beate 2077<br />

Applegate, Tanya L. 1083<br />

Apponi, Luciano 510, 2116<br />

Appourchaux, Kevin 998, 1601<br />

Apte, Udayan 650, 1767, 1774<br />

Aqel, Bashar A. 1038, 1124<br />

Arab, Juan P. 2210, 2231<br />

Arachchi, Niranjan J. 442<br />

Aracil, Carles 378<br />

Aragones, Julian 930<br />

Aragri, Marianna 222<br />

Arai, Jun 1302<br />

Arai, Kumiko 226, 917<br />

Arai, Kuniaki 392, 1348, 1653<br />

Arai, Makoto 1207<br />

Arai, Masahiro 1778<br />

Arai, Taeang 291<br />

Arakawa, Tomohiro 1116<br />

Araki, Norie 513<br />

Arama, Victoria 1161<br />

Arancibia, Juan P. 595<br />

Aransay, Ana M. 953<br />

Arase, Yasuji 1062, 1191, 2020, 2241<br />

Arase, Yoshitaka 314, 383<br />

Arauz, Oscar 1264<br />

Araya, Patrick 197


1314A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Arce-Cerezo, Altamira 659<br />

Arden, Katherine J. 1870<br />

Ardevol, Alba 734<br />

Arduino, Jean Marie 717, 729, 1867,<br />

1883<br />

Arenas, Juan I. 378, 2032<br />

Arendale, Evelyn 1178<br />

Arends, Pauline 2002, 2003<br />

Arendt, Bianca M. 2184, 2238, 2254<br />

Arendt, Karen 1111, 1155, 1164<br />

Arfianti, Evi 230<br />

Argentini, Claudio 1153<br />

Argo, Curtis K. 761<br />

Ari, Alpay 1622<br />

Arias, Irwin M. 20, 1909<br />

Arii, Shigeki 1980<br />

Arima, Shiho 937<br />

Arinaga-Hino, Teruko 304, 1819<br />

Ariza, Xavier 266<br />

Armiliato, Geyza N. 2200<br />

Armstrong, Matthew J. 1216<br />

Arnold, Frances 764<br />

Arnold, Hays 106, 627<br />

Aronsohn, Andrew I. 152, 271, 539,<br />

544, 553, 1229<br />

Aronson, Sem J. 492<br />

Arora, Ankur 1340<br />

Arora, Sanjeev 1065<br />

Arora, Zubin 1265<br />

Arosemena, Leopoldo 914<br />

Arput, Jean Pierre 1072<br />

Arranz, Jose Antonio 1509<br />

Arrese, Marco 595, 969, 971, 1217,<br />

2210, 2231<br />

Arriazu, Elena 1382<br />

Arroyo, Vicente 2069, 2071, 2087<br />

Arroyo-Manzano, David 745<br />

Arsalla, Aimal 575<br />

Arslanow, Anita 2253<br />

Arteel, Gavin E. 587, 2141, 2142<br />

Arterbery, Adam 1731<br />

Arterburn, Sarah 1045, 1049<br />

Artru, Florent 1269, 1782, 2222<br />

Arya, Aman V. 1556<br />

Aryafar, Hamed 913<br />

Asagiri, Masa 1918<br />

Asahina, Kinji 1354<br />

Asahina, Yasuhiro 164, 406, 985,<br />

1018, 1027, 1959<br />

Asahina, Yoshiro 699<br />

Asai, Akihiro 676<br />

Asai, Akira 966<br />

Asano, Yasukane 1420<br />

Asaoka, Yoshinari 461<br />

Asara, John 1961<br />

Asatryan, Armen 41, 705, 710, 719,<br />

723<br />

Asch, Steven 1048<br />

Ascherl, Rudi 1222<br />

Asghar, Aliya 724<br />

Asgharpour, Amon 160, 807, 905,<br />

1309, 2163<br />

Ashfaq, Mohammad 1199<br />

Ashouri, Besher 1100<br />

Asmann, Yan W. 358<br />

Aspichueta, Patricia 953<br />

Aspinall, Esther J. 1794<br />

Asrani, Sumeet K. 528, 1210<br />

Assan, Alya 1791<br />

Asselah, Tarik 249, 251, 714, 746,<br />

998, 1036, 1179, 1601, 2044, 2045<br />

Assene, Collins 1091<br />

Assimakopoulos, Stelios F. 2094, 2095<br />

Assis, David N. 298<br />

Ata, Nesrin 2051<br />

Ataman Hatipoglu, Cigdem 2051<br />

Atanasov, Georgi 1222<br />

Atanasov, Petar K. 1176, 1205<br />

Athinarayanan, Shaminie 56<br />

Athwal, Varinder 1364<br />

Atkinson, Elizabeth J. 607, 610<br />

Atreja, Ashish 1151<br />

Atsukawa, Masanori 291, 476, 1097<br />

Attar, Nahid 1776<br />

Attarwala, Husain 36<br />

Atwell, Thomas 293<br />

Aucejo, Federico N. 1271<br />

Aucott, Rebecca 62<br />

Audimoolam, Vinod K. 1588, 1763,<br />

2054<br />

Audureau, Etienne 1700<br />

Auger, Patrice 731<br />

Augustin, Salvador 742, 743<br />

Austin, Andrew 1480, 1524<br />

Austin, Patrick W. 2204<br />

Auth, Marcus K. 1702<br />

Auton, Adam 943<br />

Auzinger, Georg 212, 1216, 1758,<br />

1775<br />

Avagnina, Alejandra 2106<br />

Avani, S. 1475<br />

Avci, Meltem 1622<br />

Avila, Diana 1312<br />

Avila, Meera B. 1513<br />

Avila, Nathaniel P. 1513<br />

Avila, Shanika 190, 817, 834<br />

Avior, Yishai 679<br />

Avitahl-Curtis, Nicole 510, 2116<br />

Avitzur, Yaron 1731<br />

Avner, Ellis D. 1718<br />

Avril, Elisabeth 1796<br />

Ay, Nazli 444<br />

Ayala, Carmen 2257<br />

Ayer, Turgay 104, 151<br />

Ayers, Andrew 646, 1098<br />

Ayonrinde, Oyekoya T. 2185<br />

Ayoub, Walid S. 1479<br />

Aytaman, Ayse 16, 380, 438<br />

Aziz, Karim 264<br />

Aziz, Natali 123<br />

Azoulay, Daniel 215, 863<br />

Azuma, Seishin 164, 406, 1959<br />

Azzam, Ruba K. 1229<br />

Baas, Frank 637<br />

Baba, Hayato 1981<br />

Baba, Hideo 364<br />

Baba, Tomohisa 2119<br />

Bababekov, Yanik 1732<br />

Babatin, Mohammed A. 1607<br />

Bacchetti, Peter 1444<br />

Bach, Nancy 1095, 1185<br />

Bachetoni, Alessandra 983<br />

Back, David J. 1131<br />

Backstedt, David W. 1043, 1109,<br />

1145, 1224<br />

Backus, Lisa I. 93, 1786<br />

Bacon, Bruce 1046, 1108<br />

Badoe, Nina 2205<br />

Badra, Gamal. A. 733, 1853<br />

Badran, Hanaa M. 1862, 1879<br />

Badri, Prajakta 1066, 1084<br />

Badshah, Maaz B. 380, 436, 457,<br />

1903<br />

Bae, Si Hyun 452, 1615, 1882<br />

Baeck, Christer 961<br />

Baeg, Joo Yeong 389, 2113<br />

Baffy, Gyorgy 1469<br />

Bagate, François 1194<br />

Baggett, Sarah 1174<br />

Bahar, Ivet 193<br />

Baheti, Saurabh 600<br />

Bahng, Joseph 871<br />

Bai, Haibo 190, 514, 813, 834, 929<br />

Bai, Lang 1562<br />

Bai, Ming 754<br />

Bai, Shasha 1914<br />

Bai, Wei 468<br />

Baichoo, Esha 397<br />

Baik, Daniel 765, 766<br />

Baik, Soon Koo 736, 1322, 1515,<br />

2080<br />

Bailey, Lorna 1104<br />

Bailey, Shannon 1313<br />

Bailly, François 1158<br />

Bain, Gretchen 141<br />

Bain, Vince 303<br />

Baiocchi, Leonardo 203<br />

Bajaj, Jasmohan S. 52, 82, 570, 1463,<br />

1472, 1485, 1491, 1521, 1527<br />

Bakalov, Vladimir 1720<br />

Baker, Alastair j. 134, 258, 1691<br />

Baker, Talia B. 71, 838<br />

Bakr, Yasser M. 1821<br />

Bala, Shashi 109, 496, 512<br />

Balart, Luis A. 715, 1096<br />

Balasubramaniyan, Natarajan 809,<br />

1688, 1692<br />

Balasubramaniyan, Vairappan 1517<br />

Balba, Gayle 1034<br />

Balcells, Mercedes 1292<br />

Baldan, Angel 65<br />

Baldassarre, Maurizio 1493<br />

Balderramo, Domingo 1752<br />

Baldo, Vincenzo 629<br />

Bale, Shyam Sundhar 25<br />

Balestrieri, Cinzia 309, 1153<br />

Baliga, Prabhakar 1272<br />

Balk, Ethan M. 1592<br />

Balkrishnan, Rajesh 395<br />

Ball, Greg 1051, 1065, 1067<br />

Balladur, Pierre 1409<br />

Ballinger, Brad 873<br />

Ballot, Eric 1782<br />

Balteau, Sylvie 1796<br />

Balwani, Manisha 2111, 2112<br />

Bambha, Kiran 1435, 2085<br />

Banales, Jesus M. 21, 2171<br />

Bañares, Juan 1520<br />

Bañares, Rafael 1295, 1503, 1520<br />

Bandi, Sriram 1764<br />

Bandsma, Robert 2184<br />

Banerjee, Dipti 1263<br />

Banerjee, Rajarshi 931, 2190<br />

Bangma, Sander 798<br />

Bani, Daniele 902<br />

Banks, Alexander S. 881<br />

Bannister, Nicole 1606<br />

Baños, Isolina 1180<br />

Bansal, Meena B. 1095, 1185, 1397


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1315A<br />

Bansal, Sanjay 1704, 1712<br />

Bantel, Heike 2195<br />

Banton, Sophia 19, 2232<br />

Baqir, Altaf 1872<br />

Baquerizo, Angeles 861, 872<br />

Barbaro, Federico 432<br />

Barbati, Zachary 1090<br />

Barbour, April M. 730<br />

Barbour, Youssef 626, 1790, 1798,<br />

1807<br />

Barca, Lucia 309, 1153<br />

Barcia, Rita 1921<br />

Barclay, Murray L. 2262<br />

Barclay, Stephen T. 1052, 1140<br />

Bari, Khurram 743, 873<br />

Barlett, Adam 2225<br />

Barman, Pranab 418<br />

Barn, Vanessa 230<br />

Barnaba, Vincenzo 958<br />

Barnard, Trisha R. 2062<br />

Barnes, David S. 1265<br />

Barnes, Eleanor 637, 2190<br />

Barnes, Maisha 1146, 1169<br />

Barnhart, Huiman X. 225, 1922, 1923,<br />

1930, 1932, 1933<br />

Baron, Todd H. 643<br />

Barr, Eliav 40, 42, 210, 251, 700,<br />

701, 703, 712, 715<br />

Barr Fritcher, Emily G. 169<br />

Barrault, Camille 1102<br />

Barreiro, Pablo 438, 1180, 1881, 1908<br />

Barrera, Francisco 1217, 2210, 2231<br />

Barret, Matthieu 2188<br />

Barrett, Terrence 1267, 2081<br />

Barritt, Alfred S. 14, 533, 2093<br />

Barros, Raffaelle K. 2237<br />

Barroso, Eduardo 1256<br />

Barry, Jeffrey 766<br />

Bartels, Michael 261, 1222<br />

Bartenschlager, Ralf 992<br />

Bartolome, Sonja 9, 290<br />

Barton, Franca B. 286<br />

Barton, Ryan W. 1347<br />

Bartres, Concepció 982<br />

Barve, Shirish 114, 1304, 1312, 1331,<br />

1334<br />

Bashir, Mustafa R. 2150<br />

Bass, Nathan M. 1485, 1761, 2161<br />

Bassissi, Firas 394<br />

Bassit, Leda C. 1544<br />

Basu, Patrick 1037, 1074, 1075<br />

Bataller, Ramon 110, 533, 1327, 1451,<br />

2148<br />

Bate, John P. 106, 627<br />

Bathgate, Andrew 1216<br />

Batra, Sachin 1470, 1513<br />

Battezzati, Pier Maria 73, 634<br />

Batwa, Faisal 1607<br />

Baudesson, Camille 987<br />

Bauer, Christina M. 427<br />

Bauer, David J. 1525<br />

Bauer, Michael 808<br />

Baugé, Eric 887<br />

Baumert, Thomas F. 1019<br />

Baumgarten, Axel 1060, 1081, 1107,<br />

1156<br />

Bautista-Osorio, Eduardo 1022<br />

Baven-Pronk, Martine A. 302<br />

Bayliss, Julianne 1551, 1557, 1563<br />

Baytarian, Michelle 16<br />

Bazinet, Michel 31<br />

Bazzo, Maria Luiza 2075<br />

Beaty, Brenda 137, 1738<br />

Bechmann, Lars 1779, 1924<br />

Becker, Christina 261<br />

Beckerman, Rachel 877, 1534<br />

Beckett, Geoff 1574<br />

Beckford, Miguel 1961<br />

Becquart, Jérôme 1409<br />

Bedard, Patricia W. 1770<br />

Bedossa, Pierre 105, 162, 480, 887,<br />

998, 1601, 2145, 2148, 2164<br />

Beets, Greet 39<br />

Behari, Jaideep 1496, 1531<br />

Behling, Cynthia A. 157, 159<br />

Behrns, Kevin E. 177<br />

Beier, Juliane I. 2141<br />

Beigelman, Leo 993<br />

Beilby, John 2099<br />

Beilin, Lawrence 2185<br />

Beinhardt, Sandra 297, 732, 1070,<br />

1092, 1137<br />

Beiser, David G. 522<br />

Bejarano-Fernandez, Eloy 333<br />

Bekki, Shigemune 408, 783, 1600<br />

Béland, Kathie 305<br />

Belghiti, Jacques 417<br />

Bell, Chaim M. 350<br />

Bell, David 424<br />

Bell, Lauren N. 161<br />

Bell, Phillip D. 962, 2137<br />

Bell, Sally 442, 483<br />

Belle, Steven H. 121, 1598<br />

Bellelli, Giuseppe 1453<br />

Bellentani, Stefano 2164<br />

Belli, Dominique 1735<br />

Belli, Luca S. 571, 1751<br />

Bellini, Giulia 1837<br />

Belloni, Laura 958<br />

Bellos, Damien 2132<br />

Belloula, Djamel 554<br />

Belperio, Pamela S. 93, 1786<br />

Belt, Patricia H. 157, 159<br />

Bemeur, Chantal 51<br />

Ben Ari, Ziv 94, 1032, 1685<br />

Ben Mosbah, Ismail 215<br />

Benbow, Jennifer H. 1389<br />

Bendersky, Noelle 373<br />

Benedetti, Antonio 21<br />

Benetti, Gianpiero 148<br />

Benfatto, Salvatore 1127, 1654<br />

Benidir, Andreanne 1713<br />

Beninati, Concetta 1127, 1654<br />

Benitez, Laura 1881<br />

Benítez, Carlos E. 1217, 2210, 2231<br />

Benítez-Gutiérrez, Laura M. 1180<br />

Benito, Yolanda 1503<br />

Benjamin, Jaya 1328, 1340, 2088,<br />

2203<br />

Benjaram, Sindhuri 420<br />

Benkert, Pascal 1884<br />

Benlloch, Salvador 1230, 1251<br />

Bennai, Yacia 206<br />

Bennett, Michael 248, 250, 1039<br />

Benson, Ariel 1878<br />

Benyashvili, Tamara 1172<br />

Benzing, Christian 1222<br />

Beppu, Toru 364<br />

Berardi, Sonia 1751<br />

Berenguer, Marina 1084, 1230, 1251<br />

Bereshchenko, Oxana 1963<br />

Beret, Antoine 394<br />

Beretta, Laura 170, 1424, 2151<br />

Berg, Christoph P. 76, 1055, 1183<br />

Berg, Thomas 37, 245, 249, 252, 261,<br />

276, 308, 1058, 1594, 1638, 1678,<br />

1861, 2003, 2004, 2012<br />

Bergeat, Damien 487<br />

Berger, Calise K. 169<br />

Bergin, Colm J. 1103<br />

Bergquist, Annika 76<br />

Berhane, Sarah 344, 851<br />

Berk, Paul D. 972<br />

Berkane, Saadi 1621<br />

Berkes, Jamie 1248<br />

Berkowski, Caterina 980, 1055, 1559<br />

Bermejo, Javier 1503<br />

Bernabucci, Veronica 1751<br />

Bernal, William 212, 740, 1775<br />

Bernard, Denis 1239<br />

Bernard-Chabert, Brigitte 1072<br />

Bernardi, Mauro 1493, 2087<br />

Bernstein, David 726, 1039, 1051,<br />

1065, 1067, 1086<br />

Berntsen, Natalie L. 819<br />

Bernuzzi, Francesca 602, 621<br />

Berres, Marie-Luise 1364<br />

Berry, Kristin 49, 1211<br />

Bertelli, Cristina 393<br />

Bertero, Alessandro 258<br />

Bertinetto, Francesca E. 1<br />

Bertoletti, Antonio 167<br />

Besch, Camille 3, 95<br />

Beschorner, Niklas 1629<br />

Besisik, Fatih 478<br />

Bessho, Ryosuke 1981<br />

Bessone, Fernando 225, 595<br />

Best, Jan 1779<br />

Beste, Lauren A. 128, 540<br />

Besur, Siddesh 17<br />

Betrapally, Naga 1463, 1527<br />

Bettencourt, Ricki 913, 1431, 2201<br />

Bettinger, Dominik 1803<br />

Beudeker, Boris 1007<br />

Beuers, Ulrich 74, 76, 317, 492, 604,<br />

623, 637, 827, 1995, 2152<br />

Beumont, Maria 39<br />

Bevans-Fonti, Shannon 1422<br />

Beyer, Jill 705<br />

Beysen, Carine 1425<br />

Bezerra, Jorge A. 133, 676<br />

Bhadoria, Ajeet S. 1268, 1328, 2203<br />

Bhalla, Ashish 259<br />

Bhamidimarri, Kalyan R. 200, 1054,<br />

1148, 1206, 1587, 2256<br />

Bhandari, Bal R. 248, 250<br />

Bhanji, Rahima A. 303, 1481<br />

Bhardwaj, Ankit 1328, 1340, 1659,<br />

2203<br />

Bhat, Sathesh P. 957, 1938<br />

Bhatia, Vikram 2070, 2088<br />

Bhatnagar, Raj K. 1659<br />

Bhatt, Archana 913, 1431<br />

Bhattacharya, Dipankar 1392<br />

Bhattacharya, Renuka 1221<br />

Bhoori, Sherrie 1751<br />

Bhore, Rafia 706, 716, 726<br />

Bhuket, Taft 7, 237, 1252, 2159<br />

Bhushan, Bharat 1767<br />

Bianchi, Ilaria 602


1316A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Bibert, Stephanie 221<br />

Bichoupan, Kian 1095, 1151, 1185,<br />

1202<br />

Bider-Canfield, Zoe 1135, 1171<br />

Bieghs, Veerle 100<br />

Bigaeva, Emilia 1434<br />

Biggins, Scott W. 82, 570, 2085<br />

Bihari, Chhagan 327, 690, 694, 1483,<br />

2078<br />

Bilasy, Shymaa E. 858, 2215<br />

Biliotti, Elisa 222, 983<br />

Billaud, Eric 568, 1093<br />

Billiar, Timothy R. 656<br />

Billingsley, Kevin 481<br />

Bilodeau, Marc 179, 204<br />

Bilzer, Manfred 1170, 1200<br />

Binder, Christoph J. 922<br />

Bindmann, Julia 98<br />

Biolato, Marco 432<br />

Birerdinc, Aybike 2205<br />

Birnbaum, Morris J. 65<br />

Bisch, Grégoire 1409<br />

Biselli, Maurizio 1493<br />

Bishnu, Saptarshi 750<br />

Bissell, D Montgomery 2111, 2112<br />

Bissonnette, Julien 772, 2148<br />

Bizzotto, Paola 775<br />

Bjoernetroe, Tonje 819<br />

Bjornsson, Einar 225<br />

Blach, Sarah 1818, 1842<br />

Black, Sylvester 874, 1260, 1280<br />

Blacklidge, Judith 200<br />

Blackman, Brett R. 175, 664, 915,<br />

1916<br />

Blanc, Etienne 30<br />

Blanc, Jean-Frédéric 132, 450, 851<br />

Blanc, Pierluigi 1802, 1844<br />

Blanco, Albert 1509<br />

Blandino, Giovanni 958<br />

Blank, Emma E. 507<br />

Blas-García, Ana 1913<br />

Blasco-Perrin, Hélène 2082<br />

Blasi, Maria 1021, 1674<br />

Blatt, Lawrence M. 993<br />

Bloch, Mark 210<br />

Block, Timothy 371<br />

Bloomer, Joseph R. 2111, 2112<br />

Blumberg, Richard S. 819, 833<br />

Blumenthal, Antje 1362<br />

Bobardt, Michael 2035<br />

Boberg, Kirsten M. 76, 602<br />

Bochud, Pierre-Yves 221<br />

Bock, Izona 1874<br />

Bockamp, Ernesto 183<br />

Bode, Konrad A. 618<br />

Boeck, Christina 444<br />

Boehlig, Albrecht 276<br />

Boehm, Stephan 245, 276, 1594, 1638<br />

Boehme, Shannon M. 885<br />

Boeker, Klaus H. 1861<br />

Boekschoten, Mark V. 1924<br />

Boemio, Adriana 1837<br />

Boesecke, Christoph 1060<br />

Boettler, Tobias 2195<br />

Boffelli, Silvia 1493<br />

Bohorquez, Humberto E. 1242<br />

Boin, Ilka F. 469<br />

Boix, Loreto 65<br />

Bojito-Marrero, Lizza 295, 1181<br />

Bolcas, Paige 823<br />

Boldt, Andreas 261<br />

Boleslawski, Emmanuel 487, 1269<br />

Bolier, Ruth 827<br />

Bollipo, Steven J. 106, 627<br />

Bolognesi, Massimo 292, 767<br />

Bombonato, Giancarlo 292, 775<br />

Bond, Geoffrey 763<br />

Bondy, Carolyn 1720<br />

Bonello, John 2251<br />

Bonkovsky, Herbert L. 597, 1023, 1923,<br />

2111, 2112<br />

Bonnefont-Rousselot, Dominique 786<br />

Bono, Patrizia 1747<br />

Bonomo, Lorenzo 432<br />

Boodram, Basmattee 1854<br />

Boone, Pamela 1229<br />

Boonsirichan, Rattana 1631, 1662<br />

Boonstra, Andre 1007, 1196, 1585,<br />

1586, 1590, 1637, 2122<br />

Boonstra, Kirsten 74, 623<br />

Boor, Peter 961<br />

Booty, Jordan 521<br />

Bopp, Tobias 1353<br />

Borad, Mitesh J. 396, 1965, 1966<br />

Borchardt, Hannes 1299, 1385<br />

Borg, Brian B. 1178<br />

Borgeaud, Nathalie 665<br />

Borghesan, Michela 1445<br />

Borgia, Guglielmo 1802, 1844<br />

Borgne, Anne 1796<br />

Borok, Zea 1945<br />

Borralho, Pedro M. 176, 897, 960<br />

Borralho Nunes, Paula 940<br />

Borst, Andrew 1316<br />

Bortoluzzi, Ilaria 1270<br />

Borude, Prachi C. 1774<br />

Borzio, Franco 355<br />

Boscarino, Joseph A. 15, 525, 1578,<br />

1789, 1799, 1800<br />

Bosch, Jaime 326, 737, 739, 745, 746,<br />

1418, 1518, 2239<br />

Bosma, Piter J. 492, 827<br />

Bosman, Dean 724<br />

Bosoi, Cristina R. 51, 1520, 1523<br />

Bossi, Krista 2166<br />

Boström, Adrian 519<br />

Bota, Simona 270, 475, 732, 755, 771<br />

Botta-Fridlund, Danielle 3, 95, 206<br />

Bottai, Matteo 116<br />

Bottari, Antonio 288<br />

Bottero, Julie 1677<br />

Botwin, Gregory J. 724, 1129<br />

Bouattour, Mohamed 417, 480<br />

Boudjema, Karim 487<br />

Boudon, Marc 1237<br />

Boueyre, Estelle 1194<br />

Boulanger, Chantal 887, 2148<br />

Boulter, Luke G. 678<br />

Bourcier, Valérie 362, 1808, 1847<br />

Bourgeois, Stefan 797, 1051, 1091,<br />

1842<br />

Bourhis, Francois 1072<br />

Bourlière, Marc 206, 249, 394, 777,<br />

1442, 1860, 2082<br />

Boursier, Jerome 117, 1439, 2188<br />

Bouvier, Antoine 263<br />

Bouvier-Alias, Magali 1123<br />

Bouyssou, Caroline 132<br />

Bovitz, Tanya 1867, 1883<br />

Bowden, Scott 1557, 1563, 1566<br />

Bowe, Andrea 687<br />

Bowen, David 1243<br />

Bowen, William C. 670, 1979<br />

Bowlus, Christopher L. 76, 77, 602,<br />

606, 609, 613, 616, 624, 630, 639,<br />

1034<br />

Bowring, Mary G. 1749<br />

Bowyer, Teresa 1588, 2053<br />

Box, Terry D. 2257<br />

Boyd, Anders 1676, 1677<br />

Boyer, James L. 23, 298, 805, 829<br />

Boyer, Nathalie 206, 998, 1601, 2044,<br />

2045<br />

Boyer, Thomas D. 278, 1459<br />

Boyett, Sherry L. 905, 2235, 2249<br />

Bozdayi, A.Mithat 2033<br />

Bozorgzadeh, Adel 845<br />

Bradley, J. Andrew 258<br />

Brady, Joanne E. 1792, 1839<br />

Brahm, Javier 595<br />

Brahmania, Mayur 542, 1556<br />

Braid-Forbes, Mary Jo 531<br />

Brain, John G. 611<br />

Brainard, Diana M. 38, 91, 96, 205,<br />

219, 247, 249, 713, 718, 746, 1034,<br />

1041, 1049, 1080, 1094, 1120,<br />

1128, 1442, 1860<br />

Brancatelli, Santa 1898<br />

Branch, Andrea D. 29, 229, 988, 1095,<br />

1112, 1151, 1185, 1202, 1289,<br />

1397, 1742<br />

Branco, Fernanda 469<br />

Brandan, Enrique 969<br />

Brandl, Katharina 59<br />

Brandman, Danielle 414, 1273<br />

Brandon-Warner, Elizabeth 1389<br />

Brantly, Mark 138<br />

Braspenning, Joris 331<br />

Bratthauer, Gary 891, 956<br />

Bratton, Charles F. 1272<br />

Bräu, Norbert 205, 380, 438, 457<br />

Braun, Felix 76<br />

Braun, Marius 875<br />

Breckenridge, David 1359<br />

Breen, Gerome 2105<br />

Brejda, John 725<br />

Bremer, Birgit 1572, 1597, 2026<br />

Bremer, Corinna M. 2026<br />

Brenner, David A. 59, 186, 187, 189,<br />

890, 1415, 1425, 1431<br />

Breuer, Monika 46<br />

Bricault, Ivan 421, 449<br />

Bridge, Robert S. 1969, 2139<br />

Bridges, Brian 352, 1291, 1814, 1815<br />

Bridges, John F. 1727, 1729<br />

Bridle, Kim 818, 975<br />

Brigham, Benjamin S. 36<br />

Brigstock, David 192, 1356, 1404<br />

Brimacombe, Michael B. 1774<br />

Briones-Orta, Marco A. 1978, 2143<br />

Briot, Charline 264<br />

Brisac, Cynthia 25, 991, 1015, 1630,<br />

1644<br />

Britton, Laurence J. 818, 975<br />

Brivio, Simone 1950<br />

Brixko, Christian 1842<br />

Brock, Guy N. 2233<br />

Brodsky, Allison 1864, 1880<br />

Bronich, Tatiana 589<br />

Bronk, Steven 825, 895


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1317A<br />

Bronowicki, Jean-Pierre 206, 2082<br />

Brookes, Steve 424<br />

Brookmeyer, Ron 1068<br />

Brooks, Alexandria 536<br />

Brooks, Louis 1805<br />

Brosch, Mario 617<br />

Broschewitz, Johannes 1222<br />

Brouillet, Arthur 693, 1391<br />

Brouwer, Linda 1416<br />

Brouwer, Willem Pieter 348, 1568,<br />

1585, 1586, 1590, 1637, 2001, 2003<br />

Brown, Ashley S. 1826, 2003, 2047<br />

Brown, Bob D. 510, 2116<br />

Brown, Deborah D. 712, 717<br />

Brown, Helen 44<br />

Brown, Kimberly Ann 307, 1084, 1792<br />

Brown, Maxwell A. 1754<br />

Brown, Robert 1037, 1049, 1074,<br />

1075<br />

Brown, Robert E. 905<br />

Brown, Robert S. 217, 272, 369, 1084,<br />

1454<br />

Bruce, David S. 1242<br />

Bruce, Matthew J. 1113, 1588, 1704,<br />

1744, 2053, 2054<br />

Bruce, Michael 1790, 1798<br />

Bruce, Robert 1874<br />

Bruckbauer, Antje 933<br />

Bruden, Dana J. 1790, 1798, 1807<br />

Bruggmann, Philip 1818<br />

Bruha, Radan 1487<br />

Bruix, Jordi 65, 378, 860<br />

Brumme, Chanson J. 1832<br />

Brun, Sonia 394<br />

Brunetto, Maurizia R. 1802, 1844<br />

Brunner, Jim 1839<br />

Bruno, David A. 1274<br />

Bruno, Raffaele 1802, 1844<br />

Bruns, Tony 73<br />

Brunt, Elizabeth M. 157, 159, 2150,<br />

2220<br />

Bruscella, Patrice 987<br />

Brusch, Lutz 617<br />

Brusilovskaya, Ksenia 1525<br />

Brusquant, David 417<br />

Bryan, William E. 1142<br />

Bucala, Richard 298<br />

Buchanan, Charlotte E. 1480<br />

Buchholz, Frank 1629<br />

Bucsics, Theresa 270, 771, 1204<br />

Budoff, Matthew 1488<br />

Buechler, MarkusW. 1227<br />

Buggisch, Peter 1078, 1170, 1176,<br />

1179, 1200, 1205, 1559, 1861, 2257<br />

Bugianesi, Elisabetta 973, 2164, 2209<br />

Buist-Homan, Manon 591<br />

Buitrago, Laura E. 2128<br />

Bukong, Terence N. 1413, 2198<br />

Bulkow, Lisa 125<br />

Bull, Laura 78<br />

Bulut, Cemal 2051<br />

Bunchorntavakul, Chalermrat 1631,<br />

1662<br />

Bunniran, Suvapun 1050<br />

Bunraksa, Worawan 2089<br />

Burchill, Matthew A. 28<br />

Bureau, Christophe 145, 263, 264,<br />

744, 2082<br />

Burgess, Gary 798<br />

Burke, Ann E. 1452<br />

Burks, Deborah J. 691<br />

Burns, Justin M. 50, 53<br />

Burra, Patrizia 268, 351, 1270, 1889<br />

Burrin, Douglas 194, 1689<br />

Busch, Heiner W. 1060, 1081, 1156<br />

Bushau, Adrienne M. 2141<br />

Bushyhead, Daniel W. 1218<br />

Busuttil, Ronald W. 1283, 2138<br />

Buti, Maria 241, 243, 1021, 1179,<br />

1674, 2003, 2012, 2014, 2032,<br />

2037, 2059<br />

Butt, Adeel 1488<br />

Butt, Mohammed T. 1147<br />

Butte, Atul 497, 500<br />

Butterfield, David 673<br />

Butterton, Joan R. 701, 717, 725, 730,<br />

731<br />

Butterworth, Jacqueline 652<br />

Buzzigoli, Emma 2209<br />

Byrne, Sean 1041<br />

Byun, Kwan Soo 391, 405, 1504,<br />

1897, 2017<br />

Bzeizi, Khalid I. 1607<br />

Bzowej, Natalie H. 1242<br />

Caballeria, Juan 1307, 1327, 2148<br />

Caballeria, Llorenç 73, 629, 634<br />

Cabarrou, Pauline 263<br />

Cabello-Verrugio, Claudio 969<br />

Cabellos, Laia 253<br />

Cabibbo, Giuseppe 355<br />

Cable, Rebecca 2204<br />

Cabrera, Daniel 969, 971<br />

Caccamo, Gaia 288, 1476<br />

Caccamo, Lucio 1743<br />

Cacciarelli, Thomas V. 1753<br />

Cacciola, Irene 288, 1127, 1476<br />

Cachero, Alba 742<br />

Cacoub, Patrice 2082<br />

Cadamuro, Massimiliano 1950<br />

Cadranel, Jean François D. 264, 554,<br />

2082<br />

Cai, Dachuan 991, 1015, 1630, 1644<br />

Cai, Hongwei 468<br />

Cai, Hou Ming 197<br />

Cai, Luke 1284<br />

Cai, Shi-Ying 805, 2179<br />

Cai, Yan 110, 903<br />

Caiazzo, Robert 2222<br />

Caines, Allyce 532, 559<br />

Caire, Thure 526, 1240<br />

Cakaloglu, Yilmaz 558, 2001<br />

Cakmak, Erol 2219<br />

Calderaro, Julien 421, 449, 693, 1232<br />

Caldwell, Stephen H. 13, 105, 162,<br />

737, 761, 1428, 2145, 2239<br />

Cales, Paul 105, 117, 162, 741, 2145,<br />

2188<br />

Caligiuri, Alessandra 1410<br />

Calinas, Filipe 1859<br />

Caliri, Kimberley 2257, 2259, 2260<br />

Caliskan Kartal, Aysun 2033<br />

Calleja, José Luís 746, 1511, 2012,<br />

2032, 2171<br />

Calmus, Yvon 30, 997, 1239, 1409<br />

Caloggero, Simona 288<br />

Calvisi, Diego 1934, 1948<br />

Cam, Margaret 218<br />

Camacho, Jocelyn A. 1870<br />

Cambe, Joy 1870<br />

Cambula, Linda 974<br />

Cameron, Andrew M. 1255, 1727,<br />

1729, 1749<br />

Campana, Lara 62<br />

Campanale, Mariachiara 432<br />

Campbell, Andrew L. 705<br />

Campbell, Catherine M. 818<br />

Campbell, Deanna J. 1993<br />

Campbell, Jean S. 1395<br />

Campos, Frederico F. 2189<br />

Campos, Laura 1991<br />

Campos, Priscila B. 2221<br />

Camprecios, Genis 253, 254<br />

Campsen, Jeffrey 200<br />

Canbay, Ali 1779, 1924, 1978, 2143<br />

Canchola, Jesse A. 1597<br />

Cancino, Alejandra 1217<br />

Candotti, Fabio 1717<br />

Canini, Irene 1153<br />

Canini, Laetitia 982<br />

Canizaro, Leticia 722<br />

Cannon, Ryan 1069<br />

Canva, Valerie 3, 2222<br />

Cao, Chuanhui 1939<br />

Cao, Furong 1541, 1579, 1617, 1773<br />

Cao, Jiawei 1998<br />

Cao, Sheng 1315, 1355, 1406<br />

Cao, Xiaoming 169<br />

Cao, Ying 2066<br />

Cao, Yirong 131<br />

Cao, Yumei 1048<br />

Cao, ZhuJun 1611<br />

Cao, Zongxian 1374<br />

Caparosa, Susan 87<br />

Capocelli, Kelley E. 1709, 1715<br />

Capoluongo, Nicolina 1837<br />

Cappellini, Sara 309<br />

Cappoli, Andrea 983<br />

Cappon, Andrea 973<br />

Caprio, Nunzio 1543<br />

Caraceni, Paolo 1493, 1751, 1802,<br />

1844<br />

Carbone, Marco 73, 155, 601<br />

Carbonell, Nicolas 263, 264<br />

Carcassi, Carlo 309<br />

Cardenas, Andres 2084<br />

Cardenas, Victor M. 1857<br />

Cardillo Marricco, Nadia 725, 731<br />

Cardoso, Alberto 473<br />

Cardoso, Filipe S. 1759<br />

Cardoso, Helder 1540<br />

Cardoso de Brito, Miguel 258<br />

Cardozo, Karina H. 48<br />

Carey, Elizabeth J. 106, 605, 627,<br />

1468<br />

Carey, Ivana 1098, 1113, 1117, 1132,<br />

1173, 1571, 1588, 1626, 1704,<br />

1744, 2003, 2053, 2054<br />

Caridade, Marta 897, 940<br />

Carito, Brenda 36<br />

Carl, Daniel 1521<br />

Carli, Fabrizia 2209<br />

Carlson, Rolf I. 256, 1962<br />

Carlton-Smith, Charles 1015<br />

Carmody, Ian C. 1242<br />

Carmona, Damien 1102<br />

Caro, Luzelena 707, 725, 731<br />

Carpentier, Arnaud 30<br />

Carpino, Guido 817<br />

Carr, Brian I. 437<br />

Carr, David 260, 753


1318A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Carr, Kathy 1433<br />

Carr, Rotonya M. 1314<br />

Carrat, Fabrice 206<br />

Carrera, Enrique 595<br />

Carriero, Damaris C. 1112<br />

Carrilho, Flair J. 48, 240, 469, 1033,<br />

1544, 2221<br />

Carter, Jeffrey D. 1141<br />

Carvalho, Armando 473<br />

Carvalho, Catarina C. 897<br />

Carvalho, Tânia 940<br />

Carvalho, Valdemir M. 48<br />

Casadei Gardini, Andrea 435<br />

Casado, Marta 2032<br />

Casale, Michele 309, 1153<br />

Casalino, Paolo 222<br />

Casals, Gregori 659<br />

Casanovas, Teresa 2032<br />

Cascinu, Stefanu 435<br />

Cascorbi, Ingolf 225<br />

Casey, Carol A. 1310, 1969, 2139<br />

Casey, Stephen 798<br />

Casillas, Rosario 1021, 1674<br />

Casillas-Ramírez, Araní 342<br />

Casingal, Vincent P. 17<br />

Cassader, Maurizio 2209<br />

Cassagnol, David 1112<br />

Cassiman, David 171<br />

Cassinotto, Christophe 450<br />

Cassio, Doris 816<br />

Cast, Ashley E. 680<br />

Castaing, Denis X. 487, 1237<br />

Castaño, Gustavo O. 2217<br />

Castellani, Paul 394<br />

Castellote, Jose 742<br />

Castelnau, Corinne 2044, 2045<br />

Castera, Laurent 417, 480<br />

Castiella, Agustin 583<br />

Castiello, Filomena 431<br />

Castillo, Alvaro E. 481<br />

Castillo, Eliana 1576<br />

Castro, Azucena 620, 2244<br />

Castro, Marcelo 2106<br />

Castro, Rui E. 176, 832, 897, 940, 960<br />

Castro Filho, Elio C. 84<br />

Castroagudin, Javier F. 378<br />

Casu, Stefania 309, 1153<br />

Casulleras, Mireia 2087<br />

Casuscelli di Tocco, Francesca 1898<br />

Catalano, Donna 496, 512<br />

Catalina, Maria Vega 1503<br />

Catalli, Lisa 1825<br />

Catone, Stefania 1153<br />

Cattral, Mark S. 11<br />

Caturelli, Eugenio 355<br />

Causse, Xavier 1072, 2082<br />

Cavalcanti, Marta 762<br />

Cavallin, Marta 148<br />

Cave, Matthew C. 666, 1303, 2141,<br />

2233<br />

Cavenago, Margherita 1743<br />

Caviglia, Gian Paolo 1740<br />

Cazanave, Sophie C. 160, 905<br />

Cazier, Alain 554<br />

Cazzagon, Nora 73, 299, 629<br />

Cebotarescu, Valentin 31<br />

Ceccherini-Silberstein, Francesca 222<br />

Cedars, Marcelle I. 2252<br />

Cehic, Damir G. 1021<br />

Celik, Neslihan 763<br />

Cellai, Ilaria 882, 902<br />

Ceni, Elisabetta 949, 1963<br />

Cento, Valeria 222<br />

Cereser, Biancastella 101<br />

Cerny, Andreas 221, 1884<br />

Cerocchi, Orlando 1099, 1556<br />

Cervoni, Jean Paul 264, 1036, 1187<br />

Cesana, Giancarlo 571<br />

Cesario, Valentina 432<br />

Cha, Seung-Kuy 1401<br />

Chaabna, Karima 1829<br />

Chagas, Aline 48, 469<br />

Chagneau-Derrode, Carine 2082<br />

Chahal, Barinder 1413<br />

Chainuvati, Siwaporn 467<br />

Chaiprasert, Amnart 2089<br />

Chait, Alan 923<br />

Chaiteerakij, Roongruedee 127, 360,<br />

397, 416, 1497, 1965, 1966<br />

Chakrabari, Subrata 1246<br />

Chakraborty, Shweta 1814, 1815<br />

Chakraborty, Souvik 197<br />

Chalasani, Naga P. 225, 236, 239,<br />

436, 607, 624, 948, 1316, 1332,<br />

1903, 1922, 1923, 1926, 1930,<br />

1933, 2154, 2166, 2175, 2220<br />

Challies, Tracy L. 2160<br />

Chamberland, Robin 1535<br />

Champion, Mia D. 1965<br />

Chamulitrat, Walee 1411<br />

Chan, Albert 70, 1746<br />

Chan, Anthony W. 238, 2181<br />

Chan, Che-Chang 965<br />

Chan, Henry Lik-Yuen 235, 238, 241,<br />

349, 1120, 1548, 1557, 1563, 1612,<br />

1651, 2014, 2025, 2181<br />

Chan, Kelvin K. 350<br />

Chan, Sarah Y. 307<br />

Chan, See-Ching 70, 1746, 2000<br />

Chan, Stephen L. 344<br />

Chandhok, Gursimran 684, 2103<br />

Chandra, Partha K. 501, 506, 1096<br />

Chandrasekar, Edwin 579<br />

Chandrasekaran, Suki 2091<br />

Chandrashekaran, Varun 962<br />

Chang, Albert 471<br />

Chang, Binxia 903<br />

Chang, Chao-Hsuan 619<br />

Chang, Charissa Y. 1095, 1185<br />

Chang, Chi-Yang 1547<br />

Chang, Chia-Lin 1593, 1604<br />

Chang, Christine Y. 123, 359, 388,<br />

1182, 1189, 1806, 1816, 1848, 2009<br />

Chang, Chung-Chou 1488<br />

Chang, David C. 1732<br />

Chang, Ellen T. 1668<br />

Chang, Hye Young 1399<br />

Chang, Jason 451<br />

Chang, Kuo-Chin 1873, 1942<br />

Chang, Kyong-Mi 1001<br />

Chang, Mei-Hwei 1701, 1705<br />

Chang, Na 1396<br />

Chang, Paul 765, 766<br />

Chang, Sanders 1202<br />

Chang, Ting-Tsung 1554<br />

Chang, U Im 1615<br />

Chang, Will 2247<br />

Chang, Ya-Ju 654<br />

Chang, Yung-Sheng 1639<br />

Chang, Young 474, 479, 1610, 2016<br />

Channick, Richard 9<br />

Chao, Xiaojuan 598<br />

Chapkin, Robert 56<br />

Chapman, Roger W. 74, 616, 622,<br />

624, 637<br />

Chapman, William C. 704<br />

Charatcharoenwitthaya, Phunchai 467<br />

Charbek, Edward 1535<br />

Charier, Florian 95<br />

Charisse, Klaus B. 36<br />

Charles, Edgar 712<br />

Charles, Hoppel 518, 651, 667<br />

Charlot, Hélène 1102<br />

Charlton, Carmen L. 1576<br />

Charlton, Michael R. 96, 616, 746,<br />

1045, 1049, 1130, 1226, 1359,<br />

1442, 2149<br />

Charnaux, Nathalie 362<br />

Charuworn, Prista 241, 2025<br />

Chasan, Rachel 1870<br />

Chascsa, David M. 1498<br />

Chatterjee, Saswata 750<br />

Chatterjee, Saurabh 962, 2133, 2137<br />

Chatterji, Udayan 2035<br />

Chattieng, Piyanat 360<br />

Chaturvedi, Prasoon 36<br />

Chau, Gar-Yang 439<br />

Chaubey, Pururawa M. 806<br />

Chaudhry, Asad A. 1872<br />

Chaudhry, Hunana 1864, 1880<br />

Chaudhry, Sulemon 2129<br />

Chauhan, Ranjit 1641<br />

Chaung, Kevin T. 207, 368, 1561<br />

Chava, Srinivas 501, 506, 1096<br />

Chavan, Shailesh 200<br />

Chavez, Deborah 32, 2022<br />

Chavin, Kenneth D. 1272<br />

Chawla, Yogesh K. 214<br />

Chayama, Kazuaki 472, 663, 707,<br />

986, 1056, 1150, 1627, 1645, 1649,<br />

1670, 1679, 1681, 1858, 1893,<br />

1895, 1944, 2120, 2156, 2199, 2202<br />

Chazouillères, Olivier 76, 95, 624, 642<br />

Che, Li 1934<br />

Cheetham, T. Craig 87, 1135, 1171,<br />

1788<br />

Chekuri, Sweta 1095, 1202<br />

Chelis, Leonidas 1900<br />

Chelvaratnam, Uthayanan 1264<br />

Chemaitelly, Hiam 1829<br />

Chemello, Liliana 1802, 1844<br />

Chemnitz, Jan 1629<br />

Chen, Anping 911, 1361, 1365<br />

Chen, Bin 497, 500<br />

Chen, Chi-Hua 1431<br />

Chen, Chi-Ling 1996<br />

Chen, Chi-Yi 1996<br />

Chen, Chia-Lin 1945<br />

Chen, Chien-Hung 1873, 2010<br />

Chen, Chien-Jen 207, 1561, 1593,<br />

1604, 1828<br />

Chen, Chien-Yu 1945<br />

Chen, Ding-Shinn 2027<br />

Chen, Dongyu 510, 2116<br />

Chen, Erluo 700<br />

Chen, Feng 680<br />

Chen, Guang 1656, 1683, 1684<br />

Chen, Gui-Hua 419, 440<br />

Chen, Guofeng 1536, 1537, 2008<br />

Chen, Hank H. 1450


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1319A<br />

Chen, Hong-Song 1973<br />

Chen, Huan 353, 375<br />

Chen, Huey-Ling 1701, 1705<br />

Chen, Jian 130, 682, 2118<br />

Chen, Jianhong 1595<br />

Chen, Jiegen 1327<br />

Chen, Jing 1536, 1537, 2008<br />

Chen, Jiun-Sheng 130, 682<br />

Chen, Li 192, 1356, 1404<br />

Chen, Liang 1998<br />

Chen, Limin 981<br />

Chen, Long 782<br />

Chen, Michael 1986<br />

Chen, Ming-Yao 1996<br />

Chen, Minjun 596<br />

Chen, Minshan 851<br />

Chen, Pei-Jer 1593, 1604, 2027<br />

Chen, Peng 111, 1319<br />

Chen, Ping-Hsien 738, 751, 757, 759<br />

Chen, Po-Hung 1214, 1749<br />

Chen, Ruju 192, 1356<br />

Chen, Shi-Jun 2023<br />

Chen, Shiyao 131<br />

Chen, Steve 2175<br />

Chen, Tao 1660<br />

Chen, Tianyan 1541, 1579, 1617,<br />

1773<br />

Chen, Ting-Yi 380, 438, 457<br />

Chen, Tsung-Ming 1996<br />

Chen, Vincent L. 367, 387, 428, 459,<br />

1668, 2009<br />

Chen, Wei 454<br />

Chen, Wei-Yang 1304<br />

Chen, Weina 686<br />

Chen, Wen-Chi 738<br />

Chen, Xiangmei 1634<br />

Chen, Xin 1934, 1948<br />

Chen, Xinyue 2048, 2061<br />

Chen, Yi-Cheng 1552<br />

Chen, Yi-Ju 345, 1589<br />

Chen, Ying 1693<br />

Chen, Ying-chun 619<br />

Chen, Yonglin 979, 2179<br />

Chen, Yu 1382, 1940<br />

Chen, Zhaochun 1646<br />

Chen, Zhi-wei 1044<br />

Chen, Zishuo 904<br />

Cheng, Frank 285<br />

Cheng, Jin 1634<br />

Cheng, Jin-Shiung 738<br />

Cheng, Jun 2013<br />

Cheng, Li-sha 1383<br />

Cheng, Linling 1430<br />

Cheng, Rex 1221<br />

Cheng, Stephen 1169<br />

Cheng, Sunny Lihua 655<br />

Cheng, Wendy 2262<br />

Cheng, Yang 167<br />

Cheon, Gab Jin 277, 484, 1515<br />

Cheon, HyeonJoo 1025<br />

Cheong, Jae Youn 379, 1567<br />

Cheraitia, Salima 1621<br />

Chéret, Antoine 568, 1093<br />

Cherk, Erika 549<br />

Chermak, Faiza 450, 2181<br />

Cherniack, Andrew D. 129<br />

Cheroni, Cristina 2011<br />

Cherqui, Daniel 449, 487, 1237<br />

Chessa, Luchino 309, 1153, 1802,<br />

1844<br />

Cheung, Angela C. 73, 634<br />

Cheung, Cindy K. 1746<br />

Cheung, Ka-Shing 126<br />

Cheung, Michael 1261, 1276<br />

Cheung, Michelle C. 1173<br />

Cheung, Tan To 70, 1746<br />

Chevaliez, Stéphane 999, 1015, 1123,<br />

1700, 1797, 1833<br />

Chhatwal, Jagpreet 104, 151<br />

Chi, Aileen 1560<br />

Chi, Benjamin 1575<br />

Chi, Heng 1998, 2002, 2003, 2012<br />

Chi, Xiumei 1149, 1193, 1603, 1869<br />

Chi-Cervera, Luis A. 1471<br />

Chiang, John 885<br />

Chiang, Kevin 87, 1788<br />

Chiaradia, Mélanie 421, 449<br />

Chiaramonte, Maria 355<br />

Chiba, Tsutomu 1989<br />

Chiba, Yousuke 1369<br />

Chibaudel, Benoist 417<br />

Chida, Takeshi 1393<br />

Chie, Wei-Chu 1701<br />

Chien, Rong-Nan 1552, 1996<br />

Chikada, Hiromi 697, 1369<br />

Chikhi, Yazid 1621<br />

Childs, Kate 1113<br />

Childs, Kate E. 646, 1098, 1117, 1132<br />

Chim, Angel ML 238, 1548, 2181<br />

Ching, Helena 2185<br />

Chinnasamy, Thirunavukkarasu 340<br />

Chiodo, Letizia 165<br />

Chiou, Yu-Wei 1639<br />

Chipumuro, Edmond 2116<br />

Chirapongsathorn, Sakkarin 281, 282,<br />

2089<br />

Cho, Eun Ju 242, 400, 447, 470, 474,<br />

479, 1610, 1907, 2016, 2042, 2055<br />

Cho, Eun-Suk 464<br />

Cho, Hyeki 242, 479, 2016<br />

Cho, Hyo Jung 379, 1567<br />

Cho, Hyosun 1001<br />

Cho, Jai Young 857, 1902<br />

Cho, Ju-Yeon 1415<br />

Cho, Junhyeon 389<br />

Cho, Kyung Joo 504<br />

Cho, Mee-Yon 1322<br />

Cho, Mong 876, 1542, 2028<br />

Cho, Soo Yeon 2056<br />

Cho, Sung Won 379, 1567, 2056<br />

Cho, Yong Kyun 924, 941, 951, 1624,<br />

1760, 1992, 2110, 2218, 2243<br />

Cho, Young Youn 242, 400, 447, 474,<br />

479, 1610, 1907, 2016<br />

Cho, Yuri 242, 400, 447, 474, 479,<br />

1610, 2016, 2042, 2055<br />

Chodavarapu, Krishna 713<br />

Chodavarapu, Ramakrishna K. 1168<br />

Choi, Angela O. 725<br />

Choi, Cheol Woong 876<br />

Choi, Dae Hee 1515<br />

Choi, Duck Joo 391, 1896<br />

Choi, Gina 122<br />

Choi, Hanlim 857<br />

Choi, Hwa Young 2113<br />

Choi, Jong Young 452, 1615, 1882<br />

Choi, Jonggi 127, 360, 381, 397<br />

Choi, Jung Eun 502, 1419<br />

Choi, Moon Seok 422, 423, 485,<br />

1591, 1960, 2049<br />

Choi, Myunghan 1043, 1109, 1145,<br />

1224, 1241, 1262<br />

Choi, Paul C. 238<br />

Choi, Sang Il 377<br />

Choi, Steve S. 80, 1142<br />

Choi, Won-Mook 1358, 1907<br />

Choi, Yong-Hun 403, 426, 749<br />

Choi, Yoo-Shin 1278<br />

Choi, Youkyung 2126<br />

Choi, YoungRok 857, 1902<br />

Chok, Kenneth S. 70, 1746, 2000<br />

Chokshi, Shilpa 163, 599, 2143<br />

Cholankeril, George 372<br />

Chong, Hui Heng 1686<br />

Chong, Yutian 120<br />

Chopda, Girish R. 2116<br />

Chopra, Kapil B. 607, 648, 1754<br />

Chotiyaputta, Watcharasak 467<br />

Chou, Hsin I 1364, 1392<br />

Chou, Wen-Wen 2264<br />

Choudhary, Manish C. 1005<br />

Choudhary, Narendra S. 2245<br />

Choudhury, Ashok K. 103, 1344, 2070,<br />

2078<br />

Chousterman, Michel 567<br />

Chow, Eric 1727, 1729<br />

Chowdhary, Vivek K. 99<br />

Chowdhury, Abhijit 750<br />

Christensen, Peer B. 792<br />

Christensen, Stefan 1052, 1060, 1081,<br />

1156, 1861<br />

Chromy, David 561, 1204<br />

Chu, Andrew 193, 197<br />

Chu, Chia-Ming 1552<br />

Chu, Po-sung 181, 569, 2131, 2140,<br />

2234<br />

Chu, Winnie C. 235<br />

Chua, Cui Wen 1812<br />

Chua, Mei-Sze 497, 500<br />

Chuang, Wan-Long 91, 241, 367, 398,<br />

459, 1080, 1105, 1120, 1168, 1828,<br />

1996, 2014, 2025, 2264<br />

Chuang, Ya-Hui 619<br />

Chuaypen, Natthaya 2036, 2063<br />

Chui, Celia B. 1832<br />

Chun, Hoon Jai 1504<br />

Chun, Youngjae 216<br />

Chung, Jung Wha 389<br />

Chung, Kyu Sik 346, 1542, 1558<br />

Chung, Randy 357<br />

Chung, Raymond T. 25, 154, 704, 778,<br />

991, 1015, 1094, 1630, 1644, 1768<br />

Chung, Sook In 504<br />

Chung, Sookja K. 1961<br />

Chung, Woo Jin 2038<br />

Chung, Young-Hwa 446, 448<br />

Churchill, Norma D. 1641<br />

Ciaccio, Antonio 571, 1453<br />

Ciancio, Alessia 1740<br />

Ciarleglio, Maria 743, 1731, 1899<br />

Cicalese, Luca 1991<br />

Ciccarese, Francesca 355<br />

Cillo, Umberto 148, 351, 356, 1270<br />

Cinar, Resat 1374<br />

Ciociaro, Demetrio 2209<br />

Cipriano, Madalena Z. 1921<br />

Cirincione, Brenda 720, 728<br />

Citti, Caitlin C. 380, 438, 457<br />

Civan, Jesse M. 10, 172, 1452, 1906<br />

Claassen, Mark 1007


1320A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Clahsen, Thomas 963<br />

Clair, Heather B. 666, 2233<br />

Clarett, Danisa M. 1257<br />

Clària, Joan 2087<br />

Claridge, Lee C. 2143<br />

Clark, Virginia C. 138, 839<br />

Clasen, Kristen M. 213<br />

Claudel, Thierry 977<br />

Clavien, Pierre A. 499, 665<br />

Cleary, Sean P. 350<br />

Clemens, Mark G. 1023<br />

Clément, Marc-André 51, 1520, 1523<br />

Cloherty, Gavin A. 1797<br />

Cloonan, Yona K. 1555<br />

Clouston, Andrew D. 784, 1362<br />

Coakley, Dion F. 1467, 1478, 1485<br />

Coakley, Eoin 36<br />

Cobanoglu, Murat C. 193<br />

Cobbold, Jeremy F. 2190<br />

Cocchio, Silvia 629<br />

Cocchis, Donatella 1<br />

Codrington, John F. 1846<br />

Coelho, Henrique Sergio M. 273, 289<br />

Coenraad, Minneke 2069<br />

Coffin, Carla S. 1576<br />

Cogliati, Bruno 586<br />

Cohen, Alice 1034<br />

Cohen, Ari J. 818, 1242<br />

Cohen, Ayelet 1235<br />

Cohen, Daniel E. 1039, 1088<br />

Cohen, David E. 322, 881, 900, 942<br />

Cohen, Eric 1039<br />

Cohen, Jose L. 215<br />

Cohen-Naftaly, Michal 875<br />

Cohn, Jennifer 1826<br />

Coilly, Audrey 3, 95, 1237, 1269,<br />

1782<br />

Cojuhari, Lilia 31<br />

Colan, Juan 2084<br />

Cole, Banumathi K. 664<br />

Colecchia, Antonio 741<br />

Coleman, Carla 1864, 1880<br />

Collado, Sol 175, 1916<br />

Colle, Isabelle 305<br />

Colledge, Danni 1557<br />

Collier, Jane 2190<br />

Collins, Allan J. 1867, 1883<br />

Collins, Christine 705<br />

Collins, Jenna 1050<br />

Collins, Peter 564<br />

Colmenero, Jordi 860<br />

Colnot, Nathalie 887<br />

Colombo, Bruno S. 2075<br />

Colombo, Massimo 229, 393, 1049,<br />

1086, 1743, 1747, 2011, 2012<br />

Colvin, Richard 2257, 2259, 2260,<br />

2261<br />

Combalia, Andres 615<br />

Comeglio, Paolo 882, 902<br />

Comerford, Megan 2166<br />

Compagnon, Philippe 215<br />

Compan, Anita 1154<br />

Concepcion, Waldo 1208<br />

Concepción, Mar 734, 1509<br />

Conde, Isabel 1230<br />

Conde de la Rosa, Laura 1382<br />

Cong, Min 187, 1404<br />

Congiu, Mario 996<br />

Congly, Stephen 1576<br />

Congly, Stephen E. 204<br />

Connelly, Margery A. 2227<br />

Conner, Elizabeth A. 677<br />

Conrads-Frank, Annette 1089<br />

Considine, Aisling B. 646, 1098, 1113,<br />

1744<br />

Conti, Filomena 30, 95, 997, 1239,<br />

1409<br />

Conti, Maria 309, 1153<br />

Contreras, Patricia C. 1315, 1522<br />

Conway, Brian 40, 249, 1161, 1198,<br />

1892<br />

Cook, Darrel 86<br />

Cooke, Graham 1826<br />

Coombes, Jason D. 163, 1978, 2143<br />

Cooper, Curtis 204, 249<br />

Cooper, David 1083<br />

Cooper, James N. 2173<br />

Cooper, Matthew A. 55<br />

Cooper, Stewart 1598<br />

Copeland, Neal 234<br />

Coppel, Ross L. 77, 613<br />

Coppleson, Yael J. 1454<br />

Coppola, Carmine 1802, 1844<br />

Coppola, Nicola 431, 1543, 1837<br />

Corbelli, Jody 435<br />

Corbett, Christopher 1321<br />

Corbin, Ian 361<br />

Cordero, Paula 1022, 1281<br />

Cordoba, Juan 1509<br />

Corey, Kathleen E. 154, 2166, 2194<br />

Cornberg, Markus 37, 252, 1058,<br />

1131, 1559, 1572, 1861, 2026<br />

Cornelissen, Jan 1803<br />

Cornella, Helena 253<br />

Cornella, Scott L. 1501<br />

Cornide-Petronio, Maria Eugenia 342<br />

Cornillie, Alexia 2079<br />

Corouge, Marion 1194<br />

Corpechot, Christophe 73, 634, 642<br />

Correia, Lurdes 473<br />

Correia, Pedro 473<br />

Correnti, Jason 1314<br />

Corrigan, Margaret 636<br />

Corsa, Amoreena C. 1678<br />

Corsello, Tiziana 1991<br />

Cortes, Andrea C. 2118<br />

Cortesi, Paolo 571, 1453<br />

Cortez-Pinto, Helena 150, 832, 897,<br />

940, 960<br />

Coseano, Hernan 1752<br />

Coskuner, Seher A. 1622<br />

Costa, Cilénia B. 150, 1876<br />

Costa, Fernando G. 2221<br />

Costa, José N. 473<br />

Costentin, Charlotte E. 132, 373, 417,<br />

421, 449, 1232<br />

Costes, Laurent 567<br />

Cotler, Scott 532, 559, 982, 1172,<br />

1248, 1854<br />

Cotoi, Corina G. 1733<br />

Cotrim, Helma P. 2237<br />

Cotte, Laurent 568, 1093<br />

Cotterell, Adrian 838<br />

Coughlan, Diarmuid 636<br />

Courcambeck, Jerome 394<br />

Cournand, Henri 1813<br />

Cousien, Anthony 1845<br />

Cousin, Vladimir 1728, 1735<br />

Couto, Claudia A. 2200<br />

Covington, Kyle 129<br />

Covinsky, Kenneth E. 841<br />

Cowie, Benjamin C. 1566, 2060<br />

Cox, Eleanor 1480, 1524<br />

Cox, I. Jane 1300<br />

Cox, Josiah 352<br />

Coyle, Catelyn 1810<br />

Coyne, Karin 1467, 1485<br />

Crabb, David W. 1332<br />

Craig, Kevin 2116<br />

Crawford, Darrell H. 818, 975<br />

Craxi, Antonio 1179, 1802, 1844<br />

Creighton, Chad 129<br />

Crespo, Javier 378, 1511, 2032, 2171<br />

Crespo Yanguas, Sara 586<br />

Crissien, Ana Maria 108<br />

Cronberg, Alexandra 113<br />

Crosbie, Orla M. 1103<br />

Crosby, Jeff R. 183, 1353<br />

Cross, Jennifer 2007<br />

Crothers, Michael W. 13<br />

Crouchet, Emilie 1019<br />

Cruz, Helder 1921<br />

Cruz, Pedro E. 1921<br />

Cruz, Ricardo 971<br />

Csak, Timea 512<br />

Csizmadia, Eva 1394<br />

Cubero, Francisco Javier 828, 1779,<br />

1924<br />

Cuervas-Mons, Valentin 1180<br />

Cuervo, Ana Maria 333<br />

Cui, Jeffrey Y. 45, 913, 1425, 1431<br />

Cui, Julia Yue 655<br />

Cui, Qingwen 1546, 1854<br />

Cui, Wei 85, 1031, 1785<br />

Cui, Yuanyuan 2107<br />

Culberson, Cathy 1389<br />

Cullaro, Giuseppe 272<br />

Culver, Emma L. 637<br />

Cummings, Oscar 160, 948<br />

Cummings, Yvette 1111, 1155, 1164<br />

Cunha, Aloisio S. 2200<br />

Cuperus, Frans 808<br />

Curry, Michael P. 704, 1045, 1046,<br />

1049, 1108, 1177<br />

Cursaro, Carmela 1888<br />

Curth, Harald M. 687<br />

Curti, Marco 337, 1388<br />

Curtis, Donna 1737<br />

Curtis, Timothy 71, 843<br />

Cusi, Kenneth 2147, 2208<br />

Cutilli, Carolyn 536<br />

Cutler, Murray J. 424<br />

Cuzin, Lise 568, 1093<br />

Cyr, Wendy 2116<br />

Czaja, Albert J. 303<br />

Czaja, Mark J. 1333<br />

Czul, Frank 1054, 2256<br />

D’Albuquerque, Luiz C. 469<br />

D’Aliberti, Deborah 1127<br />

d’Alteroche, Louis 3, 95, 145, 206<br />

D’Amico, Ronald 1067<br />

D’Souza, Rashmi 1733<br />

da Silva, Rita de Cássia M. 469<br />

da Silva, Tereza Cristina 586<br />

Da Silva, Hannah 2238, 2254<br />

da Veiga, Zulane 762, 2072, 2076<br />

Dadabhai, Alia S. 1749<br />

Daelken, Nicole 200<br />

Daffis, Stephane 35, 2009<br />

Dahari, Harel 982, 1854


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1321A<br />

Daher, Saleh N. 1878<br />

Dahlan, Yaser 1607<br />

Dai, Chia-Yen 398, 2264<br />

Dai, Erhei 209, 1587<br />

Dai, Feng 16<br />

Dai, Shenglong 1583<br />

Dai, Wing-chiu 70, 1746<br />

Daiber, Andreas 950<br />

Daikoku, Atsuko 517<br />

Daikos, Georgios L. 2094, 2095<br />

Daily, Michael F. 854<br />

Daita, Kalyani 905, 1309, 1527, 2249<br />

Daix, Valérie 974<br />

Dal Ben, Matteo 2098<br />

Dale, Cheryl 647<br />

Dalekos, George N. 73, 76, 2012<br />

Daley, Donnele 1282<br />

Dalgard, Olav 40<br />

Dalgic, Buket 1719<br />

Dalla Valle, Fabio 767<br />

Dalton, Harry 523<br />

Daltro, Carla 2237<br />

Daly, Ann K. 225, 1930, 1933<br />

Dametto, Ennia 1<br />

Damjanov, Nevena 521<br />

Dammermann, Werner 1629<br />

Dams, Els 1846<br />

Dan, Yunjie 1865<br />

Dandekar, Satya 63<br />

Danelski, Lisa 1839<br />

Danford, Christopher J. 2166<br />

Dannhorn, Emily 489, 490<br />

Danso, Hassan 987<br />

Dantas-Correa, Esther B. 2075<br />

Dao, Doan Y. 1555<br />

Dao, Thong 1072, 1187, 2082<br />

Dar, Sadaf 1005<br />

Dara, Lily 2135<br />

Darling, Jama M. 217, 1057<br />

Daruich, Jorge 438, 2106<br />

Das, Kausik 750<br />

Das, Kshaunish 750<br />

Das, Sukanta 581, 1344<br />

Das, Suvarthi 962, 2133, 2137<br />

Dasarathy, Jaividhya 2162, 2186<br />

Dasarathy, Srinivasan 330, 334, 518,<br />

651, 667, 668, 671, 1308, 1468,<br />

2162, 2182, 2186<br />

Dasgupta, Aditi 1825<br />

Dasgupta, Kathleen 1229<br />

Dash, Ajit 175, 664, 915, 1916<br />

Dash, Asha 1096<br />

Dash, Srikanta 501, 506, 1096<br />

Dash, Suchismita 1911<br />

DaSilvera, Eduardo 1549<br />

Dattaroy, Diptadip 962, 2133, 2137<br />

Daud, Amna 71, 563, 843, 847, 849<br />

Daugherty, Tami 1182, 1189<br />

Daulouede, Jean-Pierre 1796<br />

Davalos-Moscol, Milagros B. 595<br />

Davenport, Mark 134, 1691<br />

David, Lawrence 2188<br />

David, Paul 1663, 2088<br />

David, Waseem 358<br />

Davidson, Brian 1373<br />

Davidson, Nicholas O. 911, 1361,<br />

1365<br />

Davies, Nathan 1300<br />

Davies, Sean S. 1407<br />

Davies, Susan 258<br />

Davila, Jessica A. 357<br />

Davila, Marta L. 130, 682<br />

Davis, Brian C. 213, 1768<br />

Davis, Eric 1178<br />

Davis, Roger J. 1924<br />

Davuluri, Gangarao 330, 334, 518,<br />

651, 667, 668, 671, 1308<br />

Dawson, George 1797<br />

Dawson, Paul 19<br />

Day, Christopher P. 2164, 2183<br />

Day, Gretchen 626<br />

Day, James W. 776<br />

Daye, John 1716<br />

Dayoub, Rania 954<br />

de Andrade, Mariza 607, 610<br />

de Araujo Horcel, Lucas 57<br />

de Assuncao, Thiago 692, 1355, 1406<br />

de Avila, Leyla 2205<br />

de Boer, Bastiaan 2185<br />

de Boer, Ian H. 2228<br />

de Branco, Gabriel 1484, 1494, 1495<br />

De Carlis, Luciano 1751<br />

de Carvalho, Juliana R. 273, 289<br />

De Chiara, Francesco 908, 1517<br />

De Coppi, Paolo 337, 1388<br />

De Francesco, Raffaele 2011<br />

De Gaetano, Anna Maria 432<br />

de Galocsy, Chantal 1091<br />

De Gottardi, Andrea 2087<br />

De Gramont, Armand 417<br />

de Jong, Koert P. 1434<br />

De Jong, Ype P. 1841<br />

de Knegt, Robert J. 445, 492, 1007,<br />

1166, 1196, 1585, 1586, 1590,<br />

1998, 2002, 2003<br />

De La Cruz-Munoz, Nestor 914<br />

De la Flor-Robledo, María 1758<br />

de la Grange, Pierre 1288<br />

de la Motte, Carol A. 2132<br />

de la Peña, Joaquin 745<br />

de la Revilla, Juan 1511, 2012<br />

de la Rosa, Brauley 1870<br />

De La Rosa, Guy 1040<br />

de la Vega-Prieto, Maria Jose 1342<br />

de Ledinghen, Victor 3, 95, 117, 206,<br />

251, 450, 777, 1439, 1796, 1847,<br />

2181<br />

De Luca, Francesca 222<br />

de Magnée, Catherine 1734<br />

de Man, Robert A. 68, 1590, 1803<br />

De Man, Joris G. 976<br />

De Maria, Vincenzo 1802, 1844<br />

De Martin, Eleonora 487, 1782<br />

de Mendoza, Carmen 1180, 1881,<br />

1908<br />

de Mesquita, Fernanda C. 326, 1418<br />

De Minicis, Samuele 21<br />

De Nicola, Stella 1747<br />

de Niet, Annikki 1995, 2046<br />

de Oliveira, Jarbas R. 1418<br />

De Pasquale, Paola 1493<br />

de Stefano, Giorgio 431<br />

De Vito, Claudio 221<br />

de Vos, Marie 1216<br />

de Vos, Paul 1427<br />

de Vree, J. Marleen 208, 2263<br />

de Vries, Annemarie C. 68<br />

de Vries, Elisabeth M. 74, 604, 623<br />

de Vries, Marianne 1238<br />

de Vries, Niek 637<br />

de Waart, Dirk R. 604<br />

De Winter, Benedicte Y. 894, 976<br />

De Wolf, Andre M. 9<br />

De-Oertel, Shampa 1080, 1105, 1120<br />

Deal, Barbara J. 401<br />

Deans, Julie 1290<br />

Debes, Jose 572, 1752<br />

Debette-Gratien, Marie Line 95<br />

Debray, Dominique 1728<br />

Debzi, Nabil 1621<br />

Decaens, Thomas 421, 449<br />

Decaris, Martin 1425<br />

Deckmyn, Olivier 786<br />

Deemer, James 527<br />

Defreitas, Andrew 401<br />

Degallaix, Nathalie 974<br />

Degrassi, Cristina 2098<br />

Degré, Delphine 1091, 2079<br />

Deguchi, Yoshio 1941<br />

DeHart, David N. 319<br />

Deheragoda, Maesha 1711, 1712,<br />

1721<br />

Dehovitz, Jack 1444<br />

Deichsel, Danilo 245<br />

Dejong, Cornelis H. 818<br />

Dekhtyar, Tanya 705<br />

Del Bello, David P. 1202<br />

Del Rio Hernandez, Armando E. 337,<br />

1388<br />

dela Cruz, Rouchelle D. 2214<br />

Dela Cruz, Anna Christina 914, 1267,<br />

2081<br />

Delaney, William E. 35, 1594, 2009<br />

Delasalle, Patrick 1072<br />

Delaunay, Jean-Louis 196<br />

Delelo, Rolland 1409<br />

Delemos, Andrew 17, 2166<br />

Deleuran, Bent 301<br />

Delgado, Evan 493<br />

Delio, Maria 943<br />

Delitto, Anthony 1244<br />

Dell’Olio, Dominic 1<br />

Della Corte, Cristina 435<br />

Deltenre, Pierre 1818<br />

Delvart, Valérie 1237<br />

Delwaide, Jean 1091<br />

DeMasi, Ralph 1163<br />

Demetris, Anthony J. 216, 1736<br />

DeMino, Mary 1650<br />

Demir, Kadir 478<br />

Demir, Münevver 2195<br />

Demiroz, Ali Pekcan 2051<br />

Demma, Shirin 1904<br />

DeMorrow, Sharon 454, 621, 835,<br />

1507<br />

Demzik, Alysen 1235<br />

den Dulk, A. Claire 1238<br />

Denarie, Michel F. 1085<br />

Deneau, Mark 1702<br />

Deng, Yanghong 1731<br />

Deng, Yanhong 743, 1899<br />

Deng, Yong-Qiong 1564, 1602<br />

Denis, Jacques 2082<br />

Denning, Jill M. 746, 1049, 1128<br />

Deny, Paul 1162<br />

DePaoli, Alex M. 106, 198, 627, 1986<br />

DePeralta, Danielle K. 957<br />

Derbala, Moutaz F. 1147, 1188<br />

Derderian, S. C. 98<br />

Desai, Archita P. 1459


1322A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Desai, Payal 1801<br />

Desai, Sachin 2261<br />

DeSanto, Jennifer 1849<br />

Deshmukh, Sandeep 1906<br />

Desmond, Paul V. 442, 996<br />

Desnick, Robert J. 2111, 2112<br />

Deterding, Katja 1131, 2003<br />

Detlefsen, Sönke 780, 789, 1440<br />

Deuffic-Burban, Sylvie 1845<br />

Deutsch, Melanie 2019<br />

Deutz, Nicolaas E. 671, 1523<br />

Dev, Anouk T. 442, 483<br />

Devara, Anupama 420<br />

Devarajan, Karthik 371<br />

Devarbhavi, Harshad 284, 2101<br />

Dever, John 549<br />

Devereaux, Michael W. 1687, 1688,<br />

2134<br />

Devière, Jacques 178<br />

Devoto, Marcella 1693<br />

Devue, Cécile 887, 2148<br />

Devun, Daniel 1242<br />

DeWitt, Sheila H. 143<br />

Dhali, Gopal K. 750<br />

Dhaliwal, Sandeep S. 639<br />

Dhamija, Ravinder Mohan 147<br />

Dhanasekaran, Renumathy 129, 1965<br />

Dharancy, Sebastien 95, 263, 1269,<br />

1808, 2222<br />

Dhawan, Anil 134, 1691, 1733<br />

Dhersin, Jean-Stéphane 1845<br />

Dhillon, Amar P. 793, 1530<br />

Dhillon, Ammen P. 2007, 2034<br />

Dhillon, Sonu 774<br />

Dhiman, Radha K. 214<br />

Di Benedetto, Antonino 1476<br />

Di Benedetto, Fabrizio 1751<br />

Di Bisceglie, Adrian M. 94, 121, 1057,<br />

1478, 1598<br />

Di Carlo, Antonio 1122<br />

Di Cecco, Sara R. 836<br />

Di Cocco, Silvia 958<br />

Di Gregorio, Vincenza 2067<br />

Di Leo, Alfredo 1802, 1844<br />

Di Maio, Velia Chiara 222<br />

Di Maira, Giovanni 973<br />

Di Maira, Tommaso 1230, 1251<br />

Di Marco, Mariella 355<br />

Di Marco, Vito 973<br />

Di Martino, Vincent 3, 95, 145, 206,<br />

554, 1036, 1187, 2082<br />

Di Pace, Maria 595<br />

Di Paolo, Daniela 44<br />

Di Paolo, Daniele 203, 222<br />

Di Spirito, Mike 730<br />

Diago, Moisés 2032, 2171<br />

Diago, Teresa 1271<br />

Diaz, Giacomo 166, 1646<br />

Diaz, Maria Alba 860, 1889<br />

Diaz-Meco, Maria T. 232<br />

Dickerman, Richard 1169<br />

Dickmeyer, Laura J. 401<br />

Dickson, Rolland C. 580, 1329, 1831,<br />

2073<br />

Dickstein, Aaron 1592<br />

Dieguez, Gabriela 1069<br />

Diehl, Anna Mae 66, 80, 161, 737,<br />

889, 936, 962, 1386, 1435, 1715,<br />

2137, 2143, 2169, 2188<br />

Diener, Katrina 503<br />

Dieterich, Douglas 199, 380, 457, 545,<br />

726, 1046, 1065, 1079, 1090, 1095,<br />

1106, 1108, 1112, 1151, 1185,<br />

1202, 1538<br />

Dietz, Julia 980, 1559<br />

Difrancesco, Robin 332<br />

Dige, Anders 301<br />

Dik, Ebru 1622<br />

Dillon, Jayne 11<br />

Dillon, John F. 225, 1930<br />

Dimova, Rositsa B. 1823, 1841<br />

Dincses, Elif 2180<br />

Ding, Dora 616, 624<br />

Ding, Jia 366, 935<br />

Ding, Jiguang 2048<br />

Ding, Qian 1355<br />

Ding, Shitao 1865<br />

Ding, Wen-Xing 140, 598, 1325<br />

Ding, Xiwei 1966<br />

Ding, Yanhua 2058<br />

Ding, Yifeng 140<br />

Dinh, Phillip 241, 2021, 2025<br />

Dippold, Christine 197<br />

Disabato, Mara 215<br />

Dispenzieri, Angela 1263<br />

Divine, George 1254<br />

Diwanji, Rohan 510, 2116<br />

Do, Ailinh 2250<br />

Do, Albert 283<br />

Doan, Sylvia 1702<br />

Dodge, Jennifer L. 4, 50, 53, 462,<br />

1273, 1444<br />

Doehle, Brian 91, 205, 219, 1168<br />

Dohmen, Kazufumi 2031<br />

Doi, Hiroyoshi 1302, 1941<br />

Doi, Yoshinori 1201<br />

Doignon, Isabelle 816<br />

Dole, Kiran 2259, 2260<br />

Dolganiuc, Angela 839, 1286<br />

Dolgin, Natasha 845<br />

Dolin, Christine E. 587<br />

Dollé, Laurent 2143<br />

Dombrowicz, David 928<br />

Domenicali, Marco 1493<br />

Dominguez, Edward A. 1169<br />

Dominitz, Jason A. 128<br />

Donato, Francesco 1568<br />

Donato, Maria F. 1743, 1747, 1751<br />

Donchev, Vladamir 1479<br />

Donde, Hridgandh 1331<br />

Dondossola, Daniele E. 1743, 1751<br />

Donepudi, Ajay C. 885<br />

Dong, Chao 1526<br />

Dong, Tien S. 271<br />

Dong, Weiyan B. 1832<br />

Dongiovanni, Paola 973, 2187<br />

Dongre, Neelesh 260, 753<br />

Donner, Aaron J. 507<br />

Donohue, Julie M. 104, 151<br />

Donohue, Terrence M. 589, 1389<br />

Donovan, John 1350<br />

Donovan, Julie 36<br />

Doo, Edward 213, 218, 2150, 2220<br />

Doorenspleet, Marieke E. 637<br />

Dopazo, Hernán 2217<br />

Dore, Gregory 40, 91, 543, 1083,<br />

1161, 1868<br />

Doria, Cataldo 1276<br />

Dorko, Kenneth 1291<br />

Doshi, Dilesh 1189<br />

Doss, Wahid H. 708, 1076, 1163<br />

Dossier, Claire 196<br />

Dostal, David E. 830<br />

Dottin, Stephania 1452<br />

Dou, Xiaoguang 2048<br />

Dougherty, Michelle 1864, 1880<br />

Dourakis, Spyros P. 1489, 2096<br />

Douschan, Philipp 2068<br />

Doussau, Adelaide 132<br />

Dove, Lorna M. 531, 550<br />

Doyle, Adam 11<br />

Doyle, Erin H. 229, 988, 1289, 1742<br />

Dranoff, Jonathan A. 1384, 1387<br />

Dreher, Bernd 1170, 1200<br />

Drenth, Joost 105, 162, 293, 609,<br />

2102, 2145<br />

Drescher, Hannah K. 963<br />

Dreymueller, Daniela 100<br />

Drinane, Mary 1406<br />

Drobnik, Ann 1823<br />

Du, Huijuan 638<br />

Du, Sandie Y. 2007<br />

Du, Xiaofei 2061<br />

Duan, Na-Na 935<br />

Duan, Weijia 374<br />

Duan, Xiaoqiong 981<br />

Duan, Yuyou 685<br />

Duan, Zhong-Ping 209<br />

Duarte, António 897<br />

Duarte, Ignacio 2210<br />

Duarte, Sebastião M. 2221<br />

Duarte-Rojo, Andres 445, 1099, 1166,<br />

1307, 1471, 1481, 1857<br />

Duarte-Salles, Talita 170<br />

Dubay, Derek 1250<br />

Dubin, Sterling M. 529<br />

Duboc, Denis 1194<br />

Dubray, Clarisse 394<br />

Dubreuil, Marta 615<br />

DuBrock, Hilary M. 9<br />

Duca, Ana Maria 1180<br />

Ducancelle, Alexandra 1439<br />

Duchesne, Léa 1833<br />

Duclos-Vallee, Jean-Charles 3, 95, 1237,<br />

1782<br />

Ducom, Julie 529, 549<br />

Dudek, Hank 510, 2116<br />

Dueñas, Eva 378<br />

Dufour, Jean-Francois 221, 425, 445,<br />

1049, 1166, 1884, 2164<br />

Dugas, Lara R. 1248<br />

Dugum, Alex 855, 856<br />

Dugum, Mohannad 855, 856<br />

Duksal, Faysal 2219<br />

Dumitru, Raluca 1327<br />

Dumont, Laure 98<br />

Dumortier, Jérôme 3, 95, 1158, 2082<br />

Duncan, Andrew W. 493<br />

Duncan, Oliver J. 848<br />

Duncan, Richard G. 36<br />

Dundas, Pauline 1104<br />

Dunn, Michael A. 648, 1244, 1468,<br />

1754<br />

Dunn, Winston 1814, 1815<br />

Dunnington, Katherine M. 725<br />

Dupont, Benoît 179<br />

Duran, Domingos 1859<br />

Duran-Struuck, Raimon 2129<br />

Durand, Christine M. 1749


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1323A<br />

Durand, Francois 95, 265, 480, 772,<br />

887, 1288, 1781, 2148<br />

Durand-Schneider, Anne-Marie 196<br />

Duranyildiz, Derya 478<br />

Durchschein, Franziska 2068<br />

Durel, Ryan M. 1242<br />

Durkalski, Valerie L. 213, 1766<br />

Durkin, Marian E. 677<br />

Duron, Cédric 1505<br />

Duseja, Ajay K. 214<br />

Dusheiko, Geoffrey M. 712, 2059<br />

Dutta, Amal K. 815<br />

Dutta, Chaitali 510, 2116<br />

Dutta, Sandeep 710, 719, 723<br />

Dutton, N. 2233<br />

Duvivier, Claudine C. 568, 1093<br />

Duvoor, Chitharanjan 1307<br />

Duvoux, Christophe 3, 95, 421, 449,<br />

1232<br />

Dvorak, Karel 1487<br />

Dvory, Hadas S. 247<br />

Dvory-Sobol, Hadas 91, 219, 718,<br />

1168<br />

Dweik, Nazeeh Z. 1147, 1188<br />

Dybala, Claire 2175<br />

Dyson, Jessica K. 611<br />

Eason, James 1212, 1233<br />

Eaton, John E. 607<br />

Ebinuma, Hirotoshi 181, 569, 2131,<br />

2140, 2234<br />

Ebisutani, Yusuke 906<br />

Eccleston, Jason L. 1498<br />

Eckert, Christoph 1299, 1385<br />

Eckert, Doerthe 1066<br />

Eckmann, Lars 1319<br />

Eda, Keisuke 1724<br />

Edelman, Devorah 943<br />

Edelman, Elazer R. 1292<br />

Eder, Michael 297, 1070<br />

Edlin, Brian R. 1042<br />

Edula, Raja G. Reddy 295, 1181<br />

Edwards, Angela 1063<br />

Edwards, Jeffrey 1915<br />

Edwards, Michael 503<br />

Efe, Cumali 308<br />

Egawa, Mayumi 227, 1974<br />

Egert, E. M. 521<br />

Egetemeyr, Daniel P. 1183<br />

Eghtesad, Bijan 840, 850, 855, 856,<br />

1265, 1271<br />

Eguchi, Akiko 57, 112, 1371, 1522,<br />

1971<br />

Eguchi, Junichi 1302, 1941<br />

Eguchi, Yuichiro 1097, 1632, 1819,<br />

2202, 2207<br />

Ehman, Richard 1355<br />

Ehrinpreis, Murray N. 420<br />

Ehrlich, Adam C. 765<br />

Ehrlich, Laurent 830<br />

Ehsan, Nermine 1862, 1879, 1890<br />

Einar, Bjornsson S. 1926<br />

Eisner, Sigal 875<br />

Ekbom, Anders 116<br />

Ekong, Udeme D. 1731<br />

Eksteen, Bertus 76, 616, 2143<br />

El Aggan, Hoda 1011<br />

El Deeb, Nevine M. 1011<br />

El Demiry, Sally 1011<br />

El Kasmi, Karim C. 1687, 1688, 1692,<br />

2134<br />

El Raziky, Maissa 1163<br />

El Refaie, Ahmed 1890<br />

El Sabaawy, Maha 1890<br />

El-Bendary, Mahmoud 1236<br />

El-Fert, Ashraf Y. 733<br />

El-Gazzaz, Galal 850, 1271<br />

El-Matary, Wael 1702<br />

El-Saadany, Mohamed M. 758<br />

El-Serag, Hashem B. 90, 156, 244,<br />

357, 1048<br />

El-Sherif, Omar 1103<br />

Elam, April D. 1835<br />

Elashry, Amr 1853<br />

Elbadri, Mohammed E. 1147<br />

Elbanna, Abduh M. 2215<br />

Elbasha, Elamin 727<br />

Eley, Timothy 720, 728<br />

Elfeki, Mohamed 1236<br />

Elfeky, Sarah 534<br />

Elgilani, Faysal 2<br />

Elia, Chiara 266<br />

Elizaga, Jaime 1503<br />

Elkhammas, Elmahdi 874, 1260, 1280<br />

Elkhashab, Magdy 1051, 2004, 2025<br />

Elkrief, Laure 145, 264, 772<br />

Elliott, Michael 106, 627<br />

Elmazaly, Mohamed A. 1862, 1879<br />

Elorza, Ainara 930<br />

Elsabaawy, Maha M. 1862, 1879<br />

Elsayad, Elham 1147<br />

Elshamy, Ahmed M. 229<br />

Elsharkawy, Aisha 1163<br />

Elsheikh, Elzafir 316, 892, 1186<br />

Elzohry, Hasan A. 733, 1853<br />

Eminler, Ahmet T. 846<br />

Emmert-Buck, Michael R. 166<br />

Emond, Jean C. 369, 838, 1454<br />

Emson, Claire 1425<br />

Enders, Felicity B. 605<br />

Endig, Jessica 2128<br />

Endo, Mizuki 2267<br />

Endo, Ryujin 1528<br />

Endo, Takeshi 135<br />

Enejosa, Jeffrey 1065, 1084<br />

Enestvedt, C. Kristian 69, 481<br />

Eng, Francis J. 229<br />

Eng, Lauren M. 2214<br />

Eng, Rene S. 2214<br />

Engelen, Marielle 671<br />

Engelmann, Cornelius 261, 276<br />

Engelmann, Michael 1002<br />

Engle, Ronald E. 166, 1646<br />

English, Shirley 1104<br />

Enomoto, Hirayuki 466, 770, 1533,<br />

1669, 1975<br />

Enomoto, Masaru 429, 1125, 1824<br />

Enomoto, Nobuyuki 406, 1027, 1565,<br />

1673<br />

Enooku, Kenichiro 461, 912<br />

Enrich, Carlos 64<br />

Enriquez, Aimee 1478<br />

Ensor, Katherine 1991<br />

Enweluzo, Chijioke U. 748, 769<br />

Epstein, Sara 414<br />

Erario, Madeline 575<br />

Erdelyi, Katalin 1374<br />

Erensoy, Selda 1739<br />

Ergen, Can 1756<br />

Erhart, Wiebke 617<br />

Erne, Elke M. 1802, 1844<br />

Ersoy, Baran A. 942<br />

Ersoz, Galip 246, 1247, 1739<br />

Erwin, Patricia J. 282<br />

Escajadillo, Natali 734<br />

Escalante, Shannon M. 278<br />

Escheik, Carey 534<br />

Escobedo, Miguel 1281<br />

Eshelman, Anne 556<br />

Eskind, Lon 17<br />

Eslam, Mohammed 2185<br />

Eslick, Guy D. 1343, 1450<br />

Esmadi, Mohammad 748, 769<br />

Esmat, Gamal E. 708, 1076<br />

Español-Suñer, Regina 98<br />

Espera, Hannah G. 1798, 1807<br />

Espinosa-Cuevas, Angeles 1471<br />

Espiritu, Christine 33, 2064<br />

Esplugues, Juan V. 1913<br />

Esposito, Isabella 1180<br />

Esquivel, Carlos O. 341, 1208, 2123<br />

Esteban, Rafael 1021, 1674, 2012<br />

Estep, James M. 316<br />

Estep, Michael J. 956, 1186<br />

Estes, Chris R. 1859<br />

Estévez, Pamela 2171<br />

Estrabaud, Emilie 998, 1601<br />

Eswaran, Sheila L. 1215<br />

Etschmaier, Alexandra 1466<br />

Ettorre, Giuseppe M. 1751<br />

Etzion, Ohad 779, 1498, 1720, 2144,<br />

2224<br />

Eubanks, Haleigh B. 1384<br />

Eun, Hyuk-Soo 1358, 1414<br />

Eun, Jong Ryeol 685<br />

Evans, Alison 371<br />

Evans, Helen M. 1728<br />

Evelyn, Zoehrer 1726<br />

Everson, Gregory T. 630, 746, 1045,<br />

1049, 1065, 1130, 1849<br />

Evon, Donna M. 1063, 1114<br />

Ewing, Micheal 193, 197<br />

Ezaz, Ghideon 283<br />

Fabbrizzi, Alessio 1190<br />

Faber, Klaas Nico 591, 612, 1398,<br />

1416<br />

Fabre, Thomas 1380<br />

Fabrellas, Nuria 266<br />

Fabri, Milotka J. 2004<br />

Fabris, Luca 76, 571, 811, 1950<br />

Facchetti, Floriana 435, 2011, 2012<br />

Facciuto, Marcelo 870<br />

Factor, Valentina M. 677<br />

Fadel, William 1903<br />

Fadin, Mariangela 351<br />

Fagan, Kevin 784<br />

Fagiuoli, Stefano 571<br />

Fagoaga, Omar 1254<br />

Faisal, Nabiha 204<br />

Faivre, Sandrine 417<br />

Falck-Ytter, Yngve 1154<br />

Falcon-Perez, Juan M. 953<br />

Faletti, Riccardo 2209<br />

Falguières, Thomas 196<br />

Falissard, Bruno 373<br />

Falkner, Keith C. 666, 1303, 2142,<br />

2233<br />

Fallon, Michael B. 79, 82, 570, 743,<br />

1424, 1470, 1513, 2151<br />

Fallowfield, Jonathan A. 260, 753,<br />

2168


1324A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Faloppi, Luca 435<br />

Falzano, Loredana 1802, 1844<br />

Famoso, Giulia 268<br />

Fan, Daiming 468, 754<br />

Fan, Daping 2133<br />

Fan, Jia 366<br />

Fan, Yang-Yi 56<br />

Fang, Yun 2204<br />

Fang, Zhengfeng 194, 1689<br />

Fanning, Liam J. 1103<br />

Fante, Garrett 1111, 1164<br />

Fantini, Damiano 1988<br />

Farag, Raghda 1236<br />

Farah, Taha 1607<br />

Farci, Patrizia 166, 1646<br />

Farella, Nunzia 431<br />

Farges, Olivier 373<br />

Fargion, Silvia 393, 2187<br />

Farhang Zangneh, Hooman 350<br />

Farhat, Nada 2262<br />

Farhood, Anwar 586<br />

Farinati, Fabio 351, 355, 356<br />

Färkkilä, Martti A. 76, 633<br />

Farnan, Jeanne M. 544<br />

Farooqi, Javed I. 1872<br />

Farrell, Geoffrey C. 55, 230, 1909<br />

Farrokhi, Kaveh 1012, 1297<br />

Farrugia, Helen 442<br />

Farrugia, Maria 2251<br />

Farsad, Khashayar 481<br />

Fartoux, Laetitia 417<br />

Fasano, Massimo 2003<br />

Fasola, Carlos 1169<br />

Fasolato, Silvano 148, 266, 292<br />

Fasseu, Magali 1288<br />

Fateen, Waleed 791, 800<br />

Fatela, Narcisa 150<br />

Fattovich, Giovanna 1166, 1568, 1802,<br />

1844<br />

Fauler, Günter 1726<br />

Fausther, Michel 1384<br />

Fawcett, Jonathan 818<br />

Fazel, Yousef 2192, 2212<br />

Feaver, Ryan E. 915<br />

Fedchuk, Larysa 206<br />

Federman, Alex D. 1792<br />

Fehr, Jan 1818<br />

Fei, Ran 1850, 1877, 1973<br />

Feigh, Michael 978<br />

Feilen, Nicole A. 1389<br />

Feld, Jordan J. 348, 350, 445, 542,<br />

714, 1051, 1067, 1086, 1099, 1107,<br />

1166, 1556, 1797, 1998<br />

Feldbrügge, Linda 1222, 1394, 2227<br />

Felder, Martina M. 355<br />

Feldman, Amy G. 137, 1714, 1738<br />

Feldman, Brian M. 1713<br />

Feldner, Ana Cristina A. 1484, 1494,<br />

1495<br />

Feldstein, Ariel E. 55, 57, 112, 969,<br />

1371, 1522, 1971, 2185<br />

Feldstein, Vickie A. 1570<br />

Felga, Guilherme 469<br />

Felix, Sean C. 2204<br />

Félix, Jorge 1859<br />

Felizarta, Franco 41, 1065<br />

Feng, Bo 1436, 1850, 1877<br />

Feng, Chun 731<br />

Feng, Dan 1323, 1347<br />

Feng, Dechun 182, 2179<br />

Feng, Hwa-Ping 725, 730, 731<br />

Feng, Sandy 841, 852, 1736<br />

Feng, Wenke 114, 1305, 1306<br />

Feng, Ziding 371<br />

Feng, Zongdi 1703<br />

Fenkel, Jonathan M. 1452<br />

Fenlon, Laura 1455<br />

Feranchak, Andrew P. 815<br />

Feray, Cyrille 987, 1123<br />

Ferenci, Peter 37, 252, 297, 714, 732,<br />

1058, 1067, 1070, 1092, 1137, 1885<br />

Ferguson, Beatrice 2154<br />

Ferguson, James W. 631, 1216, 1234,<br />

1710<br />

Ferlitsch, Arnulf 149, 270, 475, 732,<br />

755, 771, 1466<br />

Ferlitsch, Monika 1466<br />

Fernandes, Flavia F. 84, 762, 2072,<br />

2076<br />

Fernandes, Prabhavathi 2246<br />

Fernandez, Alejandro 2084<br />

Fernandez, Javier 266, 2084<br />

Fernández, David 953, 2244<br />

Fernández, Natalia 1511<br />

Fernandez Mena, Carolina 1295, 1520<br />

Fernandez-Bermejo, Miguel 2171<br />

Fernández-Carrillo, Carlos 1511<br />

Fernandez-Checa, Jose 899, 1320<br />

Fernandez-Hernando, Ana 65<br />

Fernandez-Hernando, Carlos 65<br />

Fernandez-Montero, Jose V. 1881, 1908<br />

Fernández-Moreira, Daniel 896, 939<br />

Fernandez-Rodriguez, Camino Maria<br />

1341<br />

Fernández-Rodríguez, Conrado M. 2171<br />

Fernández-Varo, Guillermo 659<br />

Fernandez-Zapico, Martin E. 692<br />

Fernando, Samantha 527<br />

Ferolla, Silvia M. 2200<br />

Ferrante, Shannon A. 727<br />

Ferrarese, Alberto 268, 351, 1270,<br />

1889<br />

Ferrari, Carlo 1802, 1844, 2015, 2052<br />

Ferrari, Maria de Lourdes A. 2200<br />

Ferrari, Teresa C. 2200<br />

Ferrari-Lacraz, Sylvie 1735<br />

Ferraris, Pauline 501, 506, 1096<br />

Ferraz, Maria Lucia 48, 1484, 1494,<br />

1495<br />

Ferreira, Carlos R. 1722<br />

Ferreira, Paula 150<br />

Ferrell, Linda 1736<br />

Ferrer, Maria Teresa 378<br />

Ferri, Nicola 2187<br />

Fessel, W. Jeffrey 2015, 2052<br />

Festi, Davide 741<br />

Fevery, Bart 1040<br />

Fevery, Johan 76<br />

Fey, Holger 1294, 1370, 2124<br />

Feyssa, Eyob L. 530, 535<br />

Fialla, Annette D. 780, 789, 1440<br />

Fibbi, Meghan F. 1136<br />

Fickert, Peter 662, 2068<br />

Fiel, M. Isabel 29, 253, 254, 870, 988,<br />

1228, 1289, 1364, 1397, 1742, 2124<br />

Fierer, Daniel S. 29, 89, 1090, 1094,<br />

1112, 1852<br />

Fierro, Nora A 1855<br />

Figler, Robert 175, 1916<br />

Figorilli, Francesco 309, 1153, 1300,<br />

1769, 1780, 1781<br />

Figueiredo, Fatima F. 84, 2189<br />

Figueiredo, Sabrina R. 2237<br />

Figueruela, Blanca 2032<br />

Filippi, Sandra 882, 902<br />

Filomia, Roberto 288, 1127, 1476<br />

Fimmel, Claus J. 1791<br />

Fink, Michael A. 442, 483<br />

Finnemann, Claudia 1287<br />

Fiore, Francesco 222<br />

Fiorotto, Romina 811<br />

Firpi, Roberto J. 839<br />

Fischer, Hans-Peter 2115<br />

Fischer, Janett 1638<br />

Fischer, Josiane 2075<br />

Fischer, Sandra E. 11, 2184, 2238,<br />

2254<br />

Fischl, Margaret 1444<br />

Fisher, Jonathan S. 861, 872<br />

Fisher, Robert A. 71, 843, 847<br />

Fisher-Hoch, Susan 1424<br />

Fishman, Sarah 1875<br />

Fitzgerald, Katherine A. 180<br />

Fitzhugh, Courtney 1498<br />

Fitzmorris, Paul S. 1220, 1253<br />

Fitzpatrick, Emer 1711, 1733<br />

Fitzpatrick, Thomas 577<br />

Fizanne, Lionel 2188<br />

Flaherty, John F. 243, 1678, 2004,<br />

2059<br />

Flake, Sebastian 412<br />

Flamm, Steven L. 702, 739, 1045,<br />

1046, 1049, 1108, 1130, 1442<br />

Flanigan, Colleen 573<br />

Flavell, Richard A. 1284<br />

Fleckenstein, Jaquelyn 2091<br />

Fleming, Michael F. 1231<br />

Fletcher, Linda 784<br />

Fletcher, Simon P. 35, 2009<br />

Flisiak, Robert 241, 243, 1502, 2014<br />

Flood, Michael C. 790<br />

Floreani, Annarosa 73, 76, 299, 609,<br />

629, 634<br />

Flores, Lalo 33, 2064<br />

Flores Molina, Manuel 1380<br />

Flores-Toro, Joseph A. 177<br />

Florman, Sander S. 253, 870, 988,<br />

1228, 1289, 1397<br />

Fluker, Shelly-Ann 1835<br />

Fogacas, Homero S. 762<br />

Fognani, Elisa 1190<br />

Fok, Brandon 462<br />

Foldes, Emily 275<br />

Folino, F. 268<br />

Folkert, Ian W. 868<br />

Fombuena, Blanca 1863<br />

Fondevila, Constantino 860<br />

Font-Burgada, Joan 1971<br />

Fontaine, Hélène 206, 1194, 1808<br />

Fontana, Robert J. 213, 225, 706,<br />

1768, 1850, 1877, 1922, 1923,<br />

1930, 1931, 1932, 1933<br />

Fontanges, Thierry 1072<br />

Foote, Patrick H. 169<br />

Forbes, Kevin F. 531<br />

Forbes, Stuart J. 62, 678<br />

Ford, David A. 65<br />

Ford, Jo-Ann E. 647<br />

Ford, Joel S. 153


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1325A<br />

Ford, Mary 1801<br />

Forde, Kimberly A. 122, 536, 865,<br />

871, 1470<br />

Forne, Montserrat 378<br />

Forns, Xavier 700, 703, 746, 982,<br />

1045, 1049, 1084, 1086, 1130, 1889<br />

Forsythe, Jordan M. 1804<br />

Fortea, Jose Ignacio 1520<br />

Fosby, Bjarte 819<br />

Foschi, Francesco G. 355, 435<br />

Foskett, Pierre 1721<br />

Foss, Aksel 819<br />

Foster, Andrew L. 89, 1090<br />

Foster, Graham R. 167, 241, 249,<br />

1098, 1173, 1179, 1571, 1626, 2014<br />

Foti, Michelangelo 665<br />

Fouchard-Hubert, Isabelle 117, 206<br />

Foucher, Juliette 450, 1796, 2181<br />

Fougerou-Leurent, Claire 3<br />

Fourati, Slim 999<br />

Fournier, Claire 1997<br />

Fowler, Kathryn 2150<br />

Fox, Caroline S. 2191<br />

Fox, Rena K. 16<br />

Fracanzani, Anna Ludovica 393, 2187<br />

Fraeman, Kathy 1175<br />

Frampton, Gabriel A. 1507<br />

Franceschet, Irene 73, 299, 629<br />

Francescucci, Angelica 2160<br />

Franchitto, Antonio 621, 817, 835<br />

Francis, Fadi 1753<br />

Francis, Heather L. 190, 621, 812, 817,<br />

822, 835, 929, 1317, 1349, 1363<br />

Francis, Prashanth 1210<br />

Francis, Susan 1480, 1524<br />

Francisco Ziller, Nicki M. 836<br />

Francisco-Recuero, Irene 1028<br />

Franco, Edson S. 526, 1784<br />

Franco, Ricardo A. 1804<br />

Francoz, Claire 265, 772, 1781<br />

Francque, Sven M. 105, 162, 797,<br />

894, 928, 976, 1091, 2145<br />

Frank, Brombacher 183, 1353<br />

Frank, Wacker 444<br />

Franke, Olivia 626<br />

Frankel, Angela 1680<br />

Frankova, Sona 787<br />

Franques, François 450<br />

Fransen, Floris 1427<br />

Fransen, Paul 976<br />

Frantsen, Marijke 527<br />

Franzè, Maria Stella 288<br />

Fraser, Andrew 1104<br />

Frassineti, Giovanni Luca 435<br />

Frater, Yvonne 424<br />

Frazier, Lynn M. 94, 1057<br />

Frederic, Oberti 117<br />

Frederick, R Todd 538, 1467, 1485,<br />

2104<br />

Fredericks, Emily M. 133<br />

Fredrick, Linda 722<br />

Freedman, Daniel 36<br />

Freiberg, Matthew 1488<br />

Freise, Chris 838<br />

Freissmuth, Clarissa 297, 732, 1070,<br />

1092, 1137, 1885<br />

Freitas, Luiz Antônio R. 2237<br />

Frelin, Lars 992<br />

French, Audrey 1444<br />

French, Benjamin 853<br />

French, Dorothy 632, 1359, 1367,<br />

2149<br />

French, Pim 1637<br />

Frenette, Catherine T. 4, 7, 108, 861,<br />

872<br />

Frenguelli, Luca 337, 1388<br />

Frey, Christian 2009<br />

Freyer, Nora 1921<br />

Fried, Michael 46<br />

Fried, Michael W. 94, 217, 1034,<br />

1057, 1114, 1598, 1826<br />

Friedman, Harley 1329, 2073<br />

Friedman, Nir 672<br />

Friedman, Scott L. 29, 131, 1095,<br />

1185, 1337, 1364, 1392, 1405,<br />

1779, 2174<br />

Frigo, Anna Chiara 148<br />

Frissen, Mick 100<br />

Frode, Tania S. 2075<br />

Fröhlich-Reiterer, Elke 1726<br />

Frossard, Jean-Louis 2174<br />

Fryzek, Nancy 218<br />

Fu, Bo 1051<br />

Fu, Dong 1909<br />

Fu, Rongdang 353, 375<br />

Fu, Sherry 1850, 1877<br />

Fu, Yan-Ling 1616<br />

Fu, Yiming 1536<br />

Fu, Yu 328<br />

Fuchs, Bryan C. 957, 1938<br />

Fuchs, Michael 52, 1041, 1472<br />

Fuentes, Javier 378, 2032<br />

Fujie, Satomi 513, 1912, 1987<br />

Fujii, Hideki 1956, 2125, 2156, 2207<br />

Fujii, Hironobu 906, 2156<br />

Fujii, Yohei 1059<br />

Fujikawa, Kazuyuki 1679<br />

Fujiki, Masato 1271<br />

Fujimori, Shunji 291<br />

Fujimoto, Go 707<br />

Fujimoto, Jiro 188, 466, 513, 781,<br />

1420, 1987<br />

Fujimoto, Toyoshi 880<br />

Fujinaga, Hidetaka 912<br />

Fujino, Hatsue 986, 1056, 1649, 1681,<br />

1944<br />

Fujisaka, Yasuyuki 1026, 1648, 1943,<br />

2167<br />

Fujisawa, Koichi 329<br />

Fujisawa, Toshio 505<br />

Fujitani, Naoki 1375<br />

Fujiwara, Kei 1954<br />

Fujiwara, Naoto 461<br />

Fujiya, Mikihiro 927<br />

Fujiyama, Kazuhito 1642<br />

Fujiyama, Shunichiro 1191, 2020<br />

Fukazawa, Kyota 878<br />

Fukubayashi, Kotaro 1987<br />

Fukuhara, Takayuki 472, 1056, 1893<br />

Fukui, Hiroyuki 1201, 2213<br />

Fukuizumi, Kunitaka 304, 1159<br />

Fukumitsu, Ken 1976<br />

Fukumura, Yuki 820<br />

Fukunishi, Shinya 966<br />

Fukushima, Hirofumi 1771<br />

Fukushima, Kentaro 869<br />

Fukutomi, Keisuke 1584<br />

Funaki, Masaya 495, 994, 1000<br />

Fung, James 70, 1545, 1550, 1746,<br />

2000<br />

Fung, Phoenix 295<br />

Fung, Scott 241, 1678, 2004, 2014,<br />

2025, 2184, 2238, 2254<br />

Furlage, Rosemary 332<br />

Furlini, Giuliano 1888<br />

Fürst, Dieter O. 80<br />

Furuichi, Yoshihiro 1047<br />

Furuta, Kunimaro 227, 453, 1974<br />

Furuya, Shinji 1956, 2125<br />

Fusai, Giuseppe 1407<br />

Fusco, Dahlene 991, 1015, 1630,<br />

1644<br />

Fushimi, Kazumi 392<br />

Fuster, Josep 860<br />

G. Lavoie, Elise 1387<br />

Gaarde, William A. 2135<br />

Gabr, Reham Mohamed 1821<br />

Gacusan Munson, Marie Lou 1920,<br />

1929<br />

Gaeta, Giovanni B. 241, 1802, 1844,<br />

2014<br />

Gaetano, John N. 553<br />

Gaggar, Anuj 243, 1551, 1557, 1563,<br />

1594, 1651, 1668, 2014, 2015,<br />

2037, 2043, 2052, 2059<br />

Gaggini, Melania 2209<br />

Gaglio, Paul J. 1141<br />

Gagnieu, Marie-Claude 1158, 1162<br />

Gahl, William A. 1722<br />

Gaidosh, Gabriel S. 914<br />

Gailer, Ruth E. 1455<br />

Gaisa, Michael 89<br />

Galanis, Kostas 2012<br />

Galati, Joseph S. 1057<br />

Galbraith, James W. 1804<br />

Gale, Michael 28<br />

Galen, Kelly 1266<br />

Galil, Karin 36<br />

Gallay, Philippe 2035<br />

Galle, Peter R. 376, 1949<br />

Gallego, Adolfo 378<br />

Gallego-Durán, Rocío 2171<br />

Galli, Andrea 949, 1963<br />

Galli, Claudio 1888<br />

Galli, Silvia 1888<br />

Galloway, Christopher 2247<br />

Galmozzi, Enrico 2011<br />

Gama-Carvalho, Margarida 960<br />

Gambato, Martina 982, 1889<br />

Gambino, Carmine G. 288<br />

Gambino, Roberto 2209<br />

Gamil, Mohamed 1163<br />

Gamil, Mostafa 1076<br />

Gan-Schreier, Hongying 1411<br />

Gandhi, Yash 720, 728<br />

Gane, Edward J. 38, 40, 219, 243,<br />

249, 713, 746, 1045, 1049, 1083,<br />

1120, 1128, 1130, 1557, 1563,<br />

1651, 1678, 2004, 2015, 2052,<br />

2059, 2225<br />

Ganesan, Murali 589<br />

Ganesh, Shanthi 2116<br />

Ganger, Daniel 211, 401, 1759, 1761<br />

Ganguli, Anirban 871<br />

Ganjoo, Naveen 2170<br />

Ganne, Nathalie 1847<br />

Ganne-Carrié, Nathalie 421, 449, 2082<br />

Ganz, Michael 1456, 1461<br />

Gao, Bin 110, 115, 182, 334, 903,<br />

1580, 2179


1326A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Gao, Bing 1080, 1105, 1120<br />

Gao, Guangping 496<br />

Gao, Rui 698<br />

Gao, Xiuzhu 1149, 1603, 1657, 1869<br />

Gao, Yanhang 182, 1580, 1620<br />

Gao, Zhiliang 120, 1423, 1772, 2048<br />

Gaouar, Farid 642<br />

Garabedian, Elizabeth 1717<br />

Garas, George 848, 2099<br />

Garassini, Miguel E. 595<br />

Garcia, Gabriel 1182, 1189<br />

Garcia, José H. 469<br />

Garcia, Maria 1251<br />

Garcia, Mina A. 1874<br />

Garcia, Ruel 1549<br />

García, Elisabet 2071<br />

Garcia-Alonso, Veronica 2087<br />

Garcia-Buitrago, Monica 914<br />

Garcia-Caldero, Hector 1518<br />

García-Cortés, Miren 595<br />

Garcia-Martinez, Irma 979, 1284, 2179<br />

García-Monzón, Carmelo 930, 2171<br />

Garcia-Muñoz, B. 595<br />

Garcia-Pagan, Juan Carlos 326, 745,<br />

1518<br />

Garcia-Ruiz, Carmen 899, 1320<br />

García-Ruiz, Inmaculada 896, 939<br />

Garcia-Saenz de Sicilia, Mauricio 1307<br />

García-Samaniego, Javier 1028, 2171<br />

García-Torres, María Luisa 2171<br />

Garcia-Tsao, Guadalupe 82, 570, 743,<br />

745<br />

García-Valdecasas, Juan Carlos 860<br />

Gardenier, Donald 1151, 1185, 1870<br />

Garimella, Tushar 720, 728<br />

Garioud, Armand 554, 2082<br />

Garner, Jessica S. 929, 1317, 1349<br />

Garnier, Muriel 417<br />

Garre, Carmen 378<br />

Gartner, Michael 725, 730, 731<br />

Garvin, Jennifer H. 529<br />

Gasbarrini, Antonio 355, 432, 1802,<br />

1844<br />

Gasmi, Imène 693<br />

Gaspar, Rui 1540<br />

Gassler, Nikolaus 100, 828, 1924<br />

Gastaldelli, Amalia 2209<br />

Gatselis, Nikolaos 2012<br />

Gatta, Angelo 148<br />

Gauci, James Mario 2251<br />

Gaudio, Eugenio 621, 817, 834, 835<br />

Gault, Nathalie 145<br />

Gautam, Dheeraj 2245<br />

Gautam, Manjushree 1146, 1199<br />

Gautherot, Julien 196<br />

Gauthier, Aline 1176, 1205<br />

Gautsch, Sebastian 80<br />

Gavasso, Sabrina 351<br />

Gavis, Edith A. 52, 1463, 1491, 1527<br />

Gawrieh, Samer 436, 1903, 2166<br />

Gayed, Yasser 1275<br />

Gazzin, Silvia 2098<br />

Ge, Dongliang 1651, 1668, 2037<br />

Ge, Fengxia 972<br />

Ge, Jingjing 1396<br />

Ge, Xiadong 1382, 2124, 1940<br />

Gedaly, Roberto 673, 854, 1267, 2081<br />

Geerts, Anja M. 1091<br />

Geevarghese, Sunil K. 1223<br />

Geffers, Robert 1659, 2128<br />

Gehrke, Nadine 1949<br />

Geier, Andreas 37, 46, 2195<br />

Geiser, Mary 78<br />

Geldenhuys, Werner J. 1338<br />

Gelli, Maximiliano 487<br />

Gelman, Ekaterina 1597<br />

Gelson, William 258<br />

Gely, Cristina 1509<br />

Genco, Petrina V. 1236<br />

Genda, Takuya 1851<br />

Genescà, Joan 742, 1509<br />

Genovese, Domenico 1153<br />

Genovese, Federica 1437<br />

George, Jacob 119, 441, 1077, 1868,<br />

2185<br />

Gerada, Jurgen 2251<br />

Gerada, Martina 2251<br />

Geraghty, Estella M. 630<br />

Geratikornsupuk, Nopavut 360<br />

Gerber, Naomi Lynn 316<br />

Gerken, Guido 1779<br />

German, Polina 1133, 1707<br />

Germani, Giacomo 1270, 1889<br />

Gershwin, M. Eric 77, 602, 606, 613,<br />

619, 621, 639, 640, 1285<br />

Gerstoft, Jan 251<br />

Geskus, Ronald 74, 623<br />

Gevers, Tom J. 293<br />

Geyer, Peter R. 1170, 1200<br />

Ghabril, Marwan 1467, 1478, 1903,<br />

1931, 1932<br />

Ghali, Peter 204<br />

Ghalib, Reem H. 715, 739, 1040,<br />

2239<br />

Ghanekar, Anand 11<br />

Ghanta, Mythili 1122<br />

Ghany, Marc G. 218, 1598, 1722<br />

Ghare, Smita 1331<br />

Gharib, Ahmed M. 1498<br />

Gharibian, Derenik 1135<br />

Ghesquiere, Wayne 1161<br />

Ghobrial, R. Mark 1215<br />

Ghosh, Dipu 269<br />

Ghosh, Martha S. 521<br />

Ghosh, Siddhartha 1521<br />

Ghoshal, Kalpana 99, 493<br />

Ghoz, Hassan M. 169, 397<br />

Gi, Young Jin 130, 682<br />

Giabicani, Mikhael 1288<br />

Giama, Nasra H. 127, 169, 1965<br />

Gianella, Sara 1852<br />

Giannakeas, Nikos 801<br />

Giannelli, Valerio 265, 2067<br />

Giannini, Edoardo G. 355<br />

Giannone, Ferdinando A. 1493<br />

Gianserra, Laura 222<br />

Giard, Jeanne-Marie 1259<br />

Gibert, Cynthia 1488<br />

Gibson, Neil 208, 2263<br />

Gijbels, Marion J. 922<br />

Gil, Ana Isabel 1028<br />

Gil, Gregorio 804<br />

Gilbert, Melissa 1693<br />

Gill, Kirat 1111<br />

Gill, Ryan M. 157, 236, 2161<br />

Gill, Upkar S. 167, 1571, 1626<br />

Gillberg, Per-Göran 810<br />

Gilles, HoChong 52<br />

Gillespie, Brenda W. 838<br />

Gillevet, Patrick M. 905, 1463, 1527<br />

Gillies, Robert D. 29<br />

Gilmartin, Jocelyn 730<br />

Gilmour, Kimberly 1723<br />

Gilroy, Richard 565, 837, 1814, 1815<br />

Gines, Pere 266, 860, 1327, 2069,<br />

2087<br />

Giordani, Maria Teresa 1803<br />

Giostra, Emiliano 2174<br />

Giovanni, Shewit 1229<br />

Girala, Marcos A. 595<br />

Girard, Pierre-Marie 1676, 1677, 1833<br />

Girardin, Francois 1539<br />

Girgrah, Nigel 1242<br />

Girish, Veeranna 284<br />

Gish, Robert 32, 237, 1111, 1155,<br />

1164, 1549, 2022<br />

Giugliano, Silvia 1430, 1706, 1708<br />

Giuily, Nathalie 2044, 2045<br />

Giuliante, Felice 432<br />

Giusto, Michela 2067<br />

Gjuka, Donjeta 2151<br />

Glaser, Shannon S. 190, 621, 817,<br />

830, 834, 835, 929, 1317, 1349,<br />

1978<br />

Glässner, Andreas 1287<br />

Glebe, Dieter 2026<br />

Gloria, Helena 1256<br />

Glorioso, Jaime M. 2<br />

Go, Kristina L. 177<br />

Gobejishvili, Leila 1312, 1334<br />

Goblet, David 731<br />

Godlewski, Grzegorz 1374<br />

Goel, Amit 1475, 1817<br />

Goel, Aparna 545, 1337<br />

Goel, Satyender 563, 849<br />

Goeser, Felix 1287<br />

Goeser, Tobias 687<br />

Goetz, Matthew B. 1488<br />

Goetze, Oliver 46<br />

Goetzmann, Jason E. 32, 2022<br />

Gogela, Neliswa A. 1667<br />

Gogoi, Naminita 1344<br />

Gogos, Charalambos 2094, 2095<br />

Goh, Boon-Bee George 451<br />

Goh, Evan 47<br />

Gokturk, Huseyin Savas 2206<br />

Gola, Anna 1455<br />

Gola, Elisabetta 148, 292<br />

Golabi, Pegah 153, 1186, 2173, 2192<br />

Goldberg, David J. 1794<br />

Goldberg, David S. 5, 9, 427, 563,<br />

837, 849, 853, 864, 865, 868, 1218,<br />

1470, 2083<br />

Goldberg-Clein, Tova 1878<br />

Golden, Aaron 515<br />

Golden-Mason, Lucy 28, 503, 938,<br />

1430, 1706, 1708<br />

Goldenberg, Anna 1713<br />

Goldkamp, William J. 1212<br />

Goldstein, David B. 1651, 2037<br />

Goldstein, Eric 556<br />

Golin, Carol E. 1114<br />

Gomaa, Asmaa 344, 1163, 1853<br />

Gomberoff, Leon 1796<br />

Gomel, Rachel 630<br />

Gomes, Marilia B. 2189<br />

Gomes-Gouvêa, Michele S. 48<br />

Gómez Rodríguez, Rafael 2032<br />

Gomez-Camarero, Judith 2171<br />

Gomez-Moreno, E.M. 583


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1327A<br />

Gonçalves, Afonso 150<br />

Gonçalves, Nilza 1876<br />

Gong, Guo-Zhong 2023<br />

Gong, Jiao 120<br />

Gong, Li 1526<br />

Gong, Yuewen 361<br />

Gonzales, Gabriel R. 1146, 1199<br />

Gonzalez, Adam 774<br />

Gonzalez, Adriano m. 1484, 1494,<br />

1495<br />

Gonzalez, Frank J. 110<br />

Gonzalez, Patricia 982<br />

Gonzalez, Stevan A. 1146, 1199<br />

González, Javier M. 745<br />

González Ballerga, Esteban 2106<br />

Gonzalez-Aldaco, Karina 1855<br />

González-Jiménez, Andrés 595, 596<br />

Gonzalez-Reimers, Emilio 1341, 1342<br />

Gonzalez-Rodriguez, Agueda 930<br />

Good, Steven S. 2266<br />

Goodman, Zachary D. 624, 632, 737,<br />

739, 891, 892, 956, 1359, 1428,<br />

1435, 2149, 2173, 2192<br />

Goodrich, Nathan P. 1694<br />

Goossens, Nicolas 364, 1539, 1742,<br />

2174<br />

Gooz, Monika 319<br />

Gordon, Fiona H. 564<br />

Gordon, Fredric D. 412, 1106<br />

Gordon, Ronald E. 1364<br />

Gordon, Stuart C. 15, 106, 249, 525,<br />

627, 722, 1079, 1128, 1538, 1578,<br />

1789, 1800, 2015, 2052<br />

Goree, Jessica R. 1384, 1387<br />

Gores, Gregory J. 127, 381, 415, 825,<br />

895, 1124, 1315, 1994<br />

Gorgievski, Meri 1575<br />

Goria, Odile 1187<br />

Gorn, Chad 1452<br />

Gorsh, Boris 1456, 1461<br />

Gosselin, Nathalie H. 1915<br />

Gossmann, Mona 272<br />

Gotlieb, Neta 1685<br />

Goto, Fumio 164, 406<br />

Goto, Hidemi 411, 443, 880, 967,<br />

1061, 1167<br />

Goto, Kaku 505<br />

Goto, Takaaki 1573, 1599, 1618<br />

Gottfried, Michelle 211, 1759, 1766,<br />

1777<br />

Gottfriedova, Halima 787<br />

Gotthardt, Daniel 76, 618<br />

Gottwald, Mildred D. 2247<br />

Gou, Eric 2100, 2111, 2112<br />

Goukos, Dimitrios 2094, 2095<br />

Gould, John 1969, 2139<br />

Goulis, Ioannis 1126, 2012<br />

Goumard, Claire 1239, 1409<br />

Gounder, Prabhu P. 125, 1790, 1798,<br />

1807<br />

Gourari, Samir 1621<br />

Gournay, Jerôme 2082<br />

Goutte, Nathalie 373<br />

Gouw, Annette S. 1728<br />

Gove, James E. 1790, 1798, 1807<br />

Govil, Sanjay 867<br />

Gow, Paul 442, 483, 842, 1077, 1225<br />

Goyal, Vik 532, 559<br />

Gozgit, Joseph M. 1966<br />

Grabmeier-Pfistershammer, Katharina<br />

561, 1204<br />

Gracia-Sancho, Jordi 8, 326, 342,<br />

1418, 1518<br />

Graf, Rolf 499, 665<br />

Graffner, Hans 810<br />

Gragnani, Laura 1190<br />

Graham, Camilla S. 1826<br />

Graham, Emily 648<br />

Gralla, Jane 2085<br />

Grammatikopoulos, Tassos 134, 1691,<br />

1721<br />

Granato, Celso 48<br />

Grande, Lourdes 2171<br />

Grange, Jean Didier 1158<br />

Grangé, Jean-Didier 2082<br />

Grant, Charlotte R. 300, 608<br />

Grant, Claire 1480<br />

Grant, David 11, 53, 838<br />

Grant, Stephanie 1507<br />

Grant, Tiffannia 785<br />

Granzow, Michaela 80<br />

Grasset, Denis 264<br />

Graupera, Isabel 266<br />

Graw, Frederik 982<br />

Graziadei, Ivo 1092, 1885, 2087<br />

Grbec, Matjaz 292, 767<br />

Greaves, Wayne 727<br />

Grebely, Jason 40, 1083, 1868<br />

Greco, Stephanie 1282<br />

Green, Kile J. 611<br />

Green, Pamela 128<br />

Green, Richard M. 658<br />

Greenblatt, Ruth 1444, 2252<br />

Greene, Laurence 1141<br />

Greene, Tom 2112<br />

Greenhalgh, Stephen 62<br />

Greenstein, Andrew E. 1367<br />

Greenup, Astrid-Jane 1243<br />

Greenwood, Jeremy R. 957, 1938<br />

Gregori, Josep 1021, 1674<br />

Greig, Paul D. 11, 53<br />

Greinert, Robin A. 1519<br />

Greisen, Stinne 301<br />

Grenier, Carole 889<br />

Gress, Jacqueline 210<br />

Grewal, Priya 1095, 1185<br />

Grewal, Thomas 64<br />

Grieco, Antonio 432<br />

Grieco, Stefania 983<br />

Griesemer, Adam D. 2129<br />

Griffin, Sean 1243<br />

Griffiths, Jenifer 521<br />

Grigorieva, Julia 1900<br />

Grimm, Andrew A. 98<br />

Grimm, Dirk 98<br />

Grint, Paul 208, 2263<br />

Gripshover, Janet 1101<br />

Grochowski, Christopher 1693<br />

Groen, Albert 604<br />

Groessl, Erik J. 529, 549<br />

Grogan, Martha 1263<br />

Gronbaek, Henning 119, 301, 2071,<br />

2087<br />

Groothuis, Geny M. 1427<br />

Grossmann, Maria 1594<br />

Grossmann, Mathis 842<br />

Grove, Jane 225, 908<br />

Grønbæk, Henning 768<br />

Grundhoff, Adam 1629<br />

Grünhage, Frank 1429<br />

Gruss, Hans J. 2169<br />

Gschwantler, Michael 37, 252, 1058,<br />

1179, 1885, 1886<br />

Gu, Daehoe 1438<br />

Gu, Jiezhun 597, 1922, 1926, 1931<br />

Gu, Yurong 1772<br />

Gualano, Gisela 595<br />

Gualdieri, Luciano 1543<br />

Guan, Bo-Jhih 668<br />

Guañabens, Nuria 615<br />

Guardascione, Maria Anna 742<br />

Guarner, Carlos 734<br />

Guarner-Argente, Carlos 2084<br />

Guarneri, Valeria 1888<br />

Guarrera, James V. 71, 843, 847<br />

Guasti, Daniele 902<br />

Gubertini, Guido A. 222<br />

Guerin, Coralie L. 887<br />

Guerra, Bernadette 32, 2022<br />

Guerra, Giselle 1148<br />

Guerra, Juan Francisco 1217<br />

Guerra, Vito 437<br />

Guerrier, Micheleine 190<br />

Guerrieri, Francesca 165<br />

Guessab, Nawal 1621<br />

Guest, Rachel V. 678<br />

Guevara, Asdrubal 1281<br />

Gufraind, Alexander 1854<br />

Guglielmo, Flavius 1276, 1906<br />

Guha, Indra Neil 624, 1524, 2168<br />

Guicciardi, Maria Eugenia 825<br />

Guillot, Adrien 693<br />

Guimarães, Raquel a. 84<br />

Guixé-Muntet, Sergi 326, 1418<br />

Gulamhusein, Aliya 600, 610<br />

Gulati, Rishabh 1827<br />

Guldiken, Nurdan 831<br />

Gundogdu, Ceren 932<br />

Gunneson, Tim 293<br />

Gunsar, Fulya 246, 308, 1247, 1739<br />

Guo, Grace L. 919, 934<br />

Guo, Haifeng 1867, 1883<br />

Guo, Jinsheng 131, 1405<br />

Guo, Junli 1977<br />

Guo, Luyang 1947<br />

Guo, Qing 1611, 2002<br />

Guo, Simin 2002<br />

Guo, Wei 1660<br />

Guo, Wengang 754<br />

Guo, Xiao-Lin 1465<br />

Guo, Xin 1651<br />

Guo, Zifang 725<br />

Gupta, Ekta 1005, 1268<br />

Gupta, Nidhi 223<br />

Gupta, Nitika A. 12<br />

Gupta, Parul 1664<br />

Gupta, Priya 1917<br />

Gupta, Sanjeev 340, 515, 1764, 1917<br />

Gupta, Soumi 1189<br />

Gupta, Subhash 853<br />

Gupta, Tarana 214<br />

Gupta, Tripti 1242<br />

Gurakar, Ahmet 1214, 1255, 1749<br />

Gurel, Selim 2059<br />

Gustot, Thierry 178, 1091, 2079<br />

Guthke, Reinhard 954<br />

Guthrie, Deanne 1804<br />

Guthrie, Gregory J. 194, 1689<br />

Gutic, Enisa 1886


1328A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Gutierrez, Julio A. 39, 247, 1433,<br />

1900<br />

Gutkind, J. Silvio 20<br />

Guy, Cynthia D. 157, 161, 2169, 2188<br />

Guy, Jennifer 4, 462<br />

Guyader, Dominique 206, 1808, 1847<br />

Guyer, William 1041<br />

Guzman, Carlos A. 1762<br />

Guzman, Grace 1266, 1940<br />

Guzzardi, Maria Angela 672<br />

Gwak, Geum-Youn 422, 423, 485,<br />

1591, 1957, 1960, 2049<br />

Gydesen, Sofie 2178<br />

Gyongyosi, Benedek 496, 1318, 1330<br />

Ha, Yeonjung 446, 448<br />

Haagmans, Bart L. 2122<br />

Haber, Barbara A. 42, 133, 701<br />

Habersetzer, François 1158<br />

Habib, Adil 1169<br />

Habib, Aida 1391<br />

Habib, Nagy 118<br />

Habib, Naomi 672<br />

Habib, Robert 118<br />

Habic (Suciu), Alina 741<br />

Habra, Reema 556<br />

Hadengue, Alexandra 417<br />

Hadi, Salah 208, 2263<br />

Hadj-Nacer, Khaled 554<br />

Hadzic, Nedim 1711, 1712, 1723,<br />

2105<br />

Hadziyannis, Emilia 2019<br />

Hafler, David 1731<br />

Haga, Hiroaki 335, 358<br />

Haga, Yuki 1207, 1628<br />

Hagan, Michael 528<br />

Hagège, Hervé 2082<br />

Hagen, Susan J. 322<br />

Hagey, Lee R. 23, 829<br />

Hagihara, Atsushi 1125, 1824<br />

Hagiwara, Hideki 1201<br />

Hagström, Hannes 158, 308<br />

Hahn, Judith A. 1793<br />

Hahn, Katherine 286<br />

Hai, Hoang 191, 1125, 1824<br />

Hai, Seikan 466, 1420<br />

Haider, Samran 420<br />

Haifer, Craig 1225<br />

Haigh, W. G. 230, 923<br />

Hainaut, Pierre 170<br />

Hairwadzi, Henry N. 1667<br />

Haiying, Sun 2261<br />

Hajarizadeh, Behzad 1868<br />

Hajelssedig, Omer E. 1147<br />

Hajifathalian, Kaveh 1265, 1894<br />

Hakobyan, Syune 1891<br />

Halazun, Karim J. 369<br />

Halbower, Ann C. 1715<br />

Hale, Pamela 193, 197<br />

Halegoua-De Marzio, Dina 1261, 1276,<br />

1452<br />

Halff, Glenn A. 1900<br />

Halfon, Philippe 394, 1009<br />

Halilbasic, Emina 76<br />

Hall, Andrew R. 337, 793, 1388, 1530<br />

Hall, Coleen 708<br />

Hall, Russell 2169<br />

Hallal, Hacibe 583<br />

Hallberg, Par 225<br />

Haller, Wolfram 1728<br />

Halliday, John 622<br />

Hallinan, Erin K. 2220<br />

Hallsworth, Kate 2183<br />

Halm, Ethan 548<br />

Halseth, Amy 2175<br />

Halsted, Charles H. 2109<br />

Hamano, Mina 796<br />

Hamdaoui, Nabila 693<br />

Hamdeh, Shadi 748, 769<br />

Hameed, Bilal 160, 236, 1777<br />

Hametner, Stephanie 1092, 1885<br />

Hamid, Saeed 1872<br />

Hamilton, James P. 1749<br />

Hamilton, Thomas 668<br />

Hammad, Radi 1076, 1163<br />

Hammami, Muhammad B. 1535<br />

Hammond, Rachel 1551, 1563<br />

Hampe, Jochen 617<br />

Han, Chang 185, 686<br />

Han, Guo Rong 209<br />

Han, Guohong 468, 754, 851<br />

Han, Ho Seong 857, 1902<br />

Han, Huanzi 1370<br />

Han, Hui 588<br />

Han, Jude D. 1366<br />

Han, Juqiang 1595<br />

Han, Kwang-Hyub 91, 346, 504, 1080,<br />

1105, 1168, 1542, 1558, 1581, 1605<br />

Han, Lingling 249<br />

Han, Ma Ai Thanda 1720, 2144<br />

Han, Meifang 1656, 1683, 1684<br />

Han, Ping 328<br />

Han, Seungbong 446<br />

Han, Sheng-Na 1284, 2179<br />

Han, Steven-Huy B. 488, 2015, 2052<br />

Han, Yuan-Ping 904<br />

Han, Yuyan 190, 454, 835, 929,<br />

1317, 1349<br />

Han, Zhou 709<br />

Hanczko, Robert 1961<br />

Hand, Fiona M. 1216<br />

Handa, Priya 946<br />

Hanes, Justin 115<br />

Hanf, Rémy 105, 162, 974, 2145<br />

Hangartner, Thomas N. 1694<br />

Hanje, A. James 213, 874, 1260,<br />

1280, 1759<br />

Hann, Hie-Won 371, 1554, 2004<br />

Hannan, Nicholas R. 258<br />

Hanouneh, Ibrahim A. 840, 855, 856,<br />

1215, 1265<br />

Hansen, Bettina E. 68, 73, 76, 348,<br />

445, 601, 609, 634, 1166, 1196,<br />

1568, 1585, 1586, 1590, 1998,<br />

2001, 2002, 2003, 2012, 2013<br />

Hansen, Janne F. 780, 789, 792, 1440<br />

Hansi, Navjyot K. 167, 1571<br />

Hanson, Jeffrey 166<br />

Hanson, John 722<br />

Hao, Ke 253, 254<br />

Hao, Ying 451<br />

Hara, Tasuku 886, 2230<br />

Hara, Yuichi 995, 1937, 2199<br />

Harada, Masaru 593, 883, 1819<br />

Haraguchi, Masafumi 1671<br />

Harbers, Marten 208, 2263<br />

Harbonnier, Jean 1796<br />

Hardesty, Josiah 666<br />

Hardie, Claire 611<br />

Hardison, Regina M. 1714<br />

Hardtke, Svenja 1572<br />

Hardwick, James P. 925<br />

Hardy, Timothy 1408<br />

Hargrove, Laura 812, 822, 1363<br />

Haridy, James 311<br />

Harif, Yael 875<br />

Hariharan, Siddharth 316<br />

Hariri, Susan 1787<br />

Harms, Maren H. 73, 634<br />

Harmsen, Martin C. 1416<br />

Harmsen, William S. 397<br />

Harney, Sarah 136<br />

Harnois, Denise M. 50, 53, 1236, 1245<br />

Harper, Ann M. 861, 872<br />

Harrell, Sherrie M. 1099, 1857<br />

Harrigan, Richard 1832<br />

Harriman, Geraldine 957, 1938<br />

Harrington, Andrew 253<br />

Harris, Aaron M. 1574<br />

Harris, Edward N. 507<br />

Harris, Melissa 726<br />

Harrison, Stephen A. 105, 162, 737,<br />

1428, 1435, 2145, 2151, 2239<br />

Harrison-Findik, Duygu Dee 60<br />

Harry, Kathy M. 829<br />

Hartl, Astrid 2068<br />

Hartl, Johannes 523<br />

Hartleif, Steffen 1728<br />

Hartman, George D. 33, 2064<br />

Hartman, Joshua 1095<br />

Hartman-Neumann, Sandra 709<br />

Hartmann, Bolette 1689<br />

Hartmann, Dagmar 1170, 1200<br />

Hartmann, Phillipp 59<br />

Harty, Alyson 1095, 1151, 1185, 1202<br />

Haruna, Yoshimichi 173<br />

Harvey, Brandon K. 323<br />

Harvey, Ray 1050<br />

Harwood, H. James 957, 1938<br />

Hasan, Mohsin 954<br />

Hasan, Raza 1906<br />

Hasanaliyeva, Nigar 161, 2169<br />

Hasanin, Mohsen 1249, 1250<br />

Hasebe, Takumu 927<br />

Hasegawa, Daisuke 1392, 1405<br />

Hasegawa, Kunihiro 466, 770, 781,<br />

1533, 1669<br />

Hashem, Mohamed 1142<br />

Hashiba, Tomomi 699, 1953<br />

Hashimoto, Etsuko 1326, 2207<br />

Hashimoto, Koji 1271<br />

Hashimoto, Satoru 408, 783, 1600<br />

Hashimoto, Shigeatsu 551<br />

Hashmi, Haroon 36<br />

Haslett, Patrick 36<br />

Hasnain, Nadeem 1268<br />

Hassan, Ammar 1791<br />

Hassan, Mohamed A. 1101, 1184<br />

Hassanein, Tarek I. 41, 106, 627, 714,<br />

1120, 1813<br />

Hassany, Mohamad 708, 1076, 1163<br />

Hassona, Ehab M. 1011<br />

Hatano, Etsuro 402, 1213, 1918, 1976<br />

Hatooka, Masahiro 472, 1893<br />

Hatting, Maximilian 828<br />

Hattori, Noburiro 1446<br />

Hatzoglou, Maria 668<br />

Hau, Hans-Michael 1222<br />

Hauber, Joachim 1629<br />

Haufe, William 45, 913<br />

Hausding, Michael 950


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1329A<br />

Hawley, Richard C. 521<br />

Hay, Christopher 753<br />

Hayakawa, Masako 918<br />

Hayashi, Eiichi 1302, 1941<br />

Hayashi, Kazuhiko 411, 443, 880,<br />

967, 1061, 1167<br />

Hayashi, Norio 1197, 1201<br />

Hayashi, Paul H. 14, 533, 1922, 1923,<br />

1926, 1931, 2093<br />

Hayashi, Sanae 1655<br />

Hayashi, Takehiro 699, 1000, 1953<br />

Hayashi, Tomoyuki 699, 1953<br />

Hayata, Yuki 382<br />

Haydar, Ghada 1909<br />

Haydel, Brandy M. 988, 1289<br />

Hayden, Hubert 1525<br />

Hayes, C. Nelson 986, 1627, 1645,<br />

1649, 1681, 1858, 1895, 1944, 2120<br />

Hayes, Don 1280<br />

Hayes, Nelson 663, 1150, 1670, 1679<br />

Hayes, Peter C. 260, 753<br />

Haynes, Meagan 670, 1979<br />

Hazari, Sidhartha 1096<br />

Hazra, Saswati 239<br />

He, Aiwu R. 2118<br />

He, Chuangye 754<br />

He, Dengming 1865<br />

He, Hua 245<br />

He, Jia 919<br />

He, Lina 1945<br />

He, Xiao-Song 77, 602, 613, 639<br />

He, Xiuting 1580, 1620<br />

He, Yingli 1541, 1579, 1617, 1773<br />

He, Yong 182<br />

Heathcote, E. Jenny 348<br />

Heaton, Nigel 212, 1733<br />

Heba, Elhamy 45, 2150, 2201<br />

Heckmann, Melanie 1726<br />

Hede, Shalini 1456, 1461<br />

Hedskog, Charlotte 219<br />

Heegsma, Janette 612, 1398<br />

Hegde, Pushpa 1391<br />

Hegge, Julia O. 32, 2022<br />

Heidrich, Benjamin 1002, 1572<br />

Heier, Eva-Carina 1299, 1385<br />

Height, Susan 1711, 1723<br />

Heikenwälder, Mathias 2128<br />

Heim, Markus H. 221, 1884<br />

Heimbach, Julie 6, 50, 53, 836, 837,<br />

1226, 1263<br />

Heinisch, Birgit B. 149<br />

Heinz, Stefan 331<br />

Hellard, Margaret 1083<br />

Heller, Theo 218, 779, 1498, 1650,<br />

1717, 1720, 2144<br />

Hellerstein, Marc 1425<br />

Hellmich, Brigitte 1886<br />

Helmke, Steve M. 1849<br />

Helmy, Housam 733<br />

Heluwaert, Frédéric 1072<br />

Hemmingsson, Tomas 158<br />

Henao, Ricardo 161<br />

Henderson, Mark 323<br />

Hendrikx, Tim 922<br />

Heneghan, Michael A. 212, 300, 312,<br />

608, 1098, 1744, 1758<br />

Hénique, Carole 887<br />

Henkel, Anne 257<br />

Henriksen, Kim 2178<br />

Henriques Abreu, Pedro 473<br />

Henry, Linda 18, 315, 316, 1035,<br />

1460, 2212<br />

Henry, Mitchel 1280<br />

Henry, Zachary 13, 761<br />

Hentati, Hassen 215<br />

Heo, Jeong 749, 876, 2028, 2038<br />

Heo, Nae-Yun 88, 1208, 2155<br />

Hepner, Ashley 1111, 1155, 1164<br />

Herath, Chaturika 1246<br />

Herber, Adam 276<br />

Heredia, Nancy 1770<br />

Herlong, H. Franklin 682, 2118<br />

Herman, Mark A. 2227<br />

Hernaez, Ruben 1214<br />

Hernandez, Brenda Y. 1553<br />

Hernandez, Carolyn 45, 913, 1425<br />

Hernandez, Céline 30<br />

Hernandez, Dennis 709<br />

Hernandez, Maria D. 457<br />

Hernández, Nelia 225, 595<br />

Hernandez-Alejandro, Roberto 837<br />

Hernandez-Guedea, Marco A. 1281<br />

Hernandez-Luis, Ruben 1341<br />

Hernandez-Rocha, Cristian 2210<br />

Hernandez-Santos, Nydiaris 1650<br />

Herrin, Ann 549<br />

Herrine, Steven K. 1452<br />

Hertel, Paula M. 133, 1694<br />

Herve, Camille 3<br />

Herzer, Kerstin 37, 252, 1058<br />

Heubi, James E. 1694<br />

Heuman, Douglas M. 52, 1463, 1472,<br />

1491, 1527<br />

Heurgué-Berlot, Alexandra 264, 308,<br />

1072<br />

Hew, Huong 1832<br />

Hewitt, Annette 1790, 1798, 1807<br />

Heymann, Felix 100, 1756<br />

Hezode, Christophe 42, 206, 714,<br />

1086, 1102, 1107, 1123<br />

Hézode, Christophe 1442<br />

Hiasa, Yoichi 338, 788, 1293, 2146<br />

Hibi, Taizo 336<br />

Hickey, Raymond D. 2<br />

Hickman, Matthew 1794<br />

Hickson, Ford 1830<br />

Hidaka, Hisashi 756<br />

Hidalgo, Juan 1295<br />

Hidvegi, Tunda 193, 197<br />

Hige, Shuhei 1116, 2065<br />

Higuchi, Kazuhide 966<br />

Higuchi, Nobito 1159<br />

Higuera, Monica 253<br />

Hijdra, Rosanne M. 89, 1112<br />

Hikita, Hayato 26, 97, 231, 453, 669,<br />

1082, 1197, 1357, 1635, 1636,<br />

1642, 1661, 1809, 1982, 1985,<br />

2130, 2136<br />

Hilburn, David 658<br />

Hildt, Eberhard 1559<br />

Hill, Andrew M. 1826<br />

Hillaire, Sophie 145<br />

Hillman, Luke 211<br />

Hilscher, Moira 605<br />

Hinkle, Greg 36<br />

Hino, Keisuke 995, 1632, 1937, 2199<br />

Hinrichsen, Holger 1078, 1435<br />

Hinson, Jack A. 1914<br />

Hinton, Alice 874<br />

Hiraga, Nobuhiko 663, 986, 1150,<br />

1627, 1645, 1649, 1670, 1681,<br />

1858, 1895, 1944, 2120<br />

Hirahara, Kazuki 1486<br />

Hiramatsu, Akira 472, 1893, 2199<br />

Hiramatsu, Katsushi 944, 1682<br />

Hiramatsu, Naoki 26, 231, 453, 669,<br />

1082, 1195, 1197, 1201, 1357,<br />

1635, 1636, 1809, 1982, 1985,<br />

2130, 2136, 2213<br />

Hiramatu, Akira 1056<br />

Hirano, Tadamichi 1420<br />

Hiraoka, Atsushi 524<br />

Hirashima, Noboru 1097<br />

Hirayama, Tasuku 995<br />

Hiriart, Jean-Baptiste 450, 1796, 2181<br />

Hirooka, Masashi 788, 1293<br />

Hirooka, Yoshiki 411, 880, 967, 1061,<br />

1167<br />

Hirose, Shunji 314, 383<br />

Hirota, Seiichi 781<br />

Hirsch, Amy 1154<br />

Hirsch, Geri 204, 647<br />

Hirschbein, Jonah 1479<br />

Hirschfield, Gideon 73, 76, 348, 607,<br />

631, 634, 636, 1234, 1457, 1710<br />

Hirsova, Petra 825, 895, 1315<br />

Hisatake, Garrett M. 538<br />

Hiththetiya, Kanishka 80<br />

Hittatiya, Kanishka 961, 2115<br />

Hjuler, Sara T. 2178<br />

Ho, Chanda 538<br />

Ho, Hsiu-Jon 345<br />

Ho, Ja-an Annie 654<br />

Ho, Samuel B. 59, 529, 549, 1041<br />

Hoang, Joseph K. 207, 359, 368, 386,<br />

388, 458, 1561, 1805, 1806, 1816,<br />

1848, 2250<br />

Hoang, Tanya 677<br />

Hobbs, Ursula 1452<br />

Hocking, Kyle M. 1223<br />

Hodge, Alexander 47, 945<br />

Hodson, James 574, 1728<br />

Hoefs, John C. 269<br />

Hoeke, Martijn O. 612<br />

Hoekstra, Mark 612<br />

Hoener zu Siederdissen, Christoph<br />

1131, 1803<br />

Hoerauf, Achim 2077<br />

Hoermann, Rudolf 842<br />

Hoevelmeyer, Nadine 1949<br />

Hofer, Harald 297, 732, 1070, 1092,<br />

1137, 1885<br />

Hoffman, Hal M. 1371<br />

Hoffmann, Udo 2191<br />

Hoffmann, Vera 687<br />

Hofker, Marten H. 922<br />

Hofmann, Alan F. 1915<br />

Hofmann, Wolf P. 1166<br />

Hogan, Brian J. 793, 1530<br />

Hohenester, Simon 609<br />

Höijer, Jonas 116<br />

Hoki, Toshifumi 1984<br />

Holder, Beth S. 300<br />

Hollabaugh, Kimberly 1094<br />

Holland-Fischer, Peter 1482<br />

Hollenbach, Marcus 1519, 1520<br />

Hollingsworth, Kieren G. 2183<br />

Holm, Kristian 76, 833


1330A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Holmberg, Scott D. 15, 525, 1789,<br />

1799, 1800<br />

Holmes, Benjamin 510, 2116<br />

Holmes, Jacinta A. 996<br />

Holmstom, Fredrik 992<br />

Holst, Jens 1689<br />

Holt, Andrew 1216, 1321<br />

Holt, Edward W. 306, 1252, 2159<br />

Holtzman, Deborah 525<br />

Holzmayer, Vera 1797<br />

Homan, Chriss E. 1790, 1798, 1807<br />

Homs, Maria 1021, 1674<br />

Honda, Akira 970<br />

Honda, Hiroki 310<br />

Honda, Koichi 2207, 2267<br />

Honda, Masao 144, 168, 325, 495,<br />

657, 699, 947, 990, 994, 1000,<br />

1010, 1014, 1348, 1377, 1395,<br />

1632, 1653, 1953, 1968, 1970<br />

Honda, Takashi 411, 443, 880, 967,<br />

1061, 1167<br />

Honda, Takuya 1671<br />

Hong, Jian 991, 1015, 1630, 1644<br />

Hong, Leena K. 1121<br />

Hong, Qiuting 606<br />

Hong, Sung Woo 1419<br />

Hong, Thai 442, 483<br />

Hong, Youngmi 876, 2028<br />

Honma, Yuichi 593, 883<br />

Hoofnagle, Jay H. 121, 218, 236, 597,<br />

779, 1598, 1922, 1923, 1926, 1931<br />

Hooker, Catherine A. 913, 2150<br />

Hooker, Jonathan 45, 913<br />

Hooper, Lora V. 111<br />

Hooshmand-Rad, Roya 317, 609, 625,<br />

628, 644<br />

Hoque, Rafaz 979, 2179<br />

Horban, Andrzej 2004<br />

Horberg, Michael A. 1811, 1871<br />

Horie, Yoshinori 2177<br />

Horii, Rika 1014<br />

Horiike, Norio 524<br />

Horner, Mary 1113, 1588, 1704,<br />

1744, 2053, 2054<br />

Hornsey, Emma 798<br />

Horsfall, Leigh 784<br />

Horsmans, Yves J. 1051<br />

Horvath, Angela 59, 2068<br />

Hosaka, Tetsuya 1062, 1191, 2020,<br />

2241<br />

Hoshida, Yohmei 513<br />

Hoshida, Yujin 229, 253, 254, 364,<br />

778, 1337, 1392, 1742, 1940, 2174<br />

Hoshino, Takashi 477<br />

Hosho, Keiko 1499<br />

Hosoya, Mitsuaki 551<br />

Hot, Edina 893<br />

Hou, Feng-Qin 2023<br />

Hou, Jin-Lin 2013<br />

Hou, Jun 1637<br />

Hou, Mengjun 454<br />

Hou, Ming-Chih 738, 751, 757, 759<br />

Hou, Xiaohua 1441<br />

Houben, Tom 922<br />

Houghton, David 2183<br />

Houlihan, Diarmaid D. 1103, 1216,<br />

1275<br />

Houry, Mohamad 2173, 2192<br />

House, Michael J. 798<br />

Houssel-Debry, Pauline 3, 95, 1781<br />

Housset, Chantal 196, 642, 786, 1409<br />

Hov, Johannes R. 833<br />

Howarth, Rachel M. 1381, 1408<br />

Howe, Anita Y. 40, 42, 210, 700, 701,<br />

703, 707<br />

Howell, Brett A. 224, 228<br />

Howieson, Kevin 1051<br />

Hsiang, John C 1612<br />

Hsieh, Matthew 1498<br />

Hsieh, Tsai-Yuan 1996<br />

Hsieh, Wei-Yao 751<br />

Hsieh, Yun-Cheng 757, 965<br />

Hsin, I-Fang 738, 751<br />

Hsing, Ann W. 207<br />

Hsu, Chia-Yang 365, 409<br />

Hsu, Hong-Yuan 1701, 1705<br />

Hsu, Ping-I 738<br />

Hsu, Shih-Jer 1996<br />

Hsu, Shu-hao 493<br />

Hsu, Yao-Chun 1547<br />

Hu, Bingqian 79<br />

Hu, Chengcheng 1716<br />

Hu, Chi-Tan 1554, 1996<br />

Hu, Donglei 78<br />

Hu, Haihong 1811, 1871<br />

Hu, Huaidong 1647<br />

Hu, Hui-Han 1593, 1604<br />

Hu, Jiangting 1925<br />

Hu, Jinhua 1675<br />

Hu, Junjie 1934<br />

Hu, Ke-Qin 85, 1031, 1165, 1785<br />

Hu, Peng 1044, 1647<br />

Hu, Richard 904<br />

Hu, Shi 1910<br />

Hu, Tsung-Hui 1873, 1942, 1996,<br />

2010<br />

Hu, Wei 100, 1924<br />

Hu, Xiaojun 1755<br />

Hu, Xudong 879, 1324, 1338<br />

Hu, Yaoren 2048<br />

Hu, Ying 508, 649, 674<br />

Hu, Yiran 1065<br />

Hua, Rui 1149, 2107<br />

Huaman, Moises 854<br />

Huang, Changxing 984<br />

Huang, Chao-Cheng 1942<br />

Huang, Chi-Wen 1701<br />

Huang, Chia-Lin 2040<br />

Huang, Chiung-Kuei 256, 1962<br />

Huang, Chung-Feng 2264<br />

Huang, Guangyu 1865<br />

Huang, Hai 184<br />

Huang, Hongfei 2024<br />

Huang, Jee-Fu 398, 2264<br />

Huang, Jhy-Shrian 398<br />

Huang, Kai-Wen 118, 1449<br />

Huang, Li 454<br />

Huang, Shu-Pang 720<br />

Huang, Shuo 1988<br />

Huang, Vivian 1606<br />

Huang, Wen 1812<br />

Huang, Xiaobi 731<br />

Huang, Xiaohui 1772<br />

Huang, Yi 1264<br />

Huang, Yi-Hsiang 365, 409, 439, 759,<br />

965, 1996, 2050<br />

Huang, Yi-Wen 1996<br />

Huang, Ying 1441<br />

Huang, Zhanlian 1423<br />

Huard, Genevieve 1997<br />

Huber, Caroline 1144, 1462<br />

Huber, Hans 118<br />

Hubers, Lowiek M. 637<br />

Hübner, Marc P. 2077<br />

Huck, Ian 650<br />

Hucke, Florian 475<br />

Huddleston, Heather 2252<br />

Huddleston, Leslie 1121<br />

Hudock, Rebecca 1570<br />

Hudson, Benjamin E. 564, 1140, 1173<br />

Hudson, Mark 2226<br />

Huebert, Robert C. 692, 1355<br />

Huelin, Patricia 266<br />

Hueppe, Dietrich 1060, 1078, 1081,<br />

1156, 1861<br />

Hughes, Dempsey 412<br />

Hughes, Dominic A. 1691<br />

Hughes, Eric A. 702<br />

Hughes, Michael 1094<br />

Hui, Aric Josun 241, 2014, 2025<br />

Hui, Peter 2104<br />

Hull, Katherine L. 62<br />

Hullegie, Sebastiaan 1007<br />

Hultcrantz, Rolf W. 158<br />

Hum, Dean W. 105, 162, 974, 2145<br />

Humar, Abhinav 838<br />

Humar, Bostjan 499, 665<br />

Humberson, Annette 1265<br />

Humbert, Lydie 816<br />

Hummels, Hazel 1373<br />

Humphrey, Mark 1986<br />

Hunault, Gilles 2188<br />

Hung, Annie K. 462<br />

Hung, Chao-Hung 1873, 2010<br />

Hung, Hsu-Wei 1547<br />

Hung Wan, Isabel 1121<br />

Hunt, Christine M. 540<br />

Hunt, Katie 1712<br />

Hunt, Kristel K. 16<br />

Hunt, Sharon L. 18, 315, 534, 1035,<br />

1460<br />

Hunter, Jill E. 652<br />

Hunter, Stuart 1298<br />

Huntsman, Scott 78<br />

Huntzicker, Erik G. 632, 1359, 2149<br />

Huo, Teh-Ia 356, 365, 409, 439, 759<br />

Huppert, Kari A. 680<br />

Huppert, Stacey S. 680<br />

Hur, Chin 154<br />

Hur, Keun 403, 426, 749<br />

Hur, Wonhee 1419<br />

Hurt, Christopher B. 1063<br />

Hussain, Shabir 581, 1344<br />

Hussain, Zilla H. 580<br />

Hussaini, Kamran 1535<br />

Hussaini, Tarana 204<br />

Hussein, Abd Elrazek A. 858, 2215<br />

Hutabarat, Renta M. 36<br />

Hutchinson, Sharon 1794<br />

Hutton, David W. 67<br />

Huynh, Andrew 1182, 1549<br />

Huynh, Dep K. 76<br />

Hwang, Jaeseok 2038<br />

Hwang, Kyu-Hee 1401<br />

Hwang, Peggy 42, 251, 700, 701,<br />

703, 715<br />

Hwang, Seong Gyu 1542<br />

Hwang, Sunil 1923<br />

Hyland, Robert H. 38, 247, 713, 777,<br />

1034, 1045, 1120, 1130, 1442, 1860


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1331A<br />

Hylemon, Phillip B. 22, 804, 1463,<br />

1521, 1527<br />

Hynan, Linda S. 1776<br />

Hyogo, Hideyuki 2156, 2158, 2199,<br />

2202, 2207<br />

Hyun, Jeongeun 696<br />

Hyun, Jinhee 1017<br />

Iannitti, David A. 653<br />

Iaria, Pierre 2082<br />

Iavarone, Massimo 229, 393, 435<br />

Ibanez, Luisa 225<br />

Ibis, Mehmet 932<br />

Ibrahim, Samar 895<br />

Ibusuki, Rie 937, 1421, 1838<br />

Ichai, Philippe 1237, 1269, 1782<br />

Ichikawa, Tatsuki 1671<br />

Ichiki, Yasunori 2031<br />

Ida, Kinuyo 697<br />

Ide, Tatsuya 304, 1448, 1819<br />

Ide, Yasushi 296<br />

Idilman, Ramazan 558, 1822, 2012,<br />

2033<br />

Ido, Akio 799, 937, 1421, 1838<br />

Idowu, Michael 905<br />

Ieluzzi, Donatella 1166<br />

Iemmolo, Rosa Maria 1751<br />

Iezzi, Roberto 432<br />

Iguchi, Kohta 402<br />

Iida, Noriho 370, 392, 430<br />

Iida, Takashi 1696<br />

Iijima, Hiroko 466, 770, 781, 1533,<br />

1669, 1975<br />

Iijima, Sayuki 1018<br />

Iimuro, Yuji 188<br />

Iio, Etsuko 1006, 1097, 1448, 1655,<br />

1954, 1983<br />

Iio, Sadaharu 1809<br />

Ijuin, Sho 799, 937, 1838<br />

Ikarashi, Yuichi 1326<br />

Ikeda, Fusao 2236<br />

Ikeda, Hiroki 1446<br />

Ikeda, Kazuo 517, 592<br />

Ikeda, Kenji 1062, 1191, 2020, 2241<br />

Ikeda, Naoto 466, 770, 781, 1533,<br />

1669, 1975<br />

Ikeda, Yasuhiro 2<br />

Ikeda, Yoshio 2146<br />

Ikegami, Tadashi 970<br />

Ikejima, Kenichi 226, 917, 1771<br />

Ikenaga, Naoki 826, 1367, 1368,<br />

1379<br />

Ikeno, Yoshinobu 1213<br />

Ikezono, Yu 688<br />

Ilan, Yaron 46<br />

Ilani, Nadav 1878<br />

Ildstad, Suzanne 54<br />

Iliaz, Raim 478<br />

Iliceto, Sabino 268<br />

Illa, Xavi 8<br />

Ilyas, Ghulam 1333<br />

Im, Gene Y. 1095, 1185, 1337<br />

Imagawa, Kazuo 1690<br />

Imai, Norihiro 880<br />

Imai, Yasuharu 1195, 1197, 2213<br />

Imai, Yukinori 1474, 1486<br />

Imai, Yumi 807<br />

Imai, Yusuke 338, 788, 1293<br />

Imajo, Kento 2207<br />

Imam, Mohamad 76<br />

Imamine, Rinpei 402<br />

Imamura, Masatoshi 794, 1795, 2158,<br />

2193<br />

Imamura, Michio 472, 986, 1056,<br />

1150, 1627, 1645, 1649, 1670,<br />

1681, 1858, 1893, 1895, 1944, 2120<br />

Imazeki, Fumio 1207<br />

Imbert-Bismut, Françoise 786, 1510<br />

Imteyaz, Hejab 280<br />

Inada, Masami 1197, 1201<br />

Inada, Yuki 370<br />

Inagaki, Yutaka 1369<br />

Inami, Yoshihiro 1771<br />

Inao, Mie 1059, 1064, 1474, 1486,<br />

2241<br />

Iñarrairaegui, Mercedes 344, 378<br />

Indenbirken, Daniela S. 1629<br />

Ingiliz, Patrick 1058, 1060, 1081,<br />

1156<br />

Inglot, Malgorzata 1161<br />

Ingviya, Thammasin 280, 294, 560<br />

Inomata, Kenta 336<br />

Inoue, Atsuo 173, 1201, 2213<br />

Inoue, Haruhiro 1941<br />

Inoue, Jun 1026, 1648, 1943, 2167<br />

Inoue, Taisuke 1565, 1673<br />

Inoue, Takako 1573, 1599, 1618<br />

Inoue, Tomoyoshi 756<br />

Intagliata, Nicolas 761<br />

Invernizzi, Federica 2011, 2012<br />

Invernizzi, Pietro 73, 76, 602, 609,<br />

621, 634, 1654<br />

Ioannou, George N. 49, 128, 230,<br />

540, 803, 923, 1211, 2151<br />

Iodice, Valentina 431<br />

Iovanescu, Vlad F. 755<br />

Ipharraguerre, Ignacio R. 194<br />

Iracheta-Vellve, Arvin 180, 496, 1318,<br />

1330, 2198<br />

Iredale, John P. 62, 260, 329, 753<br />

Ireland, Hamish 753<br />

Iriana, Sentia 566<br />

Irie, Makoto 385, 410<br />

Irvine, Katharine 784, 1362<br />

Irving, William 1140, 1173<br />

Isakov, Vasily 721<br />

Ishiba, Hiroshi 886, 888, 2230<br />

Ishibashi, Hiromi 2031<br />

Ishibashi, Naoto 657<br />

Ishida, Hisashi 1584<br />

Ishida, Kosuke 325, 947<br />

Ishida, Yuji 1670, 1679<br />

Ishigami, Masatoshi 411, 443, 880,<br />

967, 1061, 1167<br />

Ishihara, Akio 1584<br />

Ishii, Akio 466, 770, 781, 1533, 1669<br />

Ishii, Kunihide 304<br />

Ishii, Takamichi 1976<br />

Ishikawa, Tsuyoshi 760, 1447, 1464,<br />

1492<br />

Ishizu, Yoji 411, 443, 880, 967, 1061,<br />

1167<br />

Islam, Fakhar U. 1257<br />

Isogawa, Masanori 1640, 1655<br />

Isoyama, Shigemi 1690<br />

Issachar, Assaf 875<br />

Isted, Alexander 1733<br />

Itakura, Jun 1115, 1192<br />

Itami, Saori 429<br />

Itano, Osamu 336<br />

Ito, Daisaku 382<br />

Ito, Kiyoaki 2158<br />

Ito, Koichi 135<br />

Ito, Masahiko 1393<br />

Ito, Rei A. 1420<br />

Ito, Sayaka 505<br />

Ito, Takayoshi 1302, 1941<br />

Ito, Toshifumi 796, 1197<br />

Itoh, Fumio 1006, 1446<br />

Itoh, Ikuyo 901<br />

Itoh, Jun 287, 1946<br />

Itoh, Yoshito 232, 707, 886, 888, 901,<br />

1980, 2156, 2202, 2230<br />

Itokawa, Norio 291, 476<br />

Itou, Minoru 233<br />

Itsui, Yasuhiro 164, 406, 1018, 1027,<br />

1959<br />

Iwagami, Yoshifumi 256, 1962<br />

Iwaisako, Keiko 1918<br />

Iwakiri, Katsuhiko 291<br />

Iwakiri, Yasuko 1322, 1526<br />

Iwama, Itaru 1724<br />

Iwamoto, Hideki 688<br />

Iwamoto, Junichi 970<br />

Iwamoto, Marian 725, 730, 731<br />

Iwamoto, Takuya 760, 1464, 1492<br />

Iwao, Masao 2267<br />

Iwasa, Motoh 584, 1500<br />

Iwasaki, Ryuichiro 1584<br />

Iwasaki, Tetsuya 1584<br />

Iwase, Tomoya 1618<br />

Iwata, Kaoru 410<br />

Iwata, Tomoaki 2167<br />

Iwata, Yoshinori 466, 770, 781, 1533,<br />

1669, 1975<br />

Iwelu, Chibuzor 1233<br />

Iyer, Malliga 1374<br />

Iyer, Radhakrishnan P. 2262<br />

Izmaylov, Michelle 1223<br />

Izumi, Kousuke 1771<br />

Izumi, Namiki 1115, 1192, 1632,<br />

2029<br />

Izumi, Takaaki 1672<br />

Izzo, Francesco 851<br />

J Sangwaiya, Minal 791, 800<br />

Jaber, Fadi L. 61, 340<br />

Jablkowski, Maciej S. 1678<br />

Jachymova, Marie 1487<br />

Jackson, Kathryn 563, 849<br />

Jackson, Kathy 1566<br />

Jacobs, W. Carl 1023<br />

Jacobson, Ira M. 42, 91, 1086, 1106,<br />

1841<br />

Jacobson, Karen B. 89<br />

Jacomet, Christine 1093<br />

Jacomino, Kristina 1223<br />

Jacquemin, Emmanuel 196<br />

Jacquemyn, Bert 39<br />

Jacques, Vincent 143<br />

Jadeja, Ravirajsinh 1911<br />

Jadhav, Vasant 36<br />

Jaeschke, Hartmut 585, 586, 598,<br />

1767, 1774, 1925<br />

Jafoui, Sami 1779<br />

Jafri, Syed M. 307, 1872<br />

Jahnel, Jörg 1726<br />

Jain, Mamta K. 438<br />

Jain, Surbhi 1554<br />

Jain, Vikas 1340<br />

Jalan, Rajiv 113, 908, 1300, 1482,<br />

1517, 1769, 1780, 1781, 2069


1332A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Jalan-Sakrikar, Nidhi 692<br />

James, Laura 99, 1763, 1914<br />

James, Spencer L. 580<br />

James, Theodore W. 1255<br />

Jamil, Khurram 278<br />

Jamwal, Kapil D. 2070, 2092<br />

Janardhan, Sujit V. 522<br />

Jang, Byoung Kuk 2038<br />

Jang, Eun Chul 924, 941, 1760, 2110,<br />

2218, 2243<br />

Jang, Eun Sun 389, 2113<br />

Jang, Jae Seong 857<br />

Jang, Jae Yool 857<br />

Jang, Jae Young 277, 484, 736<br />

Jang, Jeong Won 452, 1615, 1882<br />

Jang, Kiseok 2218<br />

Jang, Se Young 363, 403, 426, 749<br />

Jang, Seung Hyeon 2049<br />

Jang, Sun Kyung 363, 403, 426, 749<br />

Janjua, Naveed Z. 86, 543<br />

Jansen, Christian 773<br />

Jansen, Geraldine 1613<br />

Jansen, Jos 2102<br />

Jansen, Louis 1633, 1995, 2005, 2046<br />

Jansen, Peter L. 604<br />

Janssen, Harry L. 73, 121, 241, 243,<br />

348, 350, 445, 542, 616, 634, 1099,<br />

1166, 1556, 1585, 1586, 1590,<br />

1637, 1651, 1998, 2001, 2002,<br />

2003, 2012, 2013, 2037, 2059, 2255<br />

Janssen, Katrien 1040<br />

Jara, Evelyn 971<br />

Jarnik, Michal 20<br />

Jaroszewicz, Jerzy 1502<br />

Jarufe, Nicolás 1217<br />

Jarupongprapa, Saharat 467<br />

Jaruvongvanich, Veeravich 964<br />

Jaskowski, Lesley-Anne 975<br />

Jasser-Nitsche, Hildegard 1726<br />

Jauffret-Roustide, Marie 1845<br />

Javaherian, Kavon 1451<br />

Javaid, Amen 286<br />

Javaid, Asad 371<br />

Javle, Milind 130, 682<br />

Jayaprakash, Aravindakshan 118<br />

Jayasekera, Channa R. 199, 201, 541,<br />

1534, 1745, 1750<br />

Jeamsripong, Woramon 1631, 1662<br />

Jeddari, Safaa 165<br />

Jeen, Yoon Tae 1504<br />

Jeffers, Lennox J. 2256<br />

Jeffers, Thomas 956<br />

Jeffery, Hannah C. 1298<br />

Jeffrey, Gary P. 848, 1077, 1264,<br />

2099, 2185<br />

Jelsing, Jacob 978<br />

Jen, Chin-Lan 1593, 1604<br />

Jen, Kuang-Yu 1736<br />

Jenab, Mazda 170<br />

Jeng, Rachel Wen-Juei 1552<br />

Jenkins, Nancy 234<br />

Jenkins, Stephen 329<br />

Jensen, Donald M. 539, 544<br />

Jensen, Majken K. 1702, 2228<br />

Jeon, Christie 2083<br />

Jeon, Hoonbae 1266<br />

Jeon, Tae Joo 464<br />

Jeon, Woo Kyu 1624, 1992<br />

Jeong, Jong-Min 910, 1358<br />

Jeong, Sook-Hyang 389, 2113<br />

Jeong, Soung Won 277, 484, 736<br />

Jeong, Won-IL 910, 1358, 1414<br />

Jeong, Woo Kyoung 485<br />

Jeong, Yun Seong 130, 682<br />

Jesse, Michelle T. 556<br />

Jessen, Niels 301<br />

Jessop, Amy B. 547<br />

Jeurissen, Mike 922<br />

Jewell, Mark 66, 936, 1386<br />

Jezequel, Caroline 264<br />

Ji, Cheng 588<br />

Ji, Dong 1536, 1537, 2008<br />

Jia, Jidong 374, 1404, 1660, 2048<br />

Jia, Zhansheng 984, 1004<br />

Jiang, An 255<br />

Jiang, Deyuan 1076, 1860<br />

Jiang, Guanglong 1332<br />

Jiang, Hai-Xing 1465<br />

Jiang, Hong Xiu 209<br />

Jiang, Hui 911<br />

Jiang, Jie 1583<br />

Jiang, Jing 1149, 1193, 2018<br />

Jiang, Joy 1285<br />

Jiang, Lijuan 2257, 2259, 2260<br />

Jiang, Tao 1149<br />

Jiang, Wei 1383<br />

Jiang, Xuemin 2259, 2260, 2261<br />

Jiang, Yanjun 194, 1689<br />

Jiang, Yong 1544<br />

Jiang, Zhaoshi 1651, 1668<br />

Jiang, Zhenghui G. 2227, 2228<br />

Jiang, Zhiwei 729<br />

Jiao, Chunhua 1521<br />

Jiao, Jingjing 1370, 1424<br />

Jibara, Ghalib 851<br />

Jideh, Bilel 1450<br />

Jimbei, Paulina 31<br />

Jiménez, Wladimiro 65, 659<br />

Jimenez Exposito, Maria Jesus 37, 252,<br />

1058<br />

Jiménez-Castro, Mónica B. 342<br />

Jimeno, Carlota 1863<br />

Jin, Dongfang 1579, 1617<br />

Jin, Jinglan 1657<br />

Jin, Li 1541, 1579, 1617, 1773<br />

Jin, Michelle Q. 367, 388, 1848, 2009<br />

Jin, Mingjuan 359, 367, 388, 459,<br />

1806, 1816, 1848<br />

Jin, Quan 919<br />

Jin, Ran 2232<br />

Jin, Young-Joo 1542<br />

Jindal, Ankur 146, 2070<br />

Jindal, Rohit 25<br />

Jinushi, Masahisa 505<br />

Jirsa, Milan 787<br />

Jitraruch, Suttiruk 1711<br />

Jittorntam, Paisan 1999<br />

Jogasuria, Alvin 879, 1324, 1338<br />

Jogunoori, Wilma S 130, 2118<br />

Joh, Jae Won 400<br />

Joh, Takashi 1954<br />

John, Anil 1843<br />

John, Nimy 1037, 1074, 1075<br />

Johnson, Ann D. 2052<br />

Johnson, Casey 57, 1522, 1971<br />

Johnson, Courtney N. 2022<br />

Johnson, Daniel C. 152<br />

Johnson, Heather 648, 1741, 1754<br />

Johnson, Lynt B. 2118<br />

Johnson, Matt 579<br />

Johnson, Nirah 1801<br />

Johnson, Phillip 344, 489, 490, 851<br />

Johnson, Rena D. 521<br />

Johnson, Scott J. 1087, 1143<br />

Johnson, Sindhu 1713<br />

Jokelainen, Kalle 633<br />

Jonas, Maureen M. 136<br />

Jonas, Sven 1222<br />

Jones, Christopher T. 2257<br />

Jones, David 601, 611, 625, 631, 636,<br />

1457, 2226<br />

Jones, Dean P. 19, 2232<br />

Jones, DeAnn 1174<br />

Jones, Elizabeth C. 2144<br />

Jones, Fiona 1321<br />

Jones, Hannah 822<br />

Jones, Rachel E. 1114<br />

Jones, Robert M. 1225<br />

Jong, Simcha 1455<br />

Joo, Sae Kyung 2211<br />

Joo, Seung-Moon 464<br />

Jopson, Laura 611<br />

Jordan, Cynthia E. 1792<br />

Jorge-Ripper, Carlos 1341, 1342<br />

Josbeno, Deborah A. 1244<br />

Joseph, John 2099<br />

Joshi, Shobha 1242<br />

Joshi, Y.K 1340, 2088<br />

Joshi-Barve, Swati 1304, 1331<br />

Jouannaud, Vincent 2082<br />

Jourdan, Tony 1374<br />

Joy, Jeffrey B. 1832<br />

Joyce, Andrew R. 1309<br />

Ju, Hye Lim 504<br />

Juday, Timothy R. 1042, 1068, 1069,<br />

1087, 1139, 1143, 1144, 1152,<br />

1198, 1462<br />

Julian, Schulze zur Wiesch 1629<br />

Julich, Henrike 43, 1299, 1385<br />

Jun, Baek Gyu 484<br />

Jun, Dae Won 924, 941, 951, 1760,<br />

2110, 2218, 2243<br />

Jun, Mi-Jung 446<br />

Jung, Minoa K. 2174<br />

Jung, Yong Jin 685, 1277<br />

Jung, Young Kul 391, 405, 1504, 1897<br />

Jung, Youngmi 696<br />

Jüngst, Christoph 617<br />

Juran, Brian D. 76, 600, 607, 610<br />

Jurek, Marzena 1467, 1478, 1485<br />

Justice, Amy 1488<br />

Kaasis, Mehdi 264<br />

Kabe, Yasuaki 1936<br />

Kabiri, Mina 104, 151<br />

Kachaylo, Ekaterina 665<br />

Kachuwaire, Obert 1575<br />

Kaczmarek, Dominik J. 1287, 2077<br />

Kado, Akira 912<br />

Kadry, Zakiyah 9<br />

Kaestner, Klaus H. 494, 511<br />

Kagawa, Tatehiro 314, 383, 697<br />

Kagay, Christopher R. 2104<br />

Kage, Masayoshi 1690<br />

Kahn, Michael 1945<br />

Kai, Yugo 97, 231, 669, 1082, 1357,<br />

1661, 1985, 2130, 2136<br />

Kaido, Toshimi 1213, 1935<br />

Kaiser, Ralf 2086<br />

Kaiser, Tiffany E. 873<br />

Kaita, Kelly D. 243


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1333A<br />

Kajanachumpol, Saowanee 1999<br />

Kaji, Kiichiro 370<br />

Kaji, Kosuke 677<br />

Kajihara, Mikio 2074<br />

Kajiwara, Atsushi 1941<br />

Kajiwara, Keiko 1375<br />

Kakati, Bobby 1043, 1109, 1224<br />

Kakati, Donny 735, 1220, 1253<br />

Kakazu, Eiji 58<br />

Kakinuma, Sei 164, 406, 1018, 1027,<br />

1959<br />

Kakisaka, Keisuke 1528<br />

Kakiyama, Genta 804<br />

Kakizaki, Satoru 477<br />

Kakolyris, Stylianos 1900<br />

Kakuda, Thomas 39<br />

Kakuda, Yuko 820<br />

Kalabin, Aleksandr 1282<br />

Kalafateli, Maria 2094, 2095<br />

Kalal, Chethan 1340<br />

Kalathil, Sumodh 1121<br />

Kalergis, Alexis 971<br />

Kalia, Harmit S. 943<br />

Kaliamoorthy, Ilankumaran 859<br />

Kalinina, Iryna 2162, 2186<br />

Kalkan, Cagdas 2033<br />

Kallail, K. James 1814, 1815<br />

Kallwitz, Eric R. 1248<br />

Kalmeijer, Ronald 1040<br />

Kalpakou, Georgia 2096<br />

Kalra, Avash 2085<br />

Kalra, Naveen 214<br />

Kalsekar, Anupama 711, 1456, 1461<br />

Kalthoff, Sandra 1927, 2108<br />

Kalwaney, Shirley K. 575<br />

Kamachi, Hirofumi 866<br />

Kamada, Yoshihiro 227, 906, 1642,<br />

1974, 2156, 2207<br />

Kamar, Nassim 3, 1158<br />

Kamath, Binita M. 1694, 1713<br />

Kamath, Patrick S. 6, 82, 281, 282,<br />

293, 570, 1315, 1316, 2089<br />

Kamble, Pravin S. 1050<br />

Kamel, Yasser M. 1147, 1188<br />

Kamimura, Hiroteru 310<br />

Kamiya, Akihide 164, 697, 1369<br />

Kamiya, Kenji 551<br />

Kamiyama, Toshiya 866<br />

Kampa, Nicholas 1718<br />

Kan, Hiromi 986, 1056, 1150, 1649,<br />

1670, 1681, 1858, 1895, 1944, 2120<br />

Kanai, Takanori 181, 569, 2131,<br />

2140, 2234<br />

Kanai, Tomoya 2074<br />

Kanda, Tatsuo 1207, 1628<br />

Kane, Pauline A. 212<br />

Kaneko, Akira 1201, 2213<br />

Kaneko, Shuichi 144, 168, 325, 370,<br />

392, 430, 495, 699, 947, 990, 994,<br />

1000, 1010, 1014, 1348, 1377,<br />

1395, 1632, 1653, 1953, 1968, 1970<br />

Kaneko, Shun 164, 406, 1959<br />

Kaneko, Syun 1018<br />

Kanel, Gary C. 1350<br />

Kanemasa, Kazuyuki 2230<br />

Kaneyama, Yasunari 2097<br />

Kang, Ah-Young 464<br />

Kang, Chaeryon 1598<br />

Kang, Dae Joong 1521<br />

Kang, Daehwan 876<br />

Kang, Jong-Hon 1097, 1778<br />

Kang, Ningling 1947<br />

Kang, Ray 71<br />

Kang, Seong Hee 405, 1897, 2017<br />

Kang, SoHee 1415, 1972<br />

Kang, Sun Woo Sophie 1909<br />

Kani, Satomi 1618<br />

Kanna, Sowjanya 1827<br />

Kantarjian, Hagop 1160<br />

Kanto, Tatsuya 476, 794, 1138, 1632,<br />

1643, 1658, 1795, 2158, 2193<br />

Kanwal, Fasiha 90, 104, 151, 244,<br />

357, 1048<br />

Kanwar, Bittoo 632, 1707<br />

Kao, Chia-Ning 2252<br />

Kao, Jia-Horng 1080, 1105, 1547,<br />

1828, 1996, 2027<br />

Kao, Shih-Chu 2258<br />

Kapalko, Angela 1136<br />

Kapeller, Rosana 957, 1938<br />

Kaplan, David E. 16, 122, 438, 521,<br />

1041, 1302, 1608<br />

Kaplan, Keith 1023<br />

Kaplan, Lee M. 154<br />

Kaplowitz, Neil 590, 1350, 1922,<br />

2135<br />

Kaps, Leonard 1353, 1426<br />

Kapur, Saurabh 748, 769<br />

Kar, Premashis 1596<br />

Karaca, Cetin 478<br />

Karagozian, Raffi 1469<br />

Karakan, Tarkan 932<br />

Karakaya, Fatih 2033<br />

Karani, John B. 212<br />

Karasu, Zeki 246, 558, 1247, 1739,<br />

1822<br />

Karatapanis, Stylianos 1126<br />

Karatayli, Ersin 2033<br />

Karatayli, Senem C. 2033<br />

Kardashian, Ani A. 1273<br />

Karimova, Madina 1629<br />

Karimzadeh, Hadi 31, 1021<br />

Karin, Michael 232, 1971<br />

Karino, Yoshiyasu 707, 1116, 1655,<br />

1778, 2065<br />

Karkada, Joel G. 1099<br />

Karlsen, Tom H. 76, 258, 819, 833<br />

Karnik, Satyajit 632, 1359, 1379, 2149<br />

Karnsakul, Wikrom 280, 294, 560<br />

Karp, Seth J. 837<br />

Karpen, Saul J. 19, 133, 1694<br />

Karpi, Asia 1885, 1886<br />

Karra, Vijay 1596<br />

Karrar, Azza 316, 2173, 2192<br />

Karsdal, Morten A. 119, 441, 780,<br />

1432, 1437, 2178<br />

Karthikeyan, Palaniswamy 1691<br />

Kartoun, Uri 2194<br />

Karvellas, Constantine J. 1759<br />

Kasai, Yosuke 402<br />

Kaseb, Ahmed O. 1160<br />

Kaser, Arthur 258<br />

Kashyap, Amit 1459<br />

Kasirga, Erhun 1623<br />

Kaspar, Mathew 1840<br />

Kasperkovitz, Pia 36<br />

Kassa, Anketse D. 1945<br />

Kasumov, Takhar 330, 651, 667<br />

Katayama, Hokahiro 1976<br />

Katayama, Kazuhiro 1201<br />

Kathpalia, Priya 557<br />

Katlama, Christine 210<br />

Kato, Akinobu 1528<br />

Kato, Junji 1984<br />

Kato, Naoya 505<br />

Kato, Takanobu 985, 1003<br />

Kato, Tomoaki 2129<br />

Kato, Yojiro 2129<br />

Kattakuzhy, Sarah 92, 220<br />

Kattamis, Antonios 1126<br />

Kattentidt-Mouravieva, Anja A. 492<br />

Katz, Barry P. 1316<br />

Katz, Jason 527<br />

Katz, Lior 2118<br />

Kaur, Savneet 954<br />

Kautiainen, Hannu 633<br />

Kawabata, Kenji 1690<br />

Kawada, Norifumi 191, 429, 517, 592,<br />

707, 1125, 1420, 1748, 1824, 2156<br />

Kawaguchi, Kazunori 168, 1653<br />

Kawaguchi, Takumi 233, 918, 1819,<br />

2158, 2207<br />

Kawaguchi, Yasunori 296<br />

Kawai, Takayuki 1976<br />

Kawai-Kitahata, Fukiko 164, 406, 1018,<br />

1027, 1959<br />

Kawakami, Yoshiiku 472, 986, 1056,<br />

1649, 1681, 1893, 1895, 1944<br />

Kawamoto, Hirofumi 2030, 2199<br />

Kawamura, Etsushi 1125, 1824<br />

Kawamura, Satoshi 382<br />

Kawamura, Yusuke 1062, 1191, 2020,<br />

2241<br />

Kawanaka, Miwa 2030, 2199, 2207<br />

Kawanishi, Koki 382<br />

Kawano, Akira 2031<br />

Kawaoka, Tomokazu 472, 986, 1056,<br />

1649, 1681, 1893, 1895, 1944<br />

Kawasaki, Yukihiko 551<br />

Kawashima, Keigo 1640<br />

Kawashima, Minae 1138<br />

Kawata, Kazuhito 1393<br />

Kawazoe, Seiji 296<br />

Kawut, Steven M. 9, 1470<br />

Kaya Kilic, Esra 2051<br />

Kayali, Zeid 702, 1100<br />

Kayandabila, Johnstone 572<br />

Kaymakoglu, Sabahattin 478, 558,<br />

1822<br />

Kazankov, Konstantin 119<br />

Kazerouni, Firouzeh 554<br />

Kazim, Syed N. 1005<br />

Kazimi, Marwan 1274<br />

Ke, Bibo 1283, 2138<br />

Ke, June 2261<br />

Keaveny, Andrew P. 1038, 1124<br />

Kebapcilar, Levent 2206<br />

Kedarisetty, Chandan K. 1328, 2078,<br />

2203<br />

Keefe, Deborah L. 2147, 2208<br />

Keegan, Mathew J. 1243<br />

Kegasawa, Tadashi 26, 453, 1809,<br />

1982<br />

Keiding, Susanne 768<br />

Keim, Sofia 1163<br />

Keirstead, Natalie 36<br />

Keller, Bradley T. 19<br />

Keller, Salome A. 1884<br />

Kelley, Caroline 711<br />

Kelley, Michael J. 540


1334A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Kelley, Robin K. 414, 471<br />

Kellogg, Todd A. 836<br />

Kellum, John A. 275<br />

Kelly, Catherine 931<br />

Kelly, Deirdre A. 1234, 1710, 1728<br />

Kelly, Dympna 850, 1271<br />

Kelly, Erin 1444, 1570<br />

Kelly, Mollie 33, 2064<br />

Kelly, Sean G. 434<br />

Kemgang Fankem, Astrid Donald D. 642<br />

Kemmer, Nyingi M. 526, 1240<br />

Kemp, William W. 442, 483<br />

Kempe, Allison 137, 1738<br />

Kemper, Sherri 192, 1356<br />

Kennedy, Lindsey 190, 621, 812, 822,<br />

835, 1317, 1363<br />

Kennedy, Patrick T. 167, 1571, 1626<br />

Kennedy, Susan M. 911, 1365<br />

Kenneth, Fleming A. 2190<br />

Kenny, Thomas P. 77<br />

Kent, Robert N. 1366<br />

Keppel, Jessica A. 1215<br />

Kerbert, Annarein J. 2069<br />

Kerbouche, Rafik 1621<br />

Kerkar, Nanda 133, 1694<br />

Kerlan, Robert 471<br />

Kersey, Kathryn 1076<br />

Kershaw, Erin E. 1531<br />

Kershenobich, David 2223<br />

Keskin, Onur 2012, 2033<br />

Keum, Bora 1438, 1504<br />

Khairy, Marwa 1076, 1163<br />

Khaitova, Viktoriya 1095, 1185<br />

Khalid, Hamza 1535<br />

Khalili, Korosh 348, 542<br />

Khalili, Mandana 306<br />

Khan, Aamir G. 1872<br />

Khan, Bilal 2047<br />

Khan, Harun 1826<br />

Khan, Mohammad Qasim 1791<br />

Khan, Rashid 281<br />

Khanam, Arshi 1762<br />

Khandelwal, Niranjan 214<br />

Khanna, Rajeev 1697, 1723<br />

Kharbanda, Kusum K. 589, 1323, 1347<br />

Khatri, Amit 1039, 1088<br />

Khattar, Ramzi 1012, 1297<br />

Khawaja, Shazia 717, 729<br />

Khawar, Sabrina 764<br />

Khemichian, Saro 217<br />

Khera-Butler, Tanya 113<br />

Khillan, Vikas 2070<br />

Khlaiphuengsin, Apichaya 2036<br />

Kho, Abel 563, 849<br />

Kholmukhamedov, Andaleb 1925<br />

Khoo, Mui Joo 1812<br />

Khoo, Saye H. 1103<br />

Khor, Seik-Soon 1138<br />

Khorzad, Rebeca 847<br />

Khosla, Ritu 695<br />

Khullar, Vikas 839<br />

Khungar, Vandana 864<br />

Khurana, Sandeep 1911<br />

Ki, Mo-ran 2113<br />

Kiciak, Slawomir 1582<br />

Kida, Akihiko 370<br />

KIda, Sachiho 906<br />

Kienle, Eike 98<br />

Kijsirichareanchai, Kunut 748, 769<br />

Kikuchi, Luciana 48, 380, 438, 457,<br />

469<br />

Kikuchi, Masahiro 2177<br />

Kikuchi, Miho 2177<br />

Kil, Natalie B. 1792<br />

Kilisinska, Natalia 1502<br />

Kim, Jung Hee 422, 485, 1960, 2049<br />

Kim, Ahyoung 1894<br />

Kim, Arthur Y. 1034, 1094<br />

Kim, Beom Kyung 346, 1542, 1558,<br />

1581<br />

Kim, Bo Hyun 377, 486, 2113<br />

Kim, Boo Sung 277, 484<br />

Kim, Boyoung 2116<br />

Kim, Brian 1151<br />

Kim, Byung Ik 1624, 1992<br />

Kim, Byung Seok 2038<br />

Kim, Chang Duck 1504<br />

Kim, Chang Wook 1542<br />

Kim, Chang-Min 486<br />

Kim, Chunki 920, 925<br />

Kim, Connie 272<br />

Kim, Dayoung 504<br />

Kim, Do Young 346, 504, 1542, 1558,<br />

1581<br />

Kim, Dong Joon 736, 1515<br />

Kim, Donghee 88, 1208, 2155<br />

Kim, Eui Joo 1896<br />

Kim, Eun Sun 1504<br />

Kim, Grace 272, 1454<br />

Kim, Gyeonghwa 403, 426, 749<br />

Kim, Haeryoung 1902<br />

Kim, Hee Yeon 1542<br />

Kim, Hong Joo 1624, 1992<br />

Kim, Hong Soo 277, 484<br />

Kim, Hwa Jung 2155<br />

Kim, Hwi Young 1277<br />

Kim, Hyeyoung 1277<br />

Kim, Hyo Jin 2049<br />

Kim, Hyun Joo 1352<br />

Kim, Hyung-Don 446<br />

Kim, Hyungwook 876<br />

Kim, Irene K. 1479<br />

Kim, Ja Kyung 464, 1399<br />

Kim, Jae Keun 464<br />

Kim, Jae-Sung 177<br />

Kim, Jeong Min 2038<br />

Kim, Ji Hoon 391, 405, 1504, 1897,<br />

2017<br />

Kim, Ji Young 577<br />

Kim, Ji-Hee 1401<br />

Kim, Jieun 696<br />

Kim, Jihoon 112<br />

Kim, Jin-Wook 389<br />

Kim, Jong Man 400<br />

Kim, Jonghwa 1415, 1972<br />

Kim, Ju Hyun 391, 1896<br />

Kim, Jun Won 464<br />

Kim, Jung-Hee 502, 1419<br />

Kim, Kang Mo 446, 448<br />

Kim, Karen E. 579<br />

Kim, Kyoungmi 606<br />

Kim, Kyung Hee 486<br />

Kim, Kyung-ah 2113<br />

Kim, Kyunga 422<br />

Kim, Kyungpil 241, 1678, 2004, 2043,<br />

2059<br />

Kim, Mi Na 1605<br />

Kim, Miran 256, 1962<br />

Kim, Misung 2227<br />

Kim, Moon Young 736, 1322, 1515,<br />

2080<br />

Kim, Myung-Ho 1358<br />

Kim, Nathan G. 387, 428<br />

Kim, Rosa S. 1392<br />

Kim, Sang 988, 1289<br />

Kim, Sang Geon 1352<br />

Kim, Sang Gyune 277, 484, 736,<br />

1542, 1605<br />

Kim, Sangbae 2118<br />

Kim, Seong Hoon 400, 862<br />

Kim, Seong-Jun 968<br />

Kim, Seung Up 346, 363, 1542, 1558,<br />

1581, 1605<br />

Kim, Shin Hee 277<br />

Kim, So Yeon 910, 1358<br />

Kim, Soo-Jin 1401<br />

Kim, Soon Sun 379, 1567, 2056<br />

Kim, Sung Min 1419<br />

Kim, Sunmi 277<br />

Kim, Tae Hyun 1352<br />

Kim, Tae Hyung 433, 1438, 1504<br />

Kim, Tae Suk 405, 1897, 2017<br />

Kim, Tae Yeob 736<br />

Kim, W. Ray 6, 75, 88, 381, 415, 546,<br />

562, 1182, 1189, 1208, 1813, 2155<br />

Kim, Won 1277, 2211<br />

Kim, Yong Ook 183, 950, 955, 1353,<br />

1376, 1385<br />

Kim, Yoon Jun 242, 400, 447, 470,<br />

474, 479, 1610, 1907, 2016, 2042,<br />

2055<br />

Kim, Yoona 207<br />

Kim, Young Don 277, 484, 1515<br />

Kim, Young Seok 277, 484, 736, 1542,<br />

1605<br />

Kim, Yun Ju 218<br />

Kim, Yun Soo 391, 1896<br />

Kim-Schluger, Leona 1185<br />

Kimura, Akihiko 1696<br />

Kimura, Kiminori 476<br />

Kimura, Koichi 1983<br />

Kimura, Minoru 1369<br />

Kimura, Mutuumi 1116<br />

Kimura, Naruhiro 310<br />

Kimura, Osamu 2167<br />

Kimura, Shoko 338<br />

King, Jennifer 708<br />

King, Jesse A. 630<br />

Kini-Matondo, Willy 1102<br />

Kinikli, Sami 2051<br />

Kinoshita, Manabu 324<br />

Kinoshita, Masahiko 1748<br />

Kipp, Benjamin R. 169<br />

Kiraithe, Muthamia 1666<br />

Kirby, Brian 1128, 1707<br />

Kirchner, Gabi I. 76<br />

Kirkness, Carmen S. 774<br />

Kirkpatrick, James N. 1218<br />

Kirkpatrick, Robert B. 874, 1260, 1280<br />

Kirpich, Irina 114<br />

Kirschner, Janina 1002<br />

Kirstein, Martha 444<br />

Kiser, Jennifer 1094<br />

Kishi, Hiroyuki 1599<br />

Kishino, Kyohei 466, 770, 781, 1533,<br />

1669<br />

Kisiel, John B. 169<br />

Kiso, Shinichi 227, 1974<br />

Kisseleva, Tatiana 186, 187, 189, 890


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1335A<br />

Kissiedu, Juliana 840<br />

Kita, Sadahiko 1976<br />

Kitade, Mitsuteru 689<br />

Kitagawa, Noriyuki 869<br />

Kitagawa, Yuko 336<br />

Kitago, Minoru 336<br />

Kitahara, Masaaki 370, 392<br />

Kitamura, Kazuya 1348<br />

Kitamura, Naoto 1978, 2143<br />

Kitazono, Masaki 456<br />

Kitrinos, Kathryn M. 243, 1551, 1557,<br />

1563, 1594, 1678, 2004, 2021<br />

Kitsahawong, Bubpha 555<br />

Kittitrakul, Chatporn 555<br />

Kitzman, Geoffrey 1178<br />

Kiyono, Soichiro 1490, 1516<br />

Kiyota, Ryousuke 1584<br />

Kjærgaard, Maria 773<br />

Klaassen, Curtis 655<br />

Klair, Jagpal S. 1099<br />

Klar, Ernst 1227<br />

Klarenbeek, Paul L. 637<br />

Klassen, Carolyn 647<br />

Klatte, Ryan 840<br />

Klebanoff, Matthew 154<br />

Klee, Eric W. 1965<br />

Klein, Andrew S. 1479<br />

Klein, Niklas 43, 1299, 1385<br />

Klein, Sabine 80, 961<br />

Klein, Thomas 955, 1376<br />

Kleiner, David E. 157, 166, 218, 597,<br />

779, 1498, 1646, 1717, 1720, 1722,<br />

1923, 1926, 2144, 2150, 2220<br />

Kleinewietfeld, Markus 1731<br />

Kleinstein, Sarah E. 2037<br />

Klenerman, Paul 1298<br />

Klepper, Arielle L. 229, 988, 1289<br />

Kliethermes, Stephanie 532, 559<br />

Klinck, John R. 1216<br />

Klinker, Hartwig 252, 1058, 1559<br />

Klintman, Daniel 308<br />

Klopfer, Stephanie 210, 712<br />

Klumpp, Klaus 33, 2064<br />

Kluwe, Johannes 2195<br />

Kneiber, David 57<br />

Knight, Virginia H. 47, 442, 483<br />

Knighton, Sarah 646, 1113, 1744<br />

Knisely, A. S. 1721<br />

Knoespel, Fanny 1921<br />

Knöpfli, Marina 425<br />

Knox, Steven J. 91, 1076, 1080, 1105,<br />

1120, 1168<br />

Knudsen, Sanne M. 978<br />

Ko, Eun Jeong 377<br />

Koay, Evelyn S. 1812<br />

Kobashi, Gen 551<br />

Kobayashi, Akira 869<br />

Kobayashi, Ichizo 1809<br />

Kobayashi, Mariko 1062, 1191, 2020<br />

Kobayashi, Masahiro 364, 1062, 1191,<br />

2020, 2241<br />

Kobayashi, Tomoki 472, 1056, 1893<br />

Kobayashi, Yoshimasa 1393<br />

Kobayashi, Yoshinao 584, 1500<br />

Kobayashi, Yoshiyuki 1047<br />

Kobazi Ensari, Gokce 831, 2115<br />

Kobune, Masayoshi 1984<br />

Koc, Özgür M. 1613<br />

Koch, Sandra 376<br />

Kocher, Hemant 101<br />

Koda, Masahiko 1499, 2199<br />

Kodali, Sudha 735<br />

Kodama, Kazuhisa 1326<br />

Kodama, Nobuhiro 2097<br />

Kodama, Takahiro 231, 234, 669<br />

Kodys, Karen 109, 180, 496, 512,<br />

1016, 2127, 2198<br />

Koek, Ger H. 922, 1613<br />

Koelfat, Kiran V. 604<br />

Koenecke, Christian 2128<br />

Koenig, Aaron B. 891, 2212<br />

Koga, Hironori 688, 918, 1819, 1962,<br />

1990<br />

Kogiso, Tomomi 1326, 2207<br />

Kogure, Takayuki 1026, 1648, 1943,<br />

2167<br />

Koh, Anna 364, 2174<br />

Koh, Christopher 218, 779, 1498,<br />

1650, 1717, 1720, 2144<br />

Koh, Hong 195<br />

Koh, Kwang Cheol 422, 423, 485,<br />

1591, 1960, 2049<br />

Koh, Sarene 167<br />

Kohgo, Satoru 1655<br />

Kohgo, Yutaka 927<br />

Kohjima, Motoyuki 1157, 1159<br />

Kohli, Anita 92, 220, 1111, 1155,<br />

1164<br />

Kohli, Rohit 159, 160, 916<br />

Koido, Shigeo 2074<br />

Koike, Kazuhiko 461, 912, 1138, 1632<br />

Koike, Kazuko 339, 2236<br />

Koizumi, Jun 383<br />

Koizumi, Wasaburo 482, 756<br />

Koizumi, Yohei 788, 1293<br />

Kojima, Kentaro 382<br />

Kojima, Sei-ichiro 383<br />

Kojima, Soichi 657<br />

Kokordelis, Pavlos 1287<br />

Koksal, Aydin 846<br />

Koksal, Iftihar 2057<br />

Kolachala, Vasantha L. 12<br />

Kolb Nava, Casey M. 1831<br />

Kolbeck, Kenneth J. 481<br />

Kolli, K. Pallav 471<br />

Kolly, Philippe 425<br />

Koloda, Dmitry 721<br />

Komar, Michael 552<br />

Komatsu, Nobutoshi 1565<br />

Komolmit, Piyawat 360, 1631, 1662<br />

Komori, Atsumasa 408, 783, 1600<br />

Komorizono, Yasuji 456<br />

Komuta, Mina 1734<br />

Kon, Kazuyoshi 226, 917, 1771<br />

Kondili, Loreta A. 1802, 1844<br />

Kondo, Chisa 291<br />

Kondo, Hisayoshi 952<br />

Kondo, Mario 1484, 1494, 1495<br />

Kondo, Mayuko 461<br />

Kondo, Mitsumasa 1970<br />

Kondo, Reiichiro 465<br />

Kondo, Takayuki 1490, 1516<br />

Kondo, Yasuteru 1026, 1648, 1943,<br />

1983, 2167<br />

Kondo, Yuichi 1420<br />

Kondo, Yuji 461<br />

Konerman, Monica 576<br />

Koneru, Alaya 1787<br />

Kong, Bo 919, 934<br />

Kong, Fei 1149, 2018<br />

Kong, Wenli 2107<br />

Kong, Yuanyuan 374<br />

Kongsomgan, Anucha 1999<br />

Konidari, Anastasia 1702<br />

Konijn, Cynthia 68<br />

Koning, Ludi 1196<br />

Kono, Hiroshi 1956, 2125<br />

Konsten, Sarah 1613<br />

Kontos, Georgios 2019<br />

Koo, Imhoi 926<br />

Koola, Jejo D. 529<br />

Koorey, David J. 1243<br />

Koot, Maarten 1995<br />

Kootstra, Neeltje A. 1633, 2005, 2255,<br />

2263<br />

Kopanja, Dragana 1988<br />

Kopec, Anna K. 81<br />

Koral, Kelly 1979<br />

Korayem, Enas 733<br />

Korenaga, Keiko 794, 1795, 2158,<br />

2193<br />

Korenaga, Masaaki 794, 1138, 1658,<br />

1795, 2158, 2193<br />

Korkmaz, Huseyin 2206<br />

Kormos Hallberg, Csilla 2100<br />

Kornek, Miroslaw 43<br />

Kort, Jens 41, 248, 250, 705, 710,<br />

719, 723<br />

Korzun, William J. 1776<br />

Kosaka, Hisashi 1420<br />

Koser, Martin 510, 2116<br />

Koskinas, Ioannis 1126<br />

Kosloski, Matthew P. 710, 723<br />

Kosters, Astrid 19<br />

Kostev, Karel 1052<br />

Kota, Janaiah 436<br />

Koteish, Ayman A. 607<br />

Kotoh, Kazuhiro 1159<br />

Kottilil, Shyam 92, 220, 578, 1013,<br />

1029<br />

Kou, Xiaoni 1773<br />

Koudou, Kazuki 901<br />

Kouides, Peter 1034<br />

Koukoulioti, Eleni 1638<br />

Kountouras, Dimitrios 1126<br />

Kourakli, Alexandra 1126<br />

Kourikou, Anastasia 2012, 2019<br />

Kouroumalis, Elias A. 2094<br />

Koury, Kenneth J. 700<br />

Koutroumpakis, Efstratios 2095<br />

Kouyama, Jun-ichi 1064<br />

Kouyos, Roger 1818<br />

Kouznetsova, Maria A. 528<br />

Kovacs, Andrea 1444<br />

Kovacs, Steven J. 2261<br />

Kowdley, Kris V. 73, 107, 239, 607,<br />

616, 628, 634, 946, 1046, 1106,<br />

1108, 1120, 2153, 2154, 2161, 2220<br />

Kowgier, Matthew 174, 348, 350,<br />

1099, 1556<br />

Koyama, Yukinori 189, 1522<br />

Koyanagi, Toshimasa 1159<br />

Kozak, Ladislav 622<br />

Kozbial, Karin 297, 732, 1070, 1092,<br />

1137, 1885<br />

Kozono, Masaya 937, 1421<br />

Kozuka, Ritsuzo 1125<br />

Kraft, Anke R. 2026<br />

Krag, Aleksander 119, 773, 780, 789,<br />

792, 1440


1336A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Krajden, Mel 86, 543<br />

Kramer, Jennifer R. 156, 244, 1048<br />

Krämer, Benjamin 1008, 1287, 2077<br />

Krams, Sheri M. 341, 2123<br />

Kranidioti, Hariklia 2019<br />

Krasnewich, Donna 1722<br />

Krastev, Zahary 2059<br />

Krattinger, Regina 519, 520<br />

Krauel, Alexander 245, 1594<br />

Kraus, Michael R. R. 1170, 1200<br />

Krawczyk, Marcin 43, 1429, 2195<br />

Krawczyk, Sarah 2227<br />

Kreefft, Kim 1007<br />

Krenkel, Oliver 1756<br />

Krenzien, Felix 1222<br />

Kresge, Charles 815<br />

Kretzschmar, Georg 617<br />

Kreutzfeltd, Martin 301<br />

Krishnamurthy, Venkat 418, 2116<br />

Krishnan, Anuradha 825, 1994<br />

Krisko, Tibor I. 881, 900<br />

Kriss, Michael 503<br />

Krittanawong, Chayakrit 282<br />

Krohn, Sandra 261, 276<br />

Krok, Karen 1470<br />

Krokowski, Dawid 668<br />

Kronborg, Ian 442<br />

Kronenberger, Bernd 1055<br />

Krones, Elisabeth 2068<br />

Kronmann, Karl 1175<br />

Krowka, Michael J. 9, 1470<br />

Kroy, Daniela C. 963<br />

Kruger, Annie 25<br />

Kruglov, Emma A. 661<br />

Krull, Mona 1708<br />

Ku, Hye Jin 346, 1558<br />

Ku, Lawrence 1185<br />

Ku, Winston 1182<br />

Kuate, Seraphin 200<br />

Kubik, Jacy L. 1969, 2139<br />

Kubo, Shoji 429, 1748, 1983<br />

Kuchimanchi, Satya 36<br />

Kuchipudi, Aishwarya 307<br />

Kucirka, Lauren M. 1749<br />

Küçükay, Fahrettin 846<br />

Kudo, Masatoshi 851, 1935<br />

Kudo, Yotaro 461<br />

Kuehne, Felicitas 1089<br />

Kugelmas, Marcelo 1040<br />

Kugiyama, Yuki 408, 783, 1600<br />

Kuiken, Sjoerd 1995<br />

Kuk, Nathan T. 945<br />

Kulkarni, Supriya 23<br />

Kullak-Ublick, Gerd A. 225, 519, 520<br />

Kuller, Lewis 2228<br />

Kumada, Hiromitsu 364, 707, 1062,<br />

1191, 2020, 2241<br />

Kumada, Takashi 344<br />

Kumagai, Erina 794, 1795, 2158,<br />

2193<br />

Kumagai, Kotaro 799, 1421, 1838<br />

Kumagai, Masashi 370<br />

Kumagai, Takanori 296<br />

Kumagi, Teru 524, 634<br />

Kumar, Aditi 1234<br />

Kumar, Ambuj 1240<br />

Kumar, Anupam 690, 694, 1483<br />

Kumar, Avinash 146, 262, 330, 334,<br />

518, 651, 667, 668, 671, 1308, 2088<br />

Kumar, Dhananjay 690, 694, 1483<br />

Kumar, Divya P. 807<br />

Kumar, Guresh 1328, 1340, 1344,<br />

1663, 2092, 2203<br />

Kumar, Manoj 1328, 1659, 2070,<br />

2203<br />

Kumar, Naveen 2245<br />

Kumar, Niteen 1268<br />

Kumar, Princy N. 1049<br />

Kumar, Sachin 2092<br />

Kumar, Vipin 1345<br />

Kumari, Radhka 1182, 1189<br />

Kumer, Sean C. 352, 565<br />

Kummen, Martin 833<br />

Kumral, Dennis 13<br />

Kunasegaran, Kamini 167<br />

Kuniholm, Mark H. 1444<br />

Kunimoto, Hideo 1191, 2020, 2241<br />

Kuno, Atsushi 1448<br />

Kunos, George 1374<br />

Kuo, Alexander 94, 1057<br />

Kuo, Margot E. 86, 543<br />

Kuo, Paul C. 2143<br />

Kuo, Yong-Fang 1249<br />

Kupiec-Weglinski, Jerzy 1283, 2138<br />

Kurahashi, Tomohide 227, 1974<br />

Kuramitsu, Tomoyuki 1097<br />

Kurimoto, Ami 1420<br />

Kurioka, Ayako 1298<br />

Kuroda, Hidekatsu 1528<br />

Kuroda, Shunichi 1642<br />

Kuromatsu, Ryoko 1819<br />

Kurosaki, Masayuki 1115, 1192, 1632,<br />

2029<br />

Kurosawa, Takao 1696<br />

Kurt, Ramazan 1096<br />

Kurt-Jones, Evelyn A. 180<br />

Kurtz, Tracie L. 138<br />

Kuruvilla, Asha 1513<br />

Kusakabe, Atsunori 1097<br />

Kusama, Hiromi 226, 917<br />

Kusano, Hironori 465, 1990<br />

Kusanovic, Juan P. 78<br />

Kuscuoglu, Deniz 2115<br />

Kushner, Tatyana 1608<br />

Kutay, Huban 99<br />

Kutsenko, Alina 359, 1806<br />

Kuwahara, Reiichiro 304, 681, 1819<br />

Kuwaki, Kenji 2236<br />

Kuwata, Yasuaki 1116<br />

Kuzuya, Teiji 411, 443, 880, 967,<br />

1061, 1167<br />

Kwak, Joo Hee 1760, 2110, 2243<br />

Kwanten, Wilhelmus J. 894, 976<br />

Kwee, Sandi 1553<br />

Kweon, Young Oh 363, 403, 426, 749,<br />

2028, 2038<br />

Kwo, Paul Y. 42, 217, 248, 700, 702,<br />

703, 712, 1084, 1903<br />

Kwok, Ryan M. 1101<br />

Kwon, Hyunjoo 686<br />

Kwon, Oh Sang 391, 1896<br />

Kwon, Sang Ok 2080<br />

Kwon, Seong Uk 857<br />

Kyritsi, Tina 190, 817, 1317<br />

Kyvernitakis, Andreas 1030, 1160<br />

L’Italien, Gilbert 1828<br />

La Rosa, Katia Yu 222<br />

Labadie, Hélène 264<br />

Labarga, Pablo 1881, 1908<br />

Labreuche, Julien 1269<br />

Labriola, Raffaella 983<br />

Labropoulou-Karatza C, Chrisoula 2094,<br />

2095<br />

Lacerda, Marco A. 1903<br />

LaCerte, Carl 1915<br />

Lachlan, Neil J. 260, 753<br />

Lacombe, Karine 1676, 1677, 1833<br />

Lacoste, Benoit 179<br />

LaCreta, Frank 720, 728<br />

Ladenheim, Maya R. 387, 428<br />

Ladner, Daniela 71, 563, 843, 847,<br />

849<br />

Laengvejkal, Pavis 280<br />

Lafdil, Fouad 693<br />

Laferl, Hermann 1885<br />

Lafoz, Erica 1518<br />

Lages, Celine S. 823, 1695<br />

Laggetta, Maristella 1493<br />

Lahiri, Pooja 831<br />

Lahoti, Manohar M. 1784<br />

Lai, Chengjung 510, 2116<br />

Lai, Ching-Lung 126, 1120, 1545,<br />

1550, 1746<br />

Lai, Hsueh-Chou 2040<br />

Lai, Jennifer B. 1053<br />

Lai, Jennifer C. 72, 557, 841, 844,<br />

852, 1219, 1468<br />

Lai, Jiaming 454<br />

Lai, Michelle 1484, 1494, 1495, 2160,<br />

2196, 2197, 2227, 2228<br />

Lai, Paul B. 344<br />

Lai, Sara 309<br />

Laitinen, Tarja 225<br />

Lakatos, Isobel 1707<br />

Lakdawalla, Darius 1042, 1068<br />

Lake, John R. 1184, 2006<br />

Lal, Bikrant B. 1697, 1725<br />

Laleman, Wim 171, 1091<br />

Lalezari, Jacob P. 41, 210, 1065<br />

Laliena, Almudena 1398<br />

Lalley-Chareczko, Linden 1136<br />

Lam, Angela 33, 2064<br />

Lam, Brian P. 892, 1186<br />

Lam, Christina 1722<br />

Lam, Hung 2175<br />

Lam, Karen S. L. 898<br />

Lam, Katherine 1100<br />

Lamb, Cheri L. 1403<br />

Lambiase, Lara 222<br />

Lamblin, Géraldine 1505<br />

Lamerato, Lois 1578<br />

Lammers, Willem J. 73, 601, 634<br />

Lammert, Frank 445, 463, 617, 1166,<br />

1429, 2077, 2086, 2195, 2253<br />

Lampe, Guy 784<br />

Lampertico, Pietro 1747, 2011, 2012<br />

Lan, Tian 890<br />

Landas, Steve 1961<br />

Landeen, Lee K. 1770<br />

Landen, Michiel 976<br />

Landini, Maria Paola 1888<br />

Landis, Charles S. 706, 716, 726,<br />

1049, 1211, 1478<br />

Landsittel, Doug 1244<br />

Lane, Suzanne 1139, 1152<br />

Lanford, Robert E. 32, 2022<br />

Lang, Anna L. 2141<br />

Lang, Julia 1666<br />

Lange, Undine G. 1222<br />

Langedijk, Jacqueline 827


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1337A<br />

Langhans, Bettina 1008, 2077<br />

Langhi, Cedric 65<br />

Langiewicz, Magda 499<br />

Langlet, Philippe 1091<br />

Langworthy, James 766<br />

Lankisch, Tim 1002<br />

Lanti, Claudia 2187<br />

Lanza, Marika 1654<br />

Lanzoni, Valéria P. 2189<br />

Lapalus, Martine 998, 1601, 2044,<br />

2045<br />

Lapierre, Pascal 305<br />

Lapinski, Tadeusz W. 1502<br />

Lapuyade, Bruno 450<br />

Laraque, Fabienne 1801<br />

Larrey, Dominique G. 206, 225, 1847<br />

Larsen, Lars P. 768<br />

Larsen, Lois M. 1107<br />

Larson, Colin 503<br />

Larson, Joseph J. 6<br />

Larsson Cohen, Marita S. 510, 2116<br />

LaRusso, Nicholas F. 73<br />

LaSalle, Janine M. 2109<br />

Lascoux-combe, Caroline 1676, 1677<br />

Laskin, Debra 934<br />

Lassailly, Guillaume 1269, 2222<br />

Lasser, Luc 1091<br />

Latinne, Dominique 1734<br />

Latour, Veronique 1796<br />

Lattanzi, Barbara 2067<br />

Latz, Eicke 100<br />

Lau, Audrey H. 2123<br />

Lau, Daryl 371<br />

Lau, George K. 1536, 1537, 2008<br />

Lau, Vincent 33, 2064<br />

Lauer, Georg 332<br />

Lauer, Ulrich M. 1183<br />

Lauletta, Gianfranco 435<br />

Laura, Alvarez 1919<br />

Laurent, Alexis 421, 449, 1232<br />

Lauridsen, Mette M. 1472<br />

Lauriski, Shannon 1849<br />

Lavanchy, Daniel 1818<br />

Lavert, Marion 999<br />

Lavieri, Mariel S. 67<br />

Lavine, Joel E. 159, 2161<br />

Lavrut, Pierre-Marrie 195<br />

Law, Matthew 1868<br />

Lawitz, Eric 39, 42, 205, 219, 247,<br />

700, 702, 703, 712, 718, 726, 1039,<br />

1040, 1120, 1128, 1428, 1433,<br />

1435, 2257, 2260<br />

Lawler, Joseph 1151<br />

Lawson, Mark 175, 915, 1916<br />

Layden, Thomas J. 1248<br />

Layese, Richard 362, 1808, 1847<br />

Lazaridis, Konstantinos 76, 600, 607,<br />

610<br />

Lazaro, Leander 1425<br />

Lazarus, Arnaud 1194<br />

Le, An K. 207, 359, 368, 387, 388,<br />

428, 1561, 1806, 1816, 1848<br />

Le, Michael H. 367, 459<br />

Le, Richard H. 359, 367, 386, 388,<br />

458, 459, 1806, 1816, 1848<br />

Le, Thuy T. 191, 1824<br />

Le Bail, Brigitte 1439, 2181<br />

Le Bert, Nina 167<br />

Le Cleach, Aline 1997<br />

Le Grand, Béatrice 30<br />

Le Hello, Rolland 1796<br />

Le Marchand, Chloe 1793<br />

Leaf, David 1488<br />

Leal, Carlos 2084<br />

Leal Tassias, Aránzazu 691<br />

Learned, Marc 198, 1986<br />

Leary, Thomas 1551<br />

Leathers, James 1223<br />

Leber, Bettina 2068<br />

Leber, Werner 1052<br />

Lebofsky, Margitta 586<br />

Lebray, Pascal 3, 95, 206, 786, 1239<br />

Lebrec, Didier 145, 265, 772, 1288,<br />

2148<br />

Lebreton, Aurelien 1505<br />

Lebrilla, Carlito B. 606<br />

Lebuffe, Gilles 1269<br />

Leclair, Katherine B. 881, 900<br />

Leclercq, Isabelle A. 178, 678<br />

LeCuyer, Brian E. 658<br />

Lee, Amy C. 2007, 2034, 2062<br />

Lee, Andy Hung-Yi 71, 847<br />

Lee, Brian P. 1214, 1215<br />

Lee, Chang Hee 1897<br />

Lee, Chang Hyeong 2038<br />

Lee, Christine K. 136<br />

Lee, Chuan-Mo 1873, 2010<br />

Lee, Danbi 446, 448<br />

Lee, David D. 50, 53, 335<br />

Lee, Donghyeon 242, 400, 447, 474,<br />

479, 1610, 2016, 2042, 2055<br />

Lee, Eun Byul 502, 1419<br />

Lee, Eung Chang 862<br />

Lee, Fa-Yauh 738, 751<br />

Lee, Guan Huei 1686<br />

Lee, Hae Lim 452, 1882<br />

Lee, Hae Won 1277<br />

Lee, Han Chu 446, 448<br />

Lee, Hannah 2015, 2052<br />

Lee, Helen 32, 2022<br />

Lee, Heon Ju 2038<br />

Lee, Hong Sik 1504<br />

Lee, Hye Won 1558, 1581<br />

Lee, Hyejung 1902<br />

Lee, I-Cheng 439<br />

Lee, Ik Jae 464<br />

Lee, Jai Sun 941<br />

Lee, Jeong Min 474<br />

Lee, Jeong-Hoon 242, 400, 447, 470,<br />

474, 479, 1610, 1907, 2016, 2042,<br />

2055<br />

Lee, Jin-Woo 1542<br />

Lee, Joon Ho 1419<br />

Lee, Joon Hyeok 422, 423, 485, 1591,<br />

1960, 2049<br />

Lee, Ju-Seog 129, 2118<br />

Lee, June Sung 242, 400, 447, 470,<br />

479, 1907, 2016<br />

Lee, Jung Il 464, 1399, 1542<br />

Lee, Kuei-Chuan 757, 965, 2050<br />

Lee, Kwan Sik 464, 1399, 1542<br />

Lee, Kwang-Hun 464<br />

Lee, Kwang-Woong 400, 474, 1277,<br />

1278<br />

Lee, Kwangwon 879, 1324, 1338<br />

Lee, Major 521<br />

Lee, Mei-Hsuan 1593, 1604, 1828<br />

Lee, Mikang 920, 925<br />

Lee, Minjong 242, 400, 447, 474, 479,<br />

1610, 2016, 2055<br />

Lee, Richard 78<br />

Lee, Sae Hwan 277, 484, 991, 1630,<br />

1644<br />

Lee, Samuel S. 1051, 1529<br />

Lee, Sangbin 436<br />

Lee, Sangmin 1336<br />

Lee, Serene M. 520<br />

Lee, Seulki 115<br />

Lee, Seung Duk 862<br />

Lee, Seung Min 924, 941, 1760, 2110,<br />

2218, 2243<br />

Lee, Soonkyu 452<br />

Lee, Sooyeon 177<br />

Lee, Su Hyun 363, 403, 426, 749<br />

Lee, Sunyoung 1783<br />

Lee, Taeheon 1624<br />

Lee, Teng-Yu 345<br />

Lee, Tingfang 1364<br />

Lee, Tzu-Hao 1887<br />

Lee, William M. 99, 211, 213, 1555,<br />

1759, 1761, 1763, 1766, 1768,<br />

1774, 1776, 1777, 1914<br />

Lee, Yoonkwang 920, 925<br />

Lee, Young-Sun 405, 1897, 2017<br />

Lee, Youngmin A. 1364<br />

Lee, Yu Rim 363, 403, 426, 749<br />

Lee, Yun Bin 2042, 2055<br />

Lee, Yung Sang 446, 448<br />

Leeflang, Mariska M. 74, 623<br />

Leeming, Diana J. 119, 441, 780, 1432<br />

Lefeber, Dirk 2102<br />

Lefebvre, Eric 1756, 2247<br />

Lefèvre, Mathieu 1019<br />

Leffondre, Karen 132<br />

Lefkowitch, Jay H. 2129<br />

Leggett, Barbara A. 106, 627<br />

Legrand, Margaux 1409<br />

Lehar, Sophie M. 2009<br />

Lehmann, Patrick 1002, 1572, 1597<br />

Lehretz, Maria 2012<br />

Leipertz, Steven 128<br />

Leise, Michael D. 1038, 1124<br />

Lemasson, Matthieu 997<br />

Lemasters, John J. 319, 1925<br />

Lemesch, Sandra 2068<br />

Lemmers, Arnaud 178<br />

Lemoinne, Sara 642<br />

Lemon, Stanley M. 990, 1010<br />

Lenci, Ilaria 203, 1751<br />

Lenders, Marie-Hélène 922<br />

Lenicek, Martin 604, 1487<br />

Lennon, Michael 1491<br />

Lenoci, Maryanne 1133<br />

Lens, Sabela 982<br />

Lenz, Oliver 1040, 1179<br />

Lenzen, Henrike 76<br />

Leonard, Jennifer 204<br />

Leong, Jennifer 1095, 1185<br />

Lepida, Antonia 2079<br />

Lerner, Daniel J. 531, 550<br />

Leroy, Vincent 3, 95, 117, 206, 777,<br />

1158, 1439, 1808, 1860<br />

Lesgourgues, Bruno 1072, 2082<br />

Letenneur, Luc 132<br />

Leteurtre, Emmanuelle 2222<br />

Letoublon, Christian 421, 449<br />

Letouze, Eric 362<br />

Leung, Patrick S. 77, 602, 606, 613,<br />

619, 639<br />

Levander, Sepideh 992


1338A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Leveille-Webster, Cynthia 582<br />

Leventhal, Thomas M. 1777<br />

Levesque, Eric 215, 1269<br />

Levi, David 17<br />

Levi, Moshe 139, 142, 907, 959<br />

Levin, Douglas M. 874, 1280<br />

Levine, Matthew H. 868<br />

Levine, Vanessa 731<br />

Levitsky, Josh 5, 54, 1235<br />

Levitt, Brian S. 1549<br />

Levrero, Massimo 165, 958<br />

Levstik, Alexander 2117<br />

Levy, Adrian 1152<br />

Levy, Cynthia 76, 616, 624, 1054,<br />

1148, 1206<br />

Levy, Gahl 331, 672, 679<br />

Levy, Gary 11, 424, 1012, 1297<br />

Lewis, David L. 32, 2022<br />

Lewis, Dylan 212<br />

Lewis, James D. 865<br />

Lewis, James H. 286<br />

Li, Lei 1934<br />

Li, Xiaojiao 2058<br />

Li, Biao 1668<br />

Li, Bing 1536, 1537, 2008<br />

Li, Changyong 1283, 2138<br />

Li, Chen 1609<br />

Li, Dongxiao 328<br />

Li, Fan 1536, 1537, 2008<br />

Li, Fang 2066<br />

Li, Feng 874<br />

Li, FengDI 1611<br />

Li, Fengyuan 1305<br />

Li, Guangyu 984<br />

Li, Haiyang 1315<br />

Li, Hu 1044<br />

Li, Hui 2048<br />

Li, Jia 496<br />

Li, Jiayi 207, 368, 1561<br />

Li, Jie 2018<br />

Li, Jin 1595<br />

Li, Jun 332<br />

Li, Kelvin 1425<br />

Li, Li 35, 1359<br />

Li, Liying 1396<br />

Li, Man 829<br />

Li, Mengsen 1977<br />

Li, Minran 2066<br />

Li, Qin 815<br />

Li, Qingmei 2058<br />

Li, Shuzhao 19, 2232<br />

Li, Tiangang 140<br />

Li, Tony 1952<br />

Li, Wanyu 1657<br />

Li, Wei 1977<br />

Li, Wenwen 505<br />

Li, Xia 87<br />

Li, Xiaodong 1595, 1675<br />

Li, Xiaolei 1948<br />

Li, Xiaoxia 675<br />

Li, Xin 1557<br />

Li, Xu 2048<br />

Li, Yingxia 942<br />

Li, Yinping 1869<br />

Li, Yong 1684<br />

Li, Yue 322<br />

Li, Yunzhou 22<br />

Li, Zhiping 1214, 1215<br />

Li, Zhongbin 1537, 2008<br />

Li, Zhuan 352, 1311<br />

LI, Jie 193, 197<br />

Lian, Jianqi 984<br />

Liang, Chen 2002<br />

Liang, Cheng-Chao 1996<br />

Liang, Guang 22<br />

Liang, Lijian 454<br />

Liang, Nathan L. 216<br />

Liang, T. Jake 218, 779<br />

Liang, Yan 1004<br />

Liangpunsakul, Suthat 267, 1316,<br />

1332, 1903<br />

Liao, Lijun 100<br />

Liao, Wei-Jia 1973<br />

Liaw, Yun -Fan 1552<br />

Liberal, Rodrigo 300, 608<br />

Lichvar, Alicia B. 1741, 1754<br />

Lidofsky, Anna 25, 991, 1015, 1630,<br />

1644<br />

Liedtke, Christian 828, 831, 1924<br />

Liedtke, Wolfgang B. 962, 2137<br />

Liefferinckx, Claire 178<br />

Liew, Zhong Hong 451<br />

Liffmann, Danielle 156, 1792, 1839<br />

Lifton, Richard P. 1719<br />

Lightbourne, Teisha G. 1451<br />

Lillegard, Joseph 2<br />

Lilley, Kirthi 420<br />

Lilly, Les 11, 204<br />

Lim, Arlene 1351<br />

Lim, Joseph K. 94, 540, 1057, 1488<br />

Lim, Kieron B. 1928<br />

Lim, Mary Ann 871<br />

Lim, Rebecca 945<br />

Lim, Sarah P. 47<br />

Lim, Seng Gee 241, 1686, 1928,<br />

2014, 2025<br />

Lim, Won 749, 876, 2028<br />

Lim, Yoo Li 2080<br />

Lim, Young-Suk 367, 446, 448, 459,<br />

1080, 1105<br />

Lima, Erika C. 2200<br />

Limani, Perparim 665<br />

Limbourg, Florian 2128<br />

Limothai, Umaporn 2036<br />

Lin, Andy 1850, 1877<br />

Lin, Chia-Hsin 2040<br />

Lin, Chih-Lin 1996<br />

Lin, Chih-Wei 41, 248, 250, 719<br />

Lin, Chih-Wen 398<br />

Lin, Chun-Che 1996<br />

Lin, Clement Y. 451<br />

Lin, Derek 207, 368, 386, 458, 1561<br />

Lin, Grace 1470<br />

Lin, Han-Chieh 365, 409, 439, 738,<br />

751, 757, 759, 965, 2050<br />

Lin, Henry C. 1693<br />

Lin, Jaw-Town 345, 1547, 1589<br />

Lin, Jianguo 1361<br />

Lin, Kai 2257<br />

Lin, Lanjia 243, 2025<br />

Lin, Lanyi 1611<br />

Lin, Linda I. 1264<br />

Lin, Lung-Huang 1701<br />

Lin, Ming 219, 1130<br />

Lin, Ming-Tsung 1873<br />

Lin, Ming-Tzung 1942<br />

Lin, Pey-Ru 1942<br />

Lin, Selena 1554<br />

Lin, Steven C. 2201<br />

Lin, Su 1765<br />

Lin, Wen-Terng 1701<br />

Lin, Wenyu 991, 1015, 1630, 1644<br />

Lin, Yong 1562<br />

Lin, Ziyu 1772<br />

Linaberry, Misti 702<br />

Linas, Benjamin P. 152, 1811<br />

Linda, Sher 200<br />

Lindell, Per 2258<br />

Lindhout, Darrin 1986<br />

Lindor, Keith D. 73, 75, 76, 546, 562,<br />

605, 630, 634<br />

Line, Pål-Dag 819<br />

Ling, John 1133, 1707<br />

Ling, Lei 106, 198, 627, 1986<br />

Ling, Victor 824<br />

Lingala, Shilpa 1717, 1722<br />

Lingiah, Vivek A. 295, 1181<br />

Linthicum, Mark T. 1042<br />

Liorre, Giulia 431<br />

Liou, Iris W. 1211<br />

Lippai, Dora 512<br />

Lippincott-Schwartz, Jennifer 1909<br />

Lisboa, Quelson C. 2200<br />

Lisman, Ton 740, 2249<br />

Lison, Hortensia 264, 554, 1072, 2082<br />

Lissock, Frédéric S. 1833<br />

Littera, Roberto 309<br />

Littlejohn, Margaret 1566<br />

Litwin, Alain H. 40<br />

Liu, Andy 199, 201, 1745, 1750<br />

Liu, Benny 7, 237, 1252, 2159<br />

Liu, Changlu 2147, 2208<br />

Liu, Chao 498, 1951<br />

Liu, Chen 138, 989, 1967<br />

Liu, Chen-Hua 1828<br />

Liu, Chun-Jen 2027<br />

Liu, Chunxiao 1537<br />

Liu, Fang 730, 731<br />

Liu, Fanwei 994, 1000<br />

Liu, Feng 782, 1436<br />

Liu, Fu-Tong 1285<br />

Liu, Hannah 1455<br />

Liu, HongLing 1609<br />

Liu, Hongqun 1529<br />

Liu, Hui-Xin 63, 508, 649, 674<br />

Liu, James 506<br />

Liu, Jessica 1593, 1604<br />

Liu, Jialiang 1536, 1537, 2008<br />

Liu, Jie 1374<br />

Liu, Jinfeng 1541, 1579, 1617, 1773<br />

Liu, Jingmei 328<br />

Liu, Ju 36<br />

Liu, KeHui 1611<br />

Liu, Lawrence 1095, 1185<br />

Liu, Lei 468<br />

Liu, Li 722<br />

Liu, Liming 1305, 1306<br />

Liu, Lin 247, 824, 1128<br />

Liu, Mengyuan 1266<br />

Liu, Miao 1562<br />

Liu, Po-Hong 365, 409<br />

Liu, Qian 1647<br />

Liu, Ran 41, 248, 1106<br />

Liu, Runping 22, 1521<br />

Liu, Sichen 807<br />

Liu, Susan B. 826, 1367, 1368, 1379<br />

Liu, Sze Hang Kevin 1550<br />

Liu, Tianhui 1404<br />

Liu, Ting 1952<br />

Liu, Wanqing 56


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1339A<br />

Liu, Wei 710, 719, 723<br />

Liu, Xiao 890, 991, 1015, 1630, 1644<br />

Liu, Xiaoying 658<br />

Liu, Xiu-Ping 366<br />

Liu, Xuan 1198<br />

Liu, Xue-Jing 935<br />

Liu, Yali 2061<br />

Liu, Yan 1198, 1595, 1675<br />

Liu, Yang 1678, 2021<br />

Liu, Yingxia 2048<br />

Liu, Yuhan 1611<br />

Liu, Yun 1383<br />

Liu, Yunlong 1332<br />

Liu, Zhang-Xu 2135<br />

Liu, Zhaohui 728<br />

Liu, Ziyi 1374<br />

Liu, Li 1939<br />

Livingston, Stephen 1798, 1807<br />

Livnat, Yuval 1032<br />

Lizarzábal, Maribel 595<br />

Llach, Josep 2084<br />

Lledo, Jose Luis 378<br />

Lleo, Ana 73, 602, 634<br />

Llop, Elba 1511<br />

Llorente-Izquierdo, Cristina 59<br />

Llovet, Josep M. 253, 254, 860<br />

Llovet, Patricia 1889<br />

Lloyd, Carla 1234<br />

Lo, Chung-Mau 70, 1746, 2000<br />

Lo, Gin-Ho 745<br />

Lo, Pei-Shuan 2040<br />

Lo, Sum Team 2225<br />

Lo, Tracie 1286<br />

Lo, Victor 1998<br />

Lo, Ying-Ru 577<br />

Lo Re, Vincent 1488<br />

Lo Turco, Edson G. 240<br />

Loaeza del Castillo, Aurora 595<br />

Lobach, Iryna 841<br />

Lobritto, Steven J. 1731<br />

Locarnini, Stephen 32, 1551, 1557,<br />

1563, 1566, 2022<br />

Locatelli, Luigi 1408<br />

Locati, Massimo 602<br />

Locklear, Cameron T. 575<br />

Loehr, Hanns F. 1170, 1200<br />

Loggi, Elisabetta 1888<br />

Lohse, Ansgar W. 76, 523, 1803<br />

Lok, Anna S. 67, 94, 576, 1555, 1850,<br />

1877, 2015, 2052<br />

Lombardero, Manuel 121<br />

Lomberk, Gwen 692<br />

Londjon-Domanec, Isabelle 1163, 1179<br />

Londoño, María-Carlota 982<br />

Long, Gang 992<br />

Long, Michelle T. 2191<br />

Long, Zhengyu 2035<br />

Long, Zhifeng 166<br />

Longato, Lisa 337, 1373, 1388, 1407,<br />

1445<br />

Longerich, Thomas 2128<br />

Longhi, Maria Serena 300, 608<br />

Longo, Diane M. 224, 228<br />

Loo, Nicole 743<br />

Loomba, Rohit 45, 236, 239, 913,<br />

1425, 1428, 1431, 1435, 1720,<br />

2149, 2150, 2153, 2154, 2201,<br />

2220, 2239<br />

Loomes, Kathleen M. 133, 1693, 1694<br />

Loomis, Timothy P. 93, 1786<br />

Lopategi, Aritz 2087<br />

Lopes, Marta 1033<br />

Lopez, Rocio 855, 856, 2162, 2182,<br />

2186<br />

Lopez Gomez, Marta 1511<br />

Lopez-Andujar, Rafae 1251<br />

López-Nevot, MAngel 583<br />

Lopez-Talavera, Juan-Carlos 1106<br />

Lopez-Vicario, Cristina 2087<br />

Lorber, Margalit 1032<br />

Lorent, Kristin 814<br />

Loss, George E. 1242<br />

Lotersztajn, Sophie 1288, 1391<br />

Lotteau, Vincent 1652<br />

Lotto, Martin 1752<br />

Loughran, Patricia 184<br />

Louie, Stan G. 1936<br />

Louie, Vincent 1135<br />

Loustaud-Ratti, Veronique 777, 1162,<br />

1187<br />

Louvet, Alexandre 145, 264, 1269,<br />

2082, 2222<br />

Louwers, Lisa 850, 1271<br />

Lovell, Sandra S. 250<br />

Loveridge, Robert 1775<br />

Low, Sarah E. 1928<br />

Lowe, Andrew 1100<br />

Lowe, Patrick 496, 1318, 1330<br />

Lowy, Elliott 540<br />

Loy, Veronica 1172<br />

Loyer, Xavier 887<br />

Lozada, Maria E. 127<br />

Lozano-Sepulveda, Sonia A. 1022<br />

Lu, Fengmin 34, 1634<br />

Lu, James T. 161<br />

Lu, Jie 692<br />

Lu, Junfeng 2061<br />

Lu, Mei 15, 525, 1578, 1789, 1800<br />

Lu, Minqiang 353, 375<br />

Lu, Quiajin 613<br />

Lu, Shelly C. 953, 1952, 2244<br />

Lu, Sheng-Nan 1593, 1604, 1828,<br />

1873, 2010<br />

Lu, Sizhao 60<br />

Lu, Wei 2023<br />

Lu, Wei-Yu 678<br />

Lu, Yuanfang 1910<br />

Lua, Ingrid A. 1354<br />

Luangmonkong, Theerut 1434<br />

Lubel, John 442, 483<br />

Lucena, M. I. 225, 583, 595, 596,<br />

1933<br />

Lucey, Michael R. 1231<br />

Lucidarme, Damien 1072<br />

Lucidi, Cristina 2067<br />

Ludwig, Andreas 100<br />

Luedde, Tom 828, 1924<br />

Lueth, Stefan 1629<br />

Luetjohann, Dieter 922, 961<br />

Luetkemeyer, Anne 40, 726<br />

Luft, Olga 1297<br />

Lugea, Aurelia 904<br />

Lui, Ka Yin 353, 375<br />

Luk, Kevin 1597<br />

Lukacs-Kornek, Veronika 43, 1299,<br />

1385<br />

Luke, Amy 532, 559<br />

Luketic, Velimir A. 52, 607, 609, 628,<br />

715, 1309, 1472, 2235<br />

Lukose, Thresiamma 272, 1454<br />

Luli, Saimir 44<br />

Lumeng, Lawrence 948<br />

Luminos, Luminita 2043<br />

Lunel-Fabiani, Françoise 1162<br />

Lunghi, Giovanna 1743, 1747<br />

Luntz, Steffen P. 1227<br />

Luo, Donghan 39<br />

Luo, Jian 106, 627, 1986<br />

Luo, Jianhua 670<br />

Luo, Mei 904<br />

Luo, Qingyang 1242<br />

Luo, Xiaoping 909, 1684<br />

Luo, Xun 1727, 1729<br />

Luo, Yan 1161, 1198<br />

Luo, Yuhuan 139, 142, 907<br />

Luo, Yunjun 159<br />

Luong, Tu Vinh 337, 413, 1373, 1388<br />

Lupinacci, Lisa Chiacchierini 707<br />

Lustig, Robert H. 1730<br />

Lutchman, Glen A. 1182, 1189<br />

Lutz, Karen 635, 645<br />

Lutz, Philipp 1008, 1287, 2077<br />

Lutz, Thomas 1060, 1081, 1156<br />

Luyendyk, James P. 81<br />

Luz, Rodrigo 273, 289<br />

Lv, Juan 1149<br />

Lv, Jun 1616<br />

Lv, Yi 698<br />

Ly, Mytop 1549<br />

Lynn, Amanda M. 643<br />

Lytvyak, Ellina 641<br />

M. Yoshida, Eric 308<br />

Ma, Ariel 549<br />

Ma, Hong 1660<br />

Ma, Hsiao-Yen 186<br />

Ma, Jimmy 2091<br />

Ma, Li 500<br />

Ma, Lina 2061<br />

Ma, Linyuan 2107<br />

Ma, Mang M. 204<br />

Ma, Mang-Ming 303<br />

Ma, Xiaoli 2025<br />

Ma, Xiong 308<br />

Ma, Yun 300, 305, 608<br />

Maan, Raoel 445, 1099, 1166<br />

Maasoumy, Benjamin 1131, 1572,<br />

1597, 1797<br />

Mabileau, Guillaume 1845<br />

Macaigne, Gilles 1072, 2082<br />

MacArthur, Meaghan 11<br />

MacConell, Leigh 239, 635, 644, 645<br />

Macdonald, Douglas 1904<br />

Macdonald, Graeme A. 1264<br />

Macdonald, Stewart 113<br />

MacDonald - Ottevanger, Sigrid 1846<br />

Macé, Vincent 264<br />

Macedo, Guilherme 1107, 1540<br />

Macera, Margherita 1837<br />

MacGregor, Louis W. 1830<br />

Machado, Joao 1256<br />

Machado, Mariana V. 897, 936, 960,<br />

1386, 2188<br />

Machicao, Victor I. 1513<br />

Machida, Keigo 682, 1918, 1936,<br />

1945, 2118<br />

Machineni, Surendra 2261<br />

Machnes, Ziv 730<br />

Macías-Rodríguez, Ricardo 828, 1471,<br />

1508


1340A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Mack, Cara L. 1692, 1699, 1706,<br />

1709<br />

Mackey, Aaron J. 175, 915, 1916<br />

Mackey, Rachel H. 2228<br />

Macklin, Jared 1350<br />

Mackman, Nigel 81<br />

MacLachlan, Ian 2007, 2062<br />

MacLachlan, Jennifer H. 1566, 2060<br />

MacNaughtan, Jane 908, 1300<br />

MacQuillan, Gerry C. 848, 2099<br />

Maddox, Avery 823<br />

Madejon, Antonio 1028<br />

Madireddi, Malavi 141<br />

Madrigal, Pedro 258<br />

Madsen, Bjørn S. 773, 780, 789, 792,<br />

1440<br />

Madsen, Emilie A. 1437<br />

Maeda, Haruka 1642<br />

Maeda, Kenji 1655<br />

Maekawa, Shinya 406, 1027, 1565<br />

Maes, Michaël 586<br />

Magaldi, Lora 1864, 1880<br />

Magdaleno, Fernando 1382, 2124<br />

Magee, John C. 133, 1694<br />

Maggart, Michael 1223<br />

Maggi, Mario 882, 902<br />

Magistroni, Paola 1<br />

Magni, Carlo F. 222<br />

Magni, Paolo 2187<br />

Magnusson, Baldur 2257<br />

Mahachai, Varocha 416<br />

Mahajan, Indra 1945<br />

Mahale, Parag 1030, 1160<br />

Mahalingam, Devalingam 1900<br />

Maharshi, Sudhir 1473<br />

Maher, Jacquelyn J. 306<br />

Mahgoub, Amar 1184, 2006<br />

Mahmoud, Huda 1480<br />

Mahmoud, Sahah A. 1011<br />

Mahoney, Douglas W. 169<br />

Mahtab, Mamun A. 2157<br />

Maier, Martin A. 36<br />

Maieron, Andreas 1092, 1885<br />

Maillard, Aline 132<br />

Maillette de Buy Wenniger, Lucas 637<br />

Maimone, Sergio 288, 1476<br />

Main, Janice 1826<br />

Maini, Mala K. 1626<br />

Maione, Sabatino 1837<br />

Maire, Laurence 1083<br />

Mairiang, Pisaln 416<br />

Maitland, Hillary 761<br />

Maitland-van der Zee, Anke H. 225<br />

Maiwall, Rakhi 83, 103, 146, 262,<br />

694, 1288, 2070, 2092<br />

Major, Marian E. 1546, 1854<br />

Makar, George A. 536<br />

Makino, Yuki 231, 453, 669, 1082,<br />

1357, 1661, 1985, 2130, 2136<br />

Makoto, Sanomura 966<br />

Makris, Alexia 1784<br />

Malago, Massimo 337, 1388, 1517<br />

Malato, Yann 98<br />

Malaya, Irina 721<br />

Maldonado, Eduardo N. 319<br />

Malek, Nisar P. 1183<br />

Malhi, Harmeet 58, 1315<br />

Malhotra, Pawan 1659<br />

Maliakkal, Benedict 82, 248, 250, 570,<br />

1478<br />

Malik, Mohammad U. 1255<br />

Malinverni, Raffaele 221, 1884<br />

Malinverno, Federica 1743, 1747,<br />

1751<br />

Malipatel, Renuka 284<br />

Mallano, Alessandra 1802, 1844<br />

Mallat, Ariane 421, 449, 1123, 1232<br />

Mallet, Maxime 744, 1510<br />

Mallet, Vincent 1194, 1803<br />

Mallolas, Josep 210<br />

Malta, Fernanda 48<br />

Maltez, Fernando 1859<br />

Man, Tak Yung 678<br />

Manch, Richard A. 1111, 1155, 1164<br />

Mancinelli, Romina 817<br />

Mandorfer, Mattias 149, 270, 475,<br />

561, 732, 755, 771, 1204, 1466<br />

Mandrekar, Pranoti 1351<br />

Maneschi, Elena 882, 902<br />

Manesis, Emanuel K. 2019<br />

Mangia, Alessandra 91, 249<br />

Mangini, Chiara 629<br />

Manini, Matteo A. 1747<br />

Manka, David 175, 664, 915, 1916<br />

Manka, Paul P. 1779, 1978, 2143<br />

Manley, Jr., Michael W. 650, 1767<br />

Mann, Derek 44, 652, 1381, 1408<br />

Mann, Jelena 1381<br />

Mannalithara, Ajitha 88, 1208, 2155<br />

Mannem, Arun 1753<br />

Mannino, Federica 1127<br />

Manns, Michael P. 76, 243, 444, 445,<br />

616, 712, 1002, 1045, 1049, 1084,<br />

1130, 1131, 1166, 1572, 1861,<br />

2026, 2128<br />

Mano, Yohei 476, 1643<br />

Manoharan, Muthiah 36<br />

Manolakopoulos, Spilios 2012, 2019<br />

Manon-Jensen, Tina 1437<br />

Manousou, Pinelopi 413, 801, 1445<br />

Manser, Christine N. 76<br />

Mantovani, Lorenzo G. 571, 1453<br />

Mantry, Parvez S. 704, 722, 1039,<br />

1084, 1146, 1169<br />

Manyakin, Nikolai 599<br />

Manzano Núñez, Fátima 691<br />

Mao, Qianguo 2048<br />

Mao, Shennen A. 2<br />

Mao, Tin 2258<br />

Marabita, Francesco 602<br />

Maragkos, Konstantinos 1126<br />

Maras, Jaswinder S. 581, 1005, 1344<br />

Marasco, Giovanni 741<br />

Maraver, Marta 1863<br />

Maraver Zamora, Marta 2171<br />

Marc, Abrams 510<br />

Marcellin, Patrick 241, 243, 887, 998,<br />

1557, 1601, 1847, 1860, 2014,<br />

2025, 2037, 2044, 2045, 2059<br />

Marchesini, Giulio 2164<br />

Marconi, Vincent 1488<br />

Marcus, James H. 1745<br />

Marcus, Sonja 380, 457<br />

Mare, Ruxandra G. 795<br />

Marek, George W. 138<br />

Marenco, Ted 730<br />

Marengo, Andrea 2209<br />

Margolis, Mary Kay 1467, 1485<br />

Margotti, Marzia 1888<br />

Marhenke, Silke 2128<br />

Marianelli, Leonardo 1752<br />

Maricic, Igor 1345<br />

Marier, J.F. 1915<br />

Marin, Jose J. 506<br />

Marin, Veronica 1339, 1346, 2098<br />

Marinho, Rui T. 150, 1859<br />

Mariño, Zoe 982<br />

Marins, Ed G. 1597<br />

Mariotti, Valeria 811<br />

Marisi, Giorgia 435<br />

Mariz, Daniela M. 2072, 2076<br />

Marker, Amanda E. 1694<br />

Markmann, James F. 837, 1732<br />

Marks, Kristen M. 1094<br />

Marks, Philippa 1083<br />

Marmon, Tonya 75, 155, 317, 628,<br />

635, 644, 645, 1458<br />

Maroni, Luca 21<br />

Marotta, Paul 1246<br />

Marquardt, Jens U. 677<br />

Marques, Carlos Eduardo C. 762<br />

Marques, Dalila M. 804<br />

Marquet, Pierre 1162<br />

Marquez, Max 11<br />

Marra, Fabio 973, 1410<br />

Marra, Fiona 1052, 1131, 1140<br />

Marrero, Idania 1345<br />

Marrero, Jorge A. 371<br />

Marrero Colon, Wesley J. 67<br />

Marri, Smitha 436<br />

Marroncini, Giada 949, 1963<br />

Marrone, Aldo 1837<br />

Marrone, Giusi 337<br />

Mars, Wendy M. 670, 1979<br />

Marschall, Hanns-Ulrich 76, 308, 662,<br />

810<br />

Marsh, Christopher L. 861, 872<br />

Marsh, Philip 47<br />

Marshall, Aileen 413, 489, 490, 1904<br />

Marshall, Vanessa 1154<br />

Marshall, Vincent D. 395<br />

Marshall, William L. 725, 730<br />

Martay, Kenneth 878<br />

Marti, Francesc 673<br />

Martí-Rodrigo, Alberto 1913<br />

Martin, Aaron 1906<br />

Martin, Daniel 774<br />

Martin, Jennifer H. 225<br />

Martin, Natasha K. 1794, 1830<br />

Martin, Paul 1054, 1128, 1148, 1206,<br />

2256<br />

Martin, Ross 91<br />

Martin, Steven R 1576<br />

Martin-llahi, Marta 378<br />

Martinello, Marianne 1083<br />

Martinet, Wim 894<br />

Martinez, Allyson 830<br />

Martinez, Anthony D. 1823<br />

Martinez, Elisa 717, 729<br />

Martinez, Jose Luis 1511<br />

Martinez, Mercedes 1731, 2129<br />

Martinez, Miguel 734<br />

Martinez, Olivia M. 341, 2123<br />

Martínez, Jorge A. 1217<br />

Martinez Rodriguez, Leonardo A. 2223<br />

Martinez Wassaf, Maribel 1752<br />

Martinez-Arranz, Ibon 953, 2163, 2244<br />

Martinez-Chantar, Maria L. 953, 2244<br />

Martinez-Llordella, Marc 1117<br />

Martinez-Lopez, Erika 1855


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1341A<br />

Martini, Silvia 1<br />

Martinot-Peignoux, Michelle 998, 1601,<br />

2044, 2045<br />

Martins, Americo 1256<br />

Martins, Eduardo B. 241, 2014, 2025<br />

Martins, Flaviano S. 2200<br />

Martins, Paulo N. 845<br />

Marukian, Svetlana 175, 1916<br />

Marusawa, Hiroyuki 1989<br />

Maruyama, Hitoshi 1490, 1516<br />

Maruyama, Kageshi 1433<br />

Maruyama, Toru 1912<br />

Marx, Steven 1042, 1050, 1068, 1069<br />

Marzioni, Marco 21, 76, 602, 621,<br />

1802, 1844<br />

Masaki, Keiichi 986, 1056, 1627,<br />

1649, 1681, 1895, 1944<br />

Masaki, Naohiko 794, 1138, 1658,<br />

1795, 2158, 2193<br />

Maslov, Alexander Y. 515<br />

Masola, Valentina 299<br />

Mason, Andrew 73, 76, 303, 628,<br />

634, 641<br />

Mason, Susan 1077<br />

Massaro, Joseph M. 2191<br />

Massarwa, Muhammad 1878<br />

Massella, Maurizio 1802, 1844<br />

Massetto, Benedetta 1076, 2004, 2015,<br />

2043, 2052, 2059<br />

Massey, Veronica L. 587, 1327, 2142<br />

Massie, Allan 1727, 1729<br />

Masson, Neil 753<br />

Masson, Steven 1216<br />

Massoud, Omar I. 1174, 1792<br />

Mast, T. Christopher 717, 729<br />

Masumoto, Akihide 1157<br />

Masuoka, Howard C. 320, 884, 948,<br />

1903<br />

Masur, Henry 92, 220<br />

Masur, Jack 578<br />

Mateo, Gloria 930<br />

Matern, Hugo 1986<br />

Mateus, Elia S. 1256<br />

Mathei, Catharina 1842<br />

Matheny, Michael E. 529<br />

Matherly, Scott C. 52, 1309, 1472<br />

Mathew, Arun 1261, 1276<br />

Mathias, Anita 1133, 1707, 2265<br />

Mathias, Wilson 1484, 1494, 1495<br />

Mathur, Karan 420<br />

Mathurin, Philippe 263, 1269, 2082,<br />

2222<br />

Matilla, Ana 378<br />

Mato, Jose M. 953, 1952, 2163, 2244<br />

Matono, Tomomitsu 1499<br />

Matsubara, Tsutomu 517, 592<br />

Matsubara, Yasuo 505<br />

Matsubayashi, Sunao 2097<br />

Matsuda, Hidetaka 944, 1682<br />

Matsuda, Shuya 1565<br />

Matsuda, Takashi 1447<br />

Matsuda, Yasunobu 2165<br />

Matsui, Takeshi 1097, 1655, 1778,<br />

1983<br />

Matsumoto, Akihiro 1673<br />

Matsumoto, Nobuyuki 1446<br />

Matsumoto, Shuichi 2097<br />

Matsumoto, Tomonori 1989<br />

Matsumoto, Toshihiko 329, 1447<br />

Matsumoto, Yoshihiro 2074<br />

Matsumoto, Yoshinari 191<br />

Matsunaga, Kazuhito 760, 1464, 1492<br />

Matsunaga, Kotaro 1446<br />

Matsunami, Kayoko 1006<br />

Matsushita, Hiroshi 1327<br />

Matsuura, Bunzo 1293<br />

Matsuura, Kentaro 1006<br />

Matsuura, Yoshiharu 1643<br />

Matsuzaki, Yasushi 970<br />

Matsuzawa, Naoto 1377<br />

Matta, Bassem 1887<br />

Matta, Benjamin M. 338<br />

Matta, Laura 309, 1153<br />

Matthews, Gail 210, 1083, 1826,<br />

1868<br />

Mattos, Angelo A. 469<br />

Matwiy, Trudy 1576<br />

Mauer, Amy S. 58<br />

Maul, Timothy M. 216<br />

Maurice, James B. 113, 764<br />

Maurice, Michèle 196<br />

Maurice, Sean B. 1413<br />

Mauss, Stefan 1060, 1078, 1081,<br />

1156, 1559, 1861<br />

Maverakis, Emanual 606<br />

Mawatari, Seiichi 799, 937, 1421,<br />

1838<br />

May, Sarah B. 357<br />

Mayatepek, Ertan 1227<br />

Mayayo-Peralta, Isabel 492<br />

Maybüchen, Lara 961<br />

Mayhew, Christopher 676<br />

Maynard-Muet, Marianne 1162<br />

Mayne, Tracy J. 75, 155, 313, 317,<br />

537, 546, 562, 1457, 1458<br />

Mayo, Marlyn J. 73, 106, 627, 634<br />

Mayo, Rebeca 620, 2244<br />

Mayo, Tara 36<br />

Mayumi, Iijima 1642<br />

Mazagova, Magdalena 59, 1319<br />

Mazariegos, George V. 763<br />

Mazer, Laura 1477<br />

Mazo, Daniel F. 1033, 2221<br />

Mazza, Giuseppe 337, 1388<br />

Mazzaferro, Vincenzo 1751<br />

Mazzarelli, Chiara 1751<br />

Mazzella, Giuseppe 609<br />

Mazzotta, Francesco 249<br />

McCaleb, Michael L. 183, 1353<br />

McCallister, Katharine 548<br />

McCann, Adam 565<br />

McCaughan, Geoff 106, 627, 746,<br />

1049, 1077, 1084, 1243, 1264<br />

McCauley, Bryan 607, 610<br />

McClain, Craig J. 114, 926, 1303,<br />

1304, 1305, 1306, 1312, 1331,<br />

1334, 2141<br />

McClintock, Kevin 2007<br />

McCone, Jonathan 715<br />

McConville, John F. 544<br />

McCormick, Aiden 1275, 2069<br />

McCormick, Joseph B. 1424<br />

McCorriston, Debra 871<br />

McCoy, Annette 1709<br />

McCullough, Arthur J. 239, 1265,<br />

2153, 2154, 2162, 2182, 2186,<br />

2220, 2239<br />

McCullough, Rebecca L. 334, 1308,<br />

2132<br />

McCune, Anne 564<br />

McDaniel, Kelly 834, 835, 929, 1317,<br />

1349<br />

McDonald, Stuart A. 101<br />

McGeough, Matthew D. 1371<br />

McGill, Mitchell R. 585, 586, 598,<br />

1767<br />

McGillicuddy, John W. 1272<br />

McGilvray, Ian 11<br />

Mcglone, Cindy 530<br />

McGuire, Brendan M. 1174, 1250,<br />

1768<br />

McHutchison, John G. 38, 91, 96, 205,<br />

219, 249, 616, 624, 632, 737, 739,<br />

746, 777, 1034, 1045, 1049, 1076,<br />

1080, 1094, 1105, 1120, 1128,<br />

1130, 1428, 1435, 1442, 2015,<br />

2037, 2052, 2149, 2239<br />

McIntyre, Chris 1480<br />

McKiernan, Patrick J. 2105<br />

McKiernan, Susan 1103<br />

McKillop, Iain H. 653<br />

McLauchlan, John 1173<br />

Mclaughlin, Leo 534<br />

McLaughlin, Mary 92, 220<br />

McLeman, Lindsay 1104<br />

McLeod, Marie-Ange 1588, 2053<br />

McLin, Valerie A. 1728, 1735<br />

McMahan, Rachel 938, 959, 1708<br />

McMahon, Brian J. 125, 626, 1790,<br />

1798, 1807<br />

McMahon, Peter S. 1052<br />

McMillin, Matthew 1507<br />

McMullen, Megan R. 334, 1308, 1346,<br />

2121, 2132<br />

McNally, John 205, 249, 2265<br />

McNiven, Mark A. 1310<br />

McNulty, Gwen L. 202<br />

McPhee, Fiona 702, 709<br />

McVicker, Benita L. 1969, 2139<br />

Medeiros, Roseane P. 1033<br />

Mederacke, Ingmar 1957<br />

Medici, Valentina 2109<br />

Medina, Vivian G. 1945<br />

Medina-Cáliz, I. 583, 595, 596<br />

Medmoun, Mourad 554<br />

Meehan, Kolin 493<br />

Mega, Ines 1256<br />

Megnien, Sophie 105, 162, 2145<br />

Mehal, Wajahat Z. 805, 979, 1209,<br />

1284, 2179<br />

Mehrab-Mohseni, Marjan 1023<br />

Mehta, Anand 371<br />

Mehta, Gautam 113<br />

Mehta, Neil 4, 50, 53, 414, 471<br />

Mehta, Nicita 496<br />

Mehta, Rajiv 241, 2014, 2025<br />

Mehta, Rajni 16<br />

Mehta, Rohini 891, 956<br />

Mehta, Shivang 1513<br />

Mei, Ruqi 1657<br />

Mei, Xiaonan 854<br />

Meillon, Jean-Christophe 2266<br />

Meissner, Eric G. 1029<br />

Mejia, Alejandro 1169<br />

Mejias, Caroline A. 548<br />

Mekaru, Steven 556<br />

Melancon, Sir Norman T. 1223<br />

Mele, Caterina 432<br />

Melgar, Jennifer 78<br />

Melgar-Lesmes, Pedro 1292


1342A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Melis, Marta 166<br />

Mello, Tommaso 902, 949, 1963<br />

Mellone, Massimiliano 1978<br />

Mells, George F. 155, 601, 1457<br />

Mells, Jamie 19<br />

Melton, Phillip 2185<br />

Meltzer, David O. 152<br />

Melum, Espen 258, 819, 833<br />

Memon, Muhammad S. 1872<br />

Menchise, Alexandra 1695<br />

Mendão, Luís 1859<br />

Mendes-Braz, Mariana 342<br />

Mendez-Bocanegra, Angela 2084<br />

Méndez-Sanchéz, Nahum 595<br />

Meng, Fanyin 190, 454, 514, 621,<br />

817, 834, 835, 929, 1317, 1349<br />

Meng, Li-Na 1465<br />

Meng, Xiaochun 1755<br />

Mengual, Edgardo 595<br />

Mennone, Albert 805, 829<br />

Menon, Krishna 1775<br />

Menon, Rajeev 1066, 1088<br />

Mensa, Federico J. 41, 248, 250, 705,<br />

719<br />

Mensing, Sven 1066<br />

Merchant, Michael L. 587<br />

Mercier-Darty, Melanie 1700<br />

Merion, Robert M. 838<br />

Méritet, Jean-François 30, 997<br />

Merlen, Grégory 179<br />

Merli, Manuela 2067<br />

Merniz, Mohamed 1621<br />

Meroueh, Fadi 1796<br />

Merriman, Raphael 538<br />

Merritt, Elliot 1117<br />

Merrouche, Wassil 450, 2181<br />

Mertens, Joachim C. 1402<br />

Mesarwi, Omar 1422<br />

Meshrekey, Raymond 1076<br />

Metcalfe, Leanne N. 532, 559<br />

Metivier, Sophie 219, 777<br />

Metselaar, Herold J. 68, 1238<br />

Metwally, Sherif N. 1280<br />

Meyer, Bernhard 444<br />

Meyer, Kirstin 617<br />

Meyer, Timothy 344, 1530, 1904<br />

Meyers, Rachel 36<br />

Mezzabotta, Lavinia 2209<br />

Mgaieth, Sara 483<br />

Mi, Yu-Qiang 1465<br />

Mi, Zhiyong 2143<br />

Miailhes, Patrick 1676, 1677<br />

Mian, Ajmal 798<br />

Miao, Chuan L. 552<br />

Miao, Ruoyu 1222<br />

Micco, Lorenzo 1626<br />

Michaels, Anthony 874, 1260, 1280<br />

Michalak, Thomas I. 1641<br />

Michalopoulos, George K. 670, 1770,<br />

1979<br />

Michel, Gerd L. 1812<br />

Michel, Maurice 1519<br />

Michel-Treil, Veronique 1597<br />

Michelotti, Gregory A. 66, 936, 962,<br />

1386, 2137<br />

Michielsen, Peter P. 797, 894, 976<br />

Michl, Patrick 1519<br />

Micolochick Steuer, Holly 2009<br />

Middleton, Michael S. 2150<br />

Midorikawa, Yutaka 407<br />

Mieli-Vergani, Giorgina 300, 608,<br />

1704, 1711<br />

Miethke, Alexander G. 823, 1695<br />

Miguez, Marcio Q. 2189<br />

Mikels-Vigdal, Amanda 1379<br />

Miki, Daiki 663, 986, 1645, 1649,<br />

1681, 1858, 1895, 1944<br />

Miki, Koichiro 2031<br />

Miki, Masahiro 1981<br />

Mikolajczyk, Adam E. 544<br />

Milana, Martina 203<br />

Milani, Stefano 1963<br />

Milián-Medina, Lara 1913<br />

Milkiewicz, Piotr 76<br />

Millán, Raquel 1863<br />

Millen, Rosemary M. 996, 2262<br />

Miller, Anne 529<br />

Miller, Charles M. 840, 1271<br />

Miller, Colton M. 507<br />

Miller, George 1282<br />

Miller, Gregory 784<br />

Miller, Lesley 1835<br />

Miller, Michael D. 91, 700, 703, 713,<br />

718, 1168, 2021<br />

Milligan, Scott 1046, 1108<br />

Millis, J. Michael 1229<br />

Mills, Anthony M. 716<br />

Miloh, Tamir A. 1716<br />

Milstein, Stuart 36<br />

Miltiadous, Oriana 253, 254<br />

Min, Albert 1095, 1165<br />

Min, Hae-Ki 905<br />

Minami, Masahito 886<br />

Minami, Tatsuya 461<br />

Minczeles, Noémie 1237<br />

Mine, Tetsuya 314, 383, 697<br />

Minello, Anne 1158, 1187<br />

Ming-jong, Bair 1996<br />

Mingarelli, Eleonora 21<br />

Minguez, Beatriz 378, 380, 438, 457<br />

Minichini, Carmine 431, 1543<br />

Minnis-Lyons, Sarah E. 678<br />

Minniti, Caterina 1498<br />

Minomi, Kenjiro 1375, 1378<br />

Minsart, Charlotte 178<br />

Minteer, William B. 4, 108<br />

Minter, Freneka F. 529<br />

Minuk, Gerald Y. 361, 1149<br />

Miquel, Juan F. 2210<br />

Mirabella, Stefano 1<br />

Miraglia del Giudice, Emanuele 1837<br />

Miranda, Joana P. 1921<br />

Mirolo, Massimiliano 602<br />

Mirshahi, Faridoddin 807, 905, 1309<br />

Mirza, Darius 1234<br />

Mirza, Roshni 2078<br />

Mirzac, Angela 730<br />

Mirzazadeh, Ali 1793<br />

Misaki, Ryo 1642<br />

Mishra, Alita 575<br />

Mishra, Bibhuti 130, 682<br />

Mishra, Lopa 130, 682, 2118<br />

Misra, Sandeep 170<br />

Missale, Gabriele 355<br />

Mistry, Nipun 682<br />

Mistry, Sameer 163<br />

Misu, Hirofumi 990, 1010<br />

Misurski, Derek A. 1089<br />

Mita, Eiji 1197, 1584, 2213<br />

Mitamura, Kuniko 804<br />

Mitchell, Angela M. 1708<br />

Mitchell, Ben C. 2021<br />

Mitchell, Donald G. 1906<br />

Mitchell, Joanne 1077<br />

Mitchell, Kristen A. 1403<br />

Mitchell, Natalie 1257<br />

Mitchell, Paul D. 136<br />

Mitchell, Tim 2099<br />

Mitov, Mihail I. 673<br />

Mitra, Lalita G. 103, 2070<br />

Mitsuhashi, Shuji 1394<br />

Mitsuya, Hiroaki 1655<br />

Mitsuyoshi, Hironori 886, 888, 901,<br />

1980, 2230<br />

Mittal, Rasham 404, 1171, 1569, 1619<br />

Mittal, Sahil 244, 357<br />

Mitterhofer, Anna Paola 983<br />

Miuma, Satoshi 952, 1671<br />

Miura, Mika 1027, 1565<br />

Miyaaki, Hisamitsu 952, 1671<br />

Miyabe, Katsuyuki 396<br />

Miyadera, Hiroko 1138<br />

Miyagawa, Koichiro 593, 883<br />

Miyagawa, Shinichi 869<br />

Miyagi, Takuya 1636<br />

Miyajima, Ichiro 304, 1819<br />

Miyake, Rei 569, 2140, 2234<br />

Miyake, Teruki 1293, 2146<br />

Miyake, Yasuhiro 339, 2236<br />

Miyaki, Eisuke 986, 1150, 1649, 1670,<br />

1681, 1944, 2120<br />

Miyamoto, Yukiko 1319<br />

Miyanishi, Koji 1984<br />

Miyata, Takashi 820<br />

Miyazaki, Masayuki 1157<br />

Miyazaki, Miyono 1375<br />

Miyazaki, Teruo 970<br />

Miyoshi, Eiji 906, 1642, 1809, 2156<br />

Miyoshi, Kenichi 1499<br />

Miyoshi, Masato 164<br />

Miyoshi, Takahiro 477<br />

Mizerski, Jacek 2043<br />

Mizokami, Masashi 91, 476, 794,<br />

1080, 1105, 1138, 1448, 1577,<br />

1614, 1632, 1643, 1658, 1673,<br />

1778, 1795, 2158, 2193<br />

Mizuguchi, Hiroyuki 1690<br />

Mizukoshi, Eishiro 370, 392, 430,<br />

1348, 1653<br />

Mizuochi, Tatsuki 676, 1724<br />

Mizutani, Kayo 906, 2156<br />

Mkada, Helmi 786<br />

Mo, Frankie 344<br />

Mo, Hongmei 91, 92, 219, 220, 713,<br />

718, 1080, 1105, 1120, 1168<br />

Mo, Lein-Ray 1996<br />

Moaven, Omeed 957, 1938<br />

Mobashery, Niloufar 708, 714<br />

Mochida, Hatsune 947<br />

Mochida, Satoshi 1059, 1064, 1474,<br />

1486, 1577, 1757, 2241<br />

Modaresi Esfeh, Jamak 1894<br />

Modi, Apurva A. 1146, 1199<br />

Modi, Deval 724<br />

Moeini, Agrin 253, 254<br />

Moeller, Frederick G. 1491<br />

Moestrup, Søren K. 2071<br />

Mogalian, Erik 2265<br />

Mogilenko, Denis 928<br />

Mogul, Douglas 1727, 1729


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1343A<br />

Mohamed, Essa A. 1966<br />

Mohamed, Sofiane 1009<br />

Mohammad, Mohammad K. 1303,<br />

1334<br />

Mohankumar, Swathi 358<br />

Mohanty, Sujata 690<br />

Mohr, Ashley M. 60, 1969<br />

Mohr, Rosemary 521<br />

Mokhtarani, Masoud 1467, 1478,<br />

1485<br />

Mole, Larry A. 93, 1786<br />

Molenkamp, Richard 208, 1846, 1995<br />

Moles, Anna 652<br />

Molina, Esther 2032<br />

Molleston, Jean P. 133, 159, 1694<br />

Molokanova, Olena 955, 1353<br />

Molokhia, Mariam 225<br />

Molrine, Deborah C. 704<br />

Momen-Heravi, Fatemah 109, 512<br />

Moncsek, Anja 1402<br />

Mondal, Kalyani 1986<br />

Monegal, Ana 615<br />

Monereo Muñoz, Maria Blanca 1341<br />

Monga, Satdarshan (Paul) S. 24, 255,<br />

821<br />

Monge, Fanny 891, 892, 956, 1428,<br />

2192<br />

Monico, Jesus 1178<br />

Monico, Sara 1743, 1747, 1751<br />

Monneret, Denis 786, 1510<br />

Monraats, Melanie A. 1238<br />

Monsour, Howard P. 1215<br />

Montagnese, Sara 292<br />

Montalbano, Marzia 1751<br />

Montalbano, Mauro 1991<br />

Montalva, Eva 1251<br />

Montane, Eva 596<br />

Montano-Loza, Aldo J. 303, 308, 616,<br />

641, 1481<br />

Monteiro, Eduardo d. 2248<br />

Montemezzo, Mauricio 2248<br />

Montestruc, François 1621<br />

Monti, Monica 1190<br />

Montialoux, Helene 3, 95<br />

Montin, Umberto 1751<br />

Monto, Alexander 1041<br />

Mony, Shruti 1155<br />

Mookerjee, Rajeshwar 908, 1300,<br />

1482, 1517<br />

Moon, Hyuk 504<br />

Moon, Scott 1041<br />

Mooney, Tim 36<br />

Moonka, Dilip 307, 1254<br />

Moore, Ann 1111, 1155, 1164<br />

Moore, Kevin 113, 337, 1388, 1407<br />

Moore, Raymond M. 1965<br />

Moorjaney, Divya 1456, 1461<br />

Moorman, Anne C. 15, 525, 1789,<br />

1799, 1800<br />

Moorman, Jonathan P. 984<br />

Mor, Eytan 875<br />

Moradpour, Darius 221, 1672, 1884<br />

Morales, Amilcar 1866<br />

Morales-Ibanez, Oriol 1327<br />

Morales-Ruiz, Manuel 65, 659<br />

Moran, Lauren 36<br />

Moran-Salvador, Eva 1381<br />

Morando, Filippo 148, 292<br />

Moratalla, Alba 1300<br />

Morbey, Ana 1256<br />

Moreau, Richard 83, 145, 178, 265,<br />

1288, 1391<br />

Moreira, Rebeca 266<br />

Morel, Philippe 2174<br />

Moreland, Anita L. 1154<br />

Morelli, Annamaria 882, 902<br />

Morelli, Giuseppe 1057<br />

Morelli, Maria Cristina 1751<br />

Moreno, Christophe 178, 1051, 1091,<br />

1179, 2079<br />

Moreno, Gigi 1042, 1068<br />

Moreno-Planas, Jose María 583, 2171<br />

Moreo, Kathleen 1141<br />

Morey, Tristan 89, 1090, 1112<br />

Morgan, Jake R. 152<br />

Morgan, Phillip 1098<br />

Morgan, Sarah 1455<br />

Morgan, Timothy R. 724, 1041, 1129<br />

Morgan-Stevenson, Vicki 946<br />

Mori, Amit 17<br />

Mori, Federica 958<br />

Mori, Kohjiroh 2156, 2202<br />

Mori, Trevor A. 2185<br />

Moriguchi, Michihisa 886, 901, 1980,<br />

2230<br />

Morihara, Daisuke 385, 410<br />

Morikawa, Hiroyasu 429, 1125, 1824<br />

Morikawa, Kenichi 287, 1302, 1672,<br />

1946<br />

Morillas, Rosa María 2032<br />

Morillo, Ramón 1863<br />

Morinaga, Maki 917<br />

Morio, Kei 472, 1893<br />

Morio, Reona 1893<br />

Morisco, Filomena 355<br />

Morishita, Naoki 1082, 1195, 1197,<br />

1201, 2213<br />

Morita, Chie 2031<br />

Moriuchi, Akihiro 799, 937, 1421,<br />

1838<br />

Moriwaki, Hisataka 657<br />

Moriya, Kyoji 461, 912<br />

Moriyasu, Fuminori 1047<br />

Morizono, Syusuke 1159<br />

Morley, Roland 1920<br />

Morling, Joanne R. 2168<br />

Moro, Tadashi 1369<br />

Moroianu, Serban A. 295<br />

Moroney, Justin B. 493<br />

Morosawa, Tatsuki 1026, 1648, 1943,<br />

2167<br />

Morris, Adam R. 1346<br />

Morris, Jason 522<br />

Morris, Jeffrey S. 130<br />

Morris, Meghan D. 1793<br />

Morrison, Dee 530<br />

Morrow, Bernice 943<br />

Morse, Gene D. 332<br />

Morton, Ryan 307, 420<br />

Mosaad, Youssef 1236<br />

Mosadeghi, Sasan 7<br />

Mosca, Nicola 431<br />

Moscat, Jorge 232<br />

Moscatelli, Alessandro 355<br />

Moser, Catherine D. 169, 396, 1965,<br />

1966<br />

Moser, Stephan 1885, 1886<br />

Moshage, Han 591, 612, 969, 1398,<br />

1416<br />

Mosoian, Arevik 1397<br />

Moss, Stephen E. 64<br />

Mossanen, Jana C. 1756<br />

Mössner, Belinda 792<br />

Motamed, David B. 1095, 1185<br />

Motola, Daniel L. 778<br />

Motomura, Kenta 1157, 1159<br />

Motoyama, Hiroaki 869<br />

Motoyama, Hiroyuki 1125<br />

Mott, Justin L. 60<br />

Mottram, Carl 1470<br />

Moucari, Rami 2044, 2045<br />

Mouillot, Thomas 264<br />

Mounzer, Karam 1136<br />

Mousa, Omar 1720<br />

Moussa, Adel 2266<br />

Moustafa, Hamdy M. 858<br />

Mouzaki, Athanasia 2094, 2095<br />

Mouzaki, Marialena 2184, 2238<br />

Moy, Carolyn 1864, 1880<br />

Moya, Angel 1251<br />

Moylan, Cynthia A. 161, 540, 889,<br />

1142<br />

Møller, Holger J. 1300, 2071<br />

Møller, Linda S. 780, 789, 1440<br />

Mridha, Auvro R. 55<br />

Mrzljak, Anna 1113, 1744<br />

Mtegha, Marumbo P. 1712<br />

Mu, Fan 711<br />

Mu, Xianmin 1910<br />

Muczynski, Kimberly 1221<br />

Mudan, Satvinder 599<br />

Mueller, Olaf 2188<br />

Mueller, Tobias 76<br />

Mugavero, Michael 1804<br />

Muir, Andrew J. 96, 106, 616, 624,<br />

627, 632, 1435, 2239<br />

Mukai, Kaori 1636<br />

Mukaida, Naofumi 2119<br />

Mukaide, Motokazu 1614<br />

Mukamal, Kenneth 2227, 2228<br />

Mukherjee, Amitava 197<br />

Mukherjee, Sunil K. 1659<br />

Mukhopadhya, Ashis 1104<br />

Mukhopadhyay, Chinmay 1344<br />

Mulazzani, Lorenzo 905<br />

Mulkay, Jean-Pierre 1842<br />

Mullen, Alan 778<br />

Mullen, Michael P. 1112<br />

Muller, Michael 1924<br />

Müller, Clara 43<br />

Müller, Niklas F. 308<br />

Müllhaupt, Beat 46, 221, 1402, 1818,<br />

1884<br />

Mulligan, David C. 9<br />

Mumtaz, Khalid 350, 874, 1260, 1280<br />

Mumtaz, Shaham 532, 559<br />

Mund, Ross 1899<br />

Mungiole, Nicole S. 521<br />

Munir, Samina 2166<br />

Munoz, Santiago J. 202<br />

Muñoz, Carolina 378<br />

Muñoz, Linda E. 1022, 1281<br />

Muñoz, Nina M. 2118<br />

Muñoz-Durango, Natalia 971<br />

Muñoz-Yagüe, Teresa 896, 939<br />

Munteanu, Mona 786<br />

Munzy, Donna M. 129<br />

Murad, Mohammad H. 282<br />

Muragundla, Anjaneyulu 118<br />

Murai, Kazuhiro 1082, 1661


1344A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Murai, Kazuhisa 990, 994, 1000, 1010<br />

Murakami, Eisuke 1858<br />

Murakami, Kazunari 2267<br />

Murakami, Masashi 970<br />

Murakami, Seishi 168, 990, 1010<br />

Murakami, Shuko 1599, 1655<br />

Murakami, Yoshiki 429, 1125, 1824<br />

Murakami, Yuya 1375<br />

Murakawa, Miyako 164, 406, 1018,<br />

1027, 1959<br />

Murali, Ramachandran 1936<br />

Muraoka, Masaru 1565<br />

Murata, Kazumoto 794, 1138, 1614,<br />

1658, 1673, 1778, 1795, 2158, 2193<br />

Murata, Yasuhiro 1640<br />

Muratori, Luigi 308<br />

Muratori, Paolo 308<br />

Murayama, Asako 1003<br />

Muromachi, Kaori 799, 937, 1838<br />

Muroyama, Ryosuke 505<br />

Murphy, Brian J. 141<br />

Murphy, Rinki 2176<br />

Murphy, Susan K. 889<br />

Murphy, Trudy 1546<br />

Murray, John W. 321, 333<br />

Murray, Karen F. 133, 1694, 2043<br />

Murthy, Karnam S. 807<br />

Musa, Hawah 1122<br />

Musolino, Cristina 1654<br />

Musto, Kaitlyn R. 1245<br />

Musukuma, Kalo 1575<br />

Mut, Deniz 1739<br />

Mutchnick, Milton G. 420<br />

Mutimer, David J. 1045, 1049, 1130,<br />

2003<br />

Mutsaers, Henricus A. 1427, 1434<br />

Mutsuki, Taiji 1157<br />

Mwinyi, Jessica 519, 520<br />

Myers, Robert P. 616, 624, 632, 737,<br />

739, 746, 777, 1428, 1435, 1442,<br />

2149, 2239<br />

Myronovych, Andriy 916<br />

Nabinger, Sarah C. 436<br />

Nadal, Elena 1270<br />

Nader, Fatema 18, 315, 316, 1460<br />

Nadig, Satish 1272<br />

Nadji, Mehrdad 914<br />

Nadler, Jerry L. 807<br />

Nafady, Mohammed 858<br />

Nafisi, Shirin 1111, 1155, 1164<br />

Naga Prasad, Sathyamangla V. 651,<br />

667, 668, 671, 1308<br />

Nagahara, Akihito 1851<br />

Nagai, Masato 551<br />

Nagamatsu, Hiroaki 399<br />

Naganuma, Atsushi 477<br />

Nagaoka, Katsuya 513, 1987<br />

Nagaoka, Shinya 408, 783, 1600<br />

Nagaoki, Yuko 1056<br />

Nagaoki, Yuuko 472, 1893<br />

Nagarkatti, Mitzi 962, 2133, 2137<br />

Nagarkatti, Prakash 962, 2133, 2137<br />

Nagasawa, Hideko 995<br />

Nagase, Syoji 1159<br />

Nagase, Toshihiko 1197, 2213<br />

Nagata, Hiroko 164, 406, 1018, 1027,<br />

1959<br />

Nagata, Naoto 144, 1348, 1968<br />

Naggie, Susanna 1035, 1094, 1142<br />

Nagy, Laura E. 675, 1308, 1339,<br />

1346, 2121, 2132<br />

Nahass, Ronald 40, 219<br />

Nahmias, Yaakov 331, 672, 679<br />

Nahon, Pierre 145, 362, 1808, 1847<br />

Naik, Hetanshi 2111, 2112<br />

Naik, Jahnavi 1095<br />

Naik, Sandhia 1571<br />

Naiki, Kayoko 1064<br />

Naing, Aung 1160<br />

Nair, Devaki 113<br />

Nair, Rashmi T. 1267, 2081<br />

Nair, Satheesh 1212, 1233<br />

Nairat, Samir S. 1843<br />

Naito, Masafumi 796<br />

Naito, Tatsushi 944, 1682, 2119<br />

Naito, Yoshiki 465<br />

Nakabori, Tasuku 97, 231, 669, 1082,<br />

1357, 1661, 1985, 2130, 2136<br />

Nakagawa, Ai 291<br />

Nakagawa, Hayato 461<br />

Nakagawa, Hidetoshi 370, 392<br />

Nakagawa, Hidewaki 1895<br />

Nakagawa, Kentarou 1584<br />

Nakagawa, Mina 164, 406, 1018,<br />

1027, 1959<br />

Nakagawa, Ryo 505<br />

Nakagawa, Shigeki 364, 2174<br />

Nakagome, Yu 1648, 1943, 2167<br />

Nakagomi, Ryo 461<br />

Nakahara, Takashi 986, 1056, 1649,<br />

1681, 1895, 1944<br />

Nakai, Masato 287, 1672, 1946<br />

Nakajima, Hisakazu 901<br />

Nakajima, Shunsuke 927<br />

Nakajima, Tomoaki 1116, 2065, 2207<br />

Nakakuki, Natsuko 1115<br />

Nakamoto, Nobuhiro 181, 569, 2131,<br />

2140, 2234<br />

Nakamoto, Shingo 1207, 1628<br />

Nakamoto, Yasunari 509, 944, 1682,<br />

2119<br />

Nakamura, Ikuo 1047, 1420<br />

Nakamura, Jun 2030<br />

Nakamura, Masato 1207, 1628<br />

Nakamura, Mikiko 1014, 1395, 1653<br />

Nakamura, Minoru 640<br />

Nakamura, Ryota 1522<br />

Nakamura, Shinichiro 2236<br />

Nakamura, Takuya 1648<br />

Nakamura, Toru 688, 1819<br />

Nakamura, Yoshiko 788, 1293<br />

Nakamura, Yuki 472, 1893<br />

Nakamuta, Makoto 707, 1157, 1159<br />

Nakanishi, Hiroyuki 1115<br />

Nakano, Chikage 466, 770, 781,<br />

1533, 1669, 1975<br />

Nakano, Masahito 455, 1819<br />

Nakano, Norihito 472, 1893<br />

Nakanuma, Yasuni 347, 614, 820,<br />

1901<br />

Nakao, Kazuhiko 952, 1671<br />

Nakao, Masamitsu 1577, 1757<br />

Nakao, Sachie 1369<br />

Nakashima, Hiroyuki 324<br />

Nakashima, Kazuhisa 456<br />

Nakashima, Masahiro 324, 952<br />

Nakashima, Osamu 465, 1990<br />

Nakashita, Shunya 296<br />

Nakatani, Kazuki 592<br />

Nakatsuka, Takuma 461<br />

Nakatsura, Tetsuya 509<br />

Nakayama, Nobuaki 1059, 1064,<br />

1474, 1486, 1577, 1757<br />

Nakayama, Yasuhiro 1565<br />

Nakazawa, Manabu 1474<br />

Nakazawa, Takahide 482<br />

Nakazuru, Shoichi 1584<br />

Naksuk, Niyada 1497<br />

Nalpas, Catherine 1179<br />

Nam, Byung Ho 377, 486<br />

Nam, Joon Yeul 474, 479, 1610, 2016<br />

Nan, Yuemin 638<br />

Nangia, Saachi 280, 560<br />

Nanni, Oriana 435<br />

Naoe, Hideaki 513, 1912, 1987<br />

Napoli, Andrew A. 217<br />

Naqvi, Alissa 568, 1093<br />

Naqvi, Syed J. 98<br />

Narciso-Schiavon, Janaína L. 2075<br />

Narimatsu, Hisashi 1448<br />

Naritaka, Nakayuki 1696<br />

Narkewicz, Michael R. 1706, 1708,<br />

1714<br />

Nart, Deniz 1739<br />

Nascimbeni, Fabio 2164<br />

Nascimento e Silva, Rafaella 518, 651,<br />

667, 668<br />

Nash, Michelle 1359<br />

Nasser, Imad 2160<br />

Nasti, Alessandro 325<br />

Natarajan, Sathish Kumar 60<br />

Nathan, Hari 395<br />

Nathanson, Michael H. 661<br />

Natsuizaka, Mitsuteru 1946<br />

Nattermann, Jacob 1008, 1287, 2077<br />

Naugler, Willscott E. 10, 69, 172, 481<br />

Naumann, Uwe 1170, 1200, 1861<br />

Navaratnam, Annalan M. 764<br />

Navarra, Giuseppe 1654<br />

Navarro, J. M. 1863<br />

Navarro, Victor J. 530, 535, 597,<br />

1922, 1923, 1932<br />

Navarro-Zornoza, Maria 8<br />

Navasa, Nicolás 953<br />

Nawa, Takatoshi 26<br />

Nawaz, Arif A. 1872<br />

Nawaz, Mohammad 1827<br />

Nayak, Suman 262<br />

Nayer, Ali 914<br />

Naylor, Paul H. 420<br />

Nazario, Hector E. 1146, 1169, 1199<br />

Nazef, Naim 2116<br />

Ndugga, Nambi J. 1451<br />

Ndungu, Milka 1146, 1199<br />

Neben, Steven 208, 2263<br />

Nederveen, Aart 2152<br />

Neff, Guy 1106<br />

Neff, Maria 1452<br />

Negrin Dastis, Sergio 1091<br />

Negro, Francesco 221, 1539, 1818,<br />

1884, 2174<br />

Negus, Susan 125, 1798, 1807<br />

Neijenhuis, Myrte 293<br />

Nejak-Bowen, Kari 24, 255, 821<br />

Nekachtali, Ouardia 772<br />

Nelson, David R. 94, 716, 989, 1057,<br />

1826<br />

Nelson, Kirbylee K. 1240<br />

Nelson, Mark 210


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1345A<br />

Nemoto, Tomoyuki 944, 1682<br />

Nephew, Lauren D. 865<br />

Nerenz, David R. 1792<br />

Nesher, Eviatar 875<br />

Neubauer, Stefan 2190<br />

Neuberger, James 631<br />

Neumann, Helmut 1506<br />

Neumann-Haefelin, Christoph 1666<br />

Neurath, Markus F. 1506<br />

Neuschwander-Tetri, Brent A. 157, 239,<br />

1535, 2150, 2153, 2154, 2220<br />

Nevens, Frederik 73, 171, 609, 634,<br />

1107<br />

Neves Souza, Lara 1733<br />

Nevzorova, Yulia A. 1924<br />

Newberry, Carolyn 536<br />

Newberry, Elizabeth P. 911, 1361,<br />

1365<br />

Newell, Evan W. 167<br />

Newell, Karen 1452<br />

Newstrom, David 632, 1359, 2149<br />

Newton, Jennifer M. 579<br />

Newton, Kimberly P. 159<br />

Ng, Anthony 2214<br />

Ng, Michel 1151, 1185<br />

Ng, Teresa 41, 248, 250, 705<br />

Ng, Vicky L. 133<br />

Ng, Vivian 488<br />

Ngo, Yen 786<br />

Nguyen, Anh-Hoa 2015<br />

Nguyen, Bach-Yen T. 40, 210, 251,<br />

700, 703, 715<br />

Nguyen, Huy 1549<br />

Nguyen, Justin H. 1236, 1245<br />

Nguyen, Khanh 1549<br />

Nguyen, Kristy 945<br />

Nguyen, Linda 1848<br />

Nguyen, Long H. 386, 458<br />

Nguyen, Michael D. 386, 458, 1816<br />

Nguyen, Mindie H. 123, 207, 359,<br />

367, 368, 386, 387, 388, 428, 458,<br />

459, 1182, 1189, 1561, 1668, 1805,<br />

1806, 1816, 1848, 2009, 2250<br />

Nguyen, My T. 1182, 1549<br />

Nguyen, Nghia H. 207, 359, 368,<br />

386, 388, 458, 1182, 1189, 1561,<br />

1806, 1816, 1848, 2250<br />

Nguyen, Ngoc Tue X. 685<br />

Nguyen, Pauline 387, 428, 1848<br />

Nguyen, Peter T. 386, 388, 458, 1816,<br />

1848<br />

Nguyen, Phirum 913<br />

Nguyen, Quy P. 514, 813<br />

Nguyen, Tuyen M. 36<br />

Nguyen, Vincent G. 207, 368, 1561<br />

Nguyen-Khac, Eric 206, 264, 1072,<br />

1187<br />

Ni, Hong-Min 140, 598, 1325<br />

Ni, Liyun 1133, 1707<br />

Ni, Yen-Hsuan 1701, 1705<br />

Nicholls, Hayley T. 881<br />

Nick, Elisabeth 2086<br />

Nickenig, Georg 961<br />

Nickkholgh, Arash 1227<br />

Nicolas, Carine 1009<br />

Nicolas, Gregory 394<br />

Nicoletti, Paola 225, 1930, 1933<br />

Nicoll, Amanda J. 311, 442, 483<br />

Nie, Yuanyang 904<br />

Niederau, Claus 1861<br />

Nielsen, Mette J. 119, 780, 1432,<br />

2178<br />

Niemietz, Christoph 2103<br />

Niemietz, Christoph J. 684<br />

Nierhoff, Dirk 687<br />

Nieto, Leonardo 1674<br />

Nieto, Natalia 1382, 1940, 2124<br />

Nietzsche, Sandor 808<br />

Nigar, Sofia 272<br />

Niinomi, Takuro 411<br />

Niitsu, Yoshiro 1375, 1378, 1433<br />

Nikolopoulou, Vasiliki 2094, 2095<br />

Nimanong, Supot 467<br />

Ning, Qin 909, 1656, 1660, 1683,<br />

1684<br />

Ning, Shunbin 984<br />

Ninomiya, Masashi 1026<br />

Nio, Kouki 699, 1953<br />

Nischalke, Hans Dieter 1008, 2077<br />

Nishida, Nao 1138, 1632, 2158<br />

Nishida, Naoshi 1935<br />

Nishiguchi, Shuhei 466, 770, 781,<br />

1533, 1669, 1975<br />

Nishijima, Norihiro 1989<br />

Nishikawa, Hiroki 466, 770, 781,<br />

1533, 1669<br />

Nishikawa, Taichiro 886, 888, 901,<br />

1980<br />

Nishimura, Miyuki 1375, 1378<br />

Nishimura, Norihisa 689<br />

Nishimura, Takashi 466, 770, 781,<br />

1533, 1669, 1975<br />

Nishimura, Tatsuro 760, 1464, 1492<br />

Nishina, Sohji 995, 1937<br />

Nishino, Ken 2030, 2199<br />

Nishio, Akira 26, 1635, 1636, 1809,<br />

1982<br />

Nishio, Kumiko 1584<br />

Nishio, Takahiro 402, 1213<br />

Nishiyama, Kiyoshi 324<br />

Nishizaki, Yasuhiro 2177<br />

Nissen, Nicholas N. 1479<br />

Nitta, Sayuri 164, 406, 985, 1018,<br />

1959<br />

Nitta, Takeo 1901<br />

Nittono, Hiroshi 804, 1696<br />

Niu, Congrong 35, 2009<br />

Niu, Jing 754<br />

Niu, Junqi 1149, 1193, 1580, 1603,<br />

1620, 1657, 1869, 2018, 2058, 2107<br />

Niu, Xiaoxia 1536, 2008<br />

Nix, Sarah 335, 358<br />

Niyibizi, Nyiramugisha 1835<br />

Njouom, Richard 1833<br />

Njoya, Merlin 1597<br />

Nkuize, Marcel 1091<br />

Noel, Benoit 974<br />

Noel, Gillian 1737<br />

Nogata, Yoichi 918<br />

Noguchi, Kazunori 304<br />

Nogueira, Monize A. 240<br />

Noh, Choong Kyun 379, 1567, 2056<br />

Nojima, Masanori 1448, 1983<br />

Nojiri, Shunsuke 1448, 1954<br />

Nokes, Brandon 1459<br />

Noma, Naruto 1918<br />

Nomoto, Minoru 2165<br />

Nomura, Hideyuki 1097, 2031<br />

Nomura, Norihiro 1941<br />

Nomura, Yoshimoto 699, 1953<br />

Nong, Yunhong 1562<br />

Noon, Luke A. 691, 1364<br />

Noor, Bashir 2205<br />

Nordstrom, Beth L. 1175<br />

Norero, Blanca M. 1217<br />

Noronha Ferreira, Carlos 150<br />

Noronha-Dutra, Alberto A. 2237<br />

Norris, Suzanne 1103<br />

Northup, Patrick 761, 1258, 1501<br />

Notake, Tsuyoshi 869<br />

Notvall-Elkey, Lena M. 2022<br />

Noureddin, Mazen 2244<br />

Nousbaum, Jean-Baptiste 554, 2082<br />

Nouso, Kazuhiro 2236<br />

Novack, Victor 2224<br />

Noviello, Stephanie 706, 709, 711,<br />

716<br />

Nozaki, Yasutoshi 97, 231, 669, 1082,<br />

1357, 1661, 1985, 2130, 2136<br />

Nozaki, Yuichi 2207<br />

Nuamah, Rosamond 2105<br />

Nugent, FW 412<br />

Numthavaj, Pawin 1999<br />

Nunez, Marina 438<br />

Nuñez, Susana 899, 1320<br />

Nunn, Abigail D. 958<br />

Nusse, Roel 683<br />

Nwankwo, Chizoba 717, 727, 729<br />

Nyberg, Anders H. 87, 1788, 1825<br />

Nyberg, Lisa M. 87, 1788, 1825<br />

Nyberg, Scott L. 2<br />

Nydam, Trevor L. 838<br />

Ó Broin, Pilib 515<br />

O’Beirne, James 413, 489, 490, 764,<br />

793, 1455, 1530, 1904<br />

O’Bleness Gray, Nicole 1260<br />

O’Brien, April 830<br />

O’Brien, Christopher 2256<br />

O’Brien, Daniel R. 129, 1965<br />

O’Brien, Sean 396<br />

O’Donoghue, Pam 489, 490, 1904<br />

O’Grady, John G. 1744<br />

O’Hara, Steven P. 1994<br />

O’Leary, Jacqueline G. 82, 200, 570,<br />

1084<br />

O’Neil, Maura 352<br />

Oakes, Kath 646, 1098, 1113, 1588,<br />

2053, 2054<br />

Oakley, Fiona 44, 652, 1381, 1408<br />

Oaks, Zachary 1961<br />

Oberti, Frédéric 145, 263, 2188<br />

Obry-Roguet, Véronique 1093<br />

Ochi, Hidenori 986, 1056, 1645,<br />

1649, 1681, 1858, 1895, 1944<br />

Ochi, Noriyo 1573<br />

Ocho, Makoto 1448<br />

Oda, Kohei 799, 937, 1421, 1838<br />

Oda, Masaya 1514<br />

Odamaki, Toshitaka 2074<br />

Odena, Gemma 110, 1327<br />

Odin, Joseph A. 597, 607, 1095,<br />

1185, 1932<br />

Oe, Bibiana 316<br />

Oe, Shinji 593, 883<br />

Oehrle, Melissa 1695<br />

Oettl, Karl 2068<br />

Ofengeim, Dimitry 832<br />

Ofosu, Andrew 1749<br />

Ofotokun, Igho 1444<br />

Ofuji, Kazuya 509, 944


1346A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Ogasawara, Sachiko 1990<br />

Ogata, Kei 304, 1819<br />

Ogawa, Hirohisa 1981<br />

Ogawa, Koji 287, 1946<br />

Ogawa, Shintaro 1573, 1599<br />

Ogiso, Satoshi 1976<br />

Ogunwobi, Olorunseun 1967<br />

Ogura, Satoshi 227, 1974<br />

Ogwo, Cheri 1233<br />

Oh, Jae K. 1470<br />

Oh, Sohee 474, 1610<br />

Oh, Yumin 115<br />

Ohashi, Jun 1632<br />

Ohira, Hiromasa 551<br />

Ohira, Tetsuya 551<br />

Ohki, Takamasa 382, 2240<br />

Ohkusa, Toshifumi 2074<br />

Ohmae, Saori 1918<br />

Ohmoto, Masaki 524<br />

Ohmura, Takumi 1116<br />

Ohne, Kumiko 1573, 1599, 1618<br />

Ohno-Machado, Lucila 112<br />

Ohtake, Takaaki 927<br />

Ohtani, Masahiro 944, 1682<br />

Ohtsuka, Masato 1369<br />

Ohtsuru, Akira 551<br />

Oishi, Naoki 168, 699, 1348, 1653,<br />

1953, 1970<br />

Oja-Tebbe, Nancy 1578<br />

Ojiro, Keisuke 1001<br />

Oka, Takahito 2030<br />

Okabe, Haruka 314<br />

Okabe, Hirohisa 821<br />

Okada, Hikari 144, 1010, 1014, 1348,<br />

1377, 1395, 1653, 1953, 1968, 1970<br />

Okada, Toshihiro 188, 1420<br />

Okajima, Akira 886, 888, 2230<br />

Okajima, Hideaki 1213<br />

Okamato, Yuta 1642<br />

Okamoto, Koji 952<br />

Okamoto, Tomohiro 1420<br />

Okamoto, Toru 1643<br />

Okamoto, Toshiaki 1499<br />

Okano, Jun-ichi 1499<br />

Okanoue, Takeshi 707, 888, 2199<br />

Okazaki, Kazuichi 613<br />

Okimoto, Gordon S. 1553<br />

Okolicsanyi, Stefano 571<br />

Okoronkwo, Nneoma O. 295, 1181<br />

Okubo, Tomomi 291<br />

Okuda, Yukihiro 1213<br />

Okuno, Masayuki 402<br />

Okuse, Chiaki 1446<br />

Okushin, Kazuya 912<br />

Olcoz, Jose Luis 2171<br />

Olde Damink, Steven 604<br />

Olinga, Peter 1427, 1434<br />

Oliva, Giovanni 1476<br />

Olivares, Shantel 257<br />

Oliveira, Carlos 1900<br />

Oliveira, Claudia P. 240, 1033, 2164,<br />

2221<br />

Oliveira, Ketti G. 1855<br />

Oliveira, Mariana M. 1523<br />

Oliveira, Nuno G. 1921<br />

Oliveira, Yanaihara P. 2237<br />

Oliver, Gavin R. 1965<br />

Olivera-Martinez, Marco A. 748, 769<br />

Ollivier-Hourmand, Isabelle 145, 264,<br />

2082<br />

Oloruntoba, Omobonike O. 540<br />

Olsen, Mark 256<br />

Olson, Janet E. 397<br />

Olson, Jody C. 1291, 1814, 1815<br />

Olthoff, Kim M. 838, 853, 868<br />

Olyaee, Mojtaba S. 1814, 1815<br />

Olynyk, John K. 2185<br />

Omagari, Katsumi 1655<br />

OMahony, Megan 1172<br />

Omar, Rabab F. 1076<br />

Omata, Masao 1080, 1105<br />

Omori, Risa 1302, 1941<br />

Omura, Hitoshi 1000<br />

On, Sissi 887<br />

Onali, Simona 309, 413, 489, 490,<br />

1153<br />

Oniki, Kentaro 1912<br />

Onishi, Hideki 2236<br />

Onishi, Hiroka 799, 937, 1838<br />

Onishi, Yoshiki 26, 453, 1635, 1636,<br />

1809, 1982<br />

Onisto, Maurizio 299<br />

Onji, Morikazu 524<br />

Ono, Atsushi 986, 1649, 1681, 1895,<br />

1944<br />

Ono, Masafumi 2156, 2207<br />

Ono, Suzane K. 1544<br />

Onodera, Mio 1528<br />

Onorato, Lorenzo 431<br />

Onori, Paolo 621, 817, 834, 835<br />

Oo, Ye H. 1298, 1710, 2003<br />

Ooka, Kohtaro 1370<br />

Oosterhuis, Dorenda 1427, 1434<br />

Or, Yat Sun 2257, 2259, 2260<br />

Oribe, Junya 2267<br />

Oriishi, Tetsuharu 233<br />

Orimo, Tatsuya 866<br />

Orkin, Chloe 210<br />

Orlent, Hans 1091<br />

Orlicky, David J. 139, 142, 907, 921<br />

Orloff, Susan L. 69, 481<br />

Orman, Eric S. 1451, 1903<br />

Oró, Denise 659<br />

Orr, Anne 670, 1979<br />

Orr, David W. 2176, 2225<br />

Orr, James G. 2226<br />

Ortega, A. 583, 595, 1863<br />

Ortiz, Inmaculada 378<br />

Ortiz, Juan 78<br />

Ortiz, Pablo 2163<br />

Ortiz-Lasanta, Grisell 1128<br />

Oruç, Nevin 492<br />

Ory, Daniel S. 911<br />

Osafo-Addo, Awo 1731<br />

Osalde-Solís, Geraldine 1471<br />

Osawa, Yosuke 476<br />

Osgood, Gregory 200<br />

Oshige, Akihiko 799, 937, 1421, 1838<br />

Oshita, Masahide 2213<br />

Osinusi, Anu 92, 205, 220, 249, 713,<br />

1013, 2265<br />

Osna, Natalia A. 589, 1323, 1347<br />

Osorio, Robert W. 4, 538, 2104<br />

Ostrenko, Oleksandr 617<br />

Ota, Tsuguhito 495<br />

Otal, Philippe 263<br />

Otani, Satoshi 164, 406<br />

Otazua, Pedro 583<br />

Otero, P. A. 493<br />

Otgonsuren, Munkhzul 575, 892, 2173,<br />

2192, 2204<br />

Otsubo, Takeshi 2160<br />

Otsuka, Taiga 296<br />

Ott, Peter 768<br />

Otto, Benjamin 523<br />

Ottobrelli, Antonio 1<br />

Ottosen, Soren 2255<br />

Otvos, James 2227<br />

Ou, Xiaojuan 374<br />

Oude Elferink, Ronald 827<br />

Ouled Cheikh, Ibtissem 1621<br />

Oules, Valerie 1052<br />

Outhay, Malena 1223<br />

Ouwerkerk-Mahadevan, Sivi 39<br />

Ouyang, Elsa 1165<br />

Ouyang, Xinshou 805, 979, 1284,<br />

2179<br />

Ouzan, Denis 1052, 1847<br />

Overcash, J. Scott 41, 248, 250<br />

Overton, Edgar T. 1444, 1804<br />

Owens, Christopher M. 2257<br />

Owens, Jennifer 812, 822, 1363<br />

Owens III, Albert P. 81<br />

Owens-Grillo, Janet 644<br />

Oyakawa, Leticia 1855<br />

Ozaki, Yoshihiko 944, 1682<br />

Ozasa, Kotaro 551<br />

Ozaslan, Ersan 308<br />

Ozdogan, Osman C. 558, 1822<br />

Oze, Tsugiko 1195, 1197, 1201, 2213<br />

Ozeki, Itaru 1116, 1203, 1778, 2065<br />

Ozenne, Violaine 145<br />

Ozutemiz, Omer 246, 1247<br />

Pabst, Oliver 100<br />

Pacher, Pal 1374<br />

Pack, Michael 195, 814<br />

Padberg, Kornelius 183<br />

Padula, William V. 152<br />

Pagan, Sarah 611<br />

Page, Agata 1381<br />

Page, Kimberly 1793<br />

Pageaux, Georges-Philippe 3, 95, 1158<br />

Pai, Rish K. 107<br />

Paik, Seung Woon 422, 423, 485,<br />

1591, 1960, 2049<br />

Paik, Yong-Han 422, 485, 1415, 1591,<br />

1960, 1972, 2049<br />

Pais, Raluca 1239, 2164<br />

Pak, Judy 260, 753<br />

Pak, Stephen C. 193, 197<br />

Palaios, Emmanouil 850, 1271<br />

Palaniyappan, Naaventhan 1524<br />

Palazzo, Adam 2009<br />

Palazzo, Donatella 222, 983<br />

Palle, Sirish 12<br />

Palma, Elena 599<br />

Palmer, Daniel 344<br />

Pamecha, Viniyendra 695, 1268, 1340<br />

Pan, Calvin Q. 209, 1165, 1587,<br />

2015, 2052<br />

Pan, Chen 2023<br />

Pan, Jason J. 108<br />

Pan, Jen-Jung 1424<br />

Pan, Terry 1216<br />

Pan, Yang 919, 934<br />

Pan, Yu 1149, 1193, 1869, 2107<br />

Pan, Zhaoxing 1709, 1715, 1737<br />

Pan-ngum, Wirichada 555<br />

Panchal, Ankur 1397


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1347A<br />

Pandak, William M. 804<br />

Pande, Gaurav 274<br />

Pandey, Akshay 1988<br />

Pandit, Vanshja 1340, 2088<br />

Pandol, Stephen J. 904<br />

Panduro, Arturo 1855<br />

Pandya, Prashant K. 1814, 1815<br />

Pandya, Purva 510, 2116<br />

Panebianco, Deborah L. 730, 731<br />

Pang, Phillip S. 91, 96, 219, 247,<br />

1045, 1049, 1105, 1130, 1442, 1860<br />

Panigrahi, Rajesh 501, 506, 1096<br />

Pant, Kishor 1664<br />

Pantea, Victor 31<br />

Panzitt, Katrin 662<br />

Paoli, Pier P. 1381, 1408<br />

Papa, Salvatore 1978<br />

Papadaki, Sotiria 1489<br />

Papatheodoridis, George V. 241, 1126,<br />

2012, 2014, 2019<br />

Papiamonis, Nikolaos 2094<br />

Papp, Maria 2090<br />

Pappas, Stephen Chris 278<br />

Pappo, Orit 1878<br />

Paradis, Valerie 417, 480, 887<br />

Paraná, Raymundo 595<br />

Paranaguá-Vezozzo, Denise 1544<br />

Paranjpe, Shirish 670<br />

Pareja, María J. 2171<br />

Parekh, Krishan 1298<br />

Parés, Albert 73, 76, 308, 603, 615,<br />

620, 629, 634, 642<br />

Pariente, Alexandre 1072, 1072, 2082<br />

Parikh, Neehar D. 67, 395, 418, 1850<br />

Parisé, Hélène 1087, 1143<br />

Parisi, Ioanna 793, 1530<br />

Park, Beom Jin 433, 1438<br />

Park, Brandi 1835<br />

Park, Do Youn 1025<br />

Park, Haesuk 1079, 1538<br />

Park, Hana 1542<br />

Park, Hyeongmin 491<br />

Park, In-Whee 2056<br />

Park, James 1165<br />

Park, Jang-June 1001<br />

Park, Jin-Kyu 1322<br />

Park, Jong Sung 115<br />

Park, Joohan 1567<br />

Park, Joong-Won 377, 486, 851<br />

Park, Jun Yong 346, 1542, 1558, 1581<br />

Park, Jung Gil 363, 403, 749<br />

Park, Kyoung-Sook 950<br />

Park, Kyu-Sang 1401<br />

Park, Nomi 464<br />

Park, Ogyi 115<br />

Park, Rae Woong 2056<br />

Park, Seol-Hee 1358<br />

Park, Seong W. 658<br />

Park, So Hyun 807<br />

Park, Soo Young 363, 403, 426, 749,<br />

2028, 2038<br />

Park, Soohyun 1415, 1972<br />

Park, Su Cheol 685<br />

Park, Su-Hyung 1025<br />

Park, SuBum 876<br />

Park, Sun Seob 377<br />

Park, Sun Young 1567<br />

Park, Tae Jun 187<br />

Park, Young Nyun 1902<br />

Parker, Richard 574, 1321<br />

Parkes, Julie 776, 793, 1455, 1530,<br />

1856<br />

Parlak, Erkan 846<br />

Parrella, Korin 1151, 1870<br />

Parruti, Giustino 1107<br />

Parsons, Judith 1023<br />

Parvin-Nejad, Fatemeh P. 1364<br />

Pas, Suzan D. 1590, 2122<br />

Pascasio, Juan Manuel 2032<br />

Pascual, Sonia 378<br />

Pasetto, Maria Cristina 309, 1153<br />

Pasic, Fuad I. 1188<br />

Pasquale, Giuseppe 1543<br />

Pasquazzi, Caterina 222<br />

Pasquet, Blandine 145<br />

Passmann, Sandra 1559<br />

Passos-Castilho, Ana Maria 48<br />

Pastor, Helena 2171<br />

Pastor, Jose J. 194<br />

Patch, David W. 413, 764, 793, 1530<br />

Patel, Anna 1151<br />

Patel, Arpan A. 1512<br />

Patel, Dilip 260<br />

Patel, Forum 606<br />

Patel, Janki R. 913<br />

Patel, Keyur 205, 624, 1428, 1887,<br />

2169<br />

Patel, Neal M. 1095, 1185, 1202<br />

Patel, Rahul 1151<br />

Patel, Rajiv 1787<br />

Patel, Ruchi 1720<br />

Patel, Sameer 1775<br />

Patel, Samir 2147, 2208<br />

Patel, Trushar 1267, 2081<br />

Patel, Tushar 335, 358, 384<br />

Patel, Vishal C. 202, 740<br />

Patel, Mamta 1470<br />

Paternostro, Rafael 149, 1466<br />

Pathil, Anita 1411, 2195<br />

Patidar, Kavish 2235<br />

Patil, Mallikarjun 284, 2101<br />

Patra, Sharda 1663<br />

Pattanasirigool, Chaowalit 1631, 1662<br />

Patten, Daniel A. 1296<br />

Pattou, Francois 2222<br />

Paugam-Burtz, Catherine 1781<br />

Paul, Andreas 1779<br />

Paul, Jean-Louis 887<br />

Paul, Sonali 1592<br />

Paule, Bernard 421, 449<br />

Paulino, Lismeiry 545<br />

Pauta, Montserrat 65, 659<br />

Pauwels, Arnaud 264<br />

Pavel, Mihai-Calin 860<br />

Pavel, Oana 734<br />

Pavendranathan, Gokulan 1243<br />

Pavkov, Douglas 1172<br />

Pavletic, Steven 2144<br />

Pavlides, Michael 2190<br />

Pavlovic, Julie 1225<br />

Pavlovic, Vedran 245, 2026<br />

Pavone, Nicole 536<br />

Pavuk, Marian 2233<br />

Pawlik, Timothy M. 356<br />

Pawlinski, Rafal 81<br />

Pawlotsky, Jean-Michel 693, 987, 999,<br />

1123, 1700, 1797<br />

Pawlowska, Malgorzata 2043<br />

Payance, Audrey 480, 772, 2148<br />

Payer, Berit A. 1204<br />

Payerl, Doris 2068<br />

Payungporn, Sunchai 2063<br />

Péan, Noémie 816<br />

Pearlman, Brian 715<br />

Peck-Radosavljevic, Markus 37, 149,<br />

252, 270, 475, 561, 732, 755, 771,<br />

1049, 1058, 1092, 1204, 1466, 1525<br />

Pecriaux, Caroline 1194<br />

Peddu, Praveen 212<br />

Pedersen, Mark 1043, 1109, 1145,<br />

1224, 1241, 1262<br />

Pediconi, Natalia 958<br />

Pedley, Alison 2191<br />

Pedro, Ana Júlia 150<br />

Pedrosa, Marcos C. 16, 1106<br />

Peeples, Miranda 711<br />

Peeraphatdit, Thoetchai 397, 1497<br />

Pei, Luying 2007<br />

Peiffer, Kai-Henrik 980, 1559<br />

Peiser, Leanne 2009, 2052<br />

Pelaquier, Agnes 2082<br />

Pelemis, Mijomir 2004<br />

Pellegatta, Gaia 355<br />

Pelletier, Gilles 1237<br />

Pelletier, Shawn 1258<br />

Pellicoro, Antonella 62<br />

Peltekian, Kevork M. 204<br />

Peña, Jose M. 1881<br />

Penaranda, Guillaume 1009<br />

Pencek, Richard 155, 625, 628, 635,<br />

644, 645, 1458<br />

Pendergast, Jason 573<br />

Penders, John 100<br />

Pene, Veronique 30, 997<br />

Peng, Cheng-Yuan 40, 42, 1828, 1996,<br />

2040<br />

Peng, Haoran 353, 375<br />

Peng, Hong 1583<br />

Peng, Jie 1998<br />

Peng, Jin 828, 1924<br />

Peng, Kesong 22<br />

Peng, Lee 1122<br />

Peng, Lee F. 1015<br />

Peng, Liang 1423, 1772<br />

Peng, Lijun 131<br />

Peng, Shan S. 2107<br />

Peng, Yan-Zhong 2023<br />

Peng, Zhechu 1945<br />

Peng, Zhen-Wei 826, 1367, 1368,<br />

1379<br />

Penn, Michael 1177<br />

Penna, Giuseppe 893<br />

Pennell, Craig E. 2185<br />

Pepe, Veronica 1270<br />

Peppa, Dimitra 1626<br />

Perales, Celia 1028<br />

Peralta, Carmen 8, 326, 342<br />

Perarnau, Jean-Marc 145<br />

Perazzo, Hugo 786, 2189<br />

Perdigoto, Rui 1256<br />

Pereira, Cristiano A. 2200<br />

Pereira, Gustavo 84, 762, 2072, 2076<br />

Pereira, Isabel V. 586<br />

Pereira, João Luiz 84, 762<br />

Pereira, Marcus 272<br />

Pereira, Stephen P. 76<br />

Pereira, Thiago A. 936<br />

Perez, Renata D. 762<br />

Perez, Renata M. 84, 273, 289, 2189<br />

Perez-Cormenzana, Miriam 620


1348A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Perez-Hernandez, Onan 1341, 1342<br />

Perez-Rodriguez, Edelmiro 1281<br />

Perichon, Regis 161<br />

Perilla, Mauricio 1271<br />

Perinelli, Paola 222, 983<br />

Perini, Lisa 629<br />

Peris, Pilar 615<br />

Perito, Emily R. 1730, 1736<br />

Perkins, Neil D. 652<br />

Perkins, Susan 436<br />

Perl, Andras 1961<br />

Perlmutter, David H. 193, 197<br />

Pernazza, Alejandra 2106<br />

Perner, Dany 980, 1559<br />

Perno, Carlo Federico 203, 222<br />

Péron, Jean-Marie 263, 1803, 2082<br />

Perrone, Leonardo 1089<br />

Persico, Marcello 1802, 1844<br />

Perumalswami, Ponni V. 545, 1095,<br />

1151, 1185, 1202, 1783<br />

Perumpail, Ryan B. 199, 201, 541,<br />

877, 1534, 1745, 1750<br />

Pesci, David 1451<br />

Pessoa, Mario G. 1033<br />

Peter, Joy A. 1826<br />

Peter, Senaka 1867, 1883<br />

Peters, Marion G. 1094, 1444, 1570<br />

Peters, Yvette 644<br />

Petersen, Dennis R. 921<br />

Petersen, Joerg 37, 241, 252, 1058,<br />

1078, 1176, 1205, 1559, 2003, 2014<br />

Peterson, Ryan M. 32, 2022<br />

Petit, Laetitia Marie 1735<br />

Petoumenos, Kathy 1083<br />

Petraccia, Luisa 1190<br />

Petraksa, Supanna 1999<br />

Petrasek, Jan 180<br />

Petropoulou, Fotini 1126<br />

Petrov-Sanchez, Ventzislava 362, 1808,<br />

1847<br />

Petrtyl, Jaromir 1487<br />

Petta, Salvatore 973, 2164<br />

Peyret, Thomas 1915<br />

Peyton, Adam 1041, 2256<br />

Peyton, Diane 536<br />

Pflanz, Stefan 2009<br />

Pfreundschuh, Michael 463<br />

Phakdeekitcharoen, Bunyong 1999<br />

Pham, Huong T. 1186<br />

Phan, Jennifer 488<br />

Phaosawasdi, Kamthorn 416, 555<br />

Phelan, Kieran 676<br />

Phelps, Janet R. 2007<br />

Philip, Philip A. 420<br />

Philips, Cyriac A. 103, 2070<br />

Phillips, John D. 2111, 2112<br />

Phillips, Sandra 163<br />

Philosophe, Benjamin 1214<br />

Pho, Mai T. 152<br />

Phung, Van 1986<br />

Pi, Liya 989<br />

Pianko, Stephen 91, 249<br />

Piano, Salvatore 148, 266, 292, 767<br />

Piao, Shenglong 113<br />

Piard-Ruster, Karine 341<br />

Piazza, Nicholas 1257<br />

Picard, Nicolas 1162<br />

Picat, Marie-Quitterie 132<br />

Picchio, Gaston 39<br />

Piconese, Silvia 958<br />

Pieper, Dietmar H. 1002<br />

Pierce, Ruth 17<br />

Pieri, Giulia 1751<br />

Pilette, Christophe 1072<br />

Pillai, Anjana 369<br />

Pillardy, Jaroslaw 1841<br />

Pilot-Matias, Tami 705, 714, 1065,<br />

1084, 1086<br />

Piluso, Alessia 1190<br />

Pilutti, Chiara 148, 266, 292<br />

Pimentel, Carolina F. 1484, 1494,<br />

1495, 2160, 2227<br />

Pine, Karen 1452<br />

Pineton de Chambrun, Marc 1288<br />

Pinho, João Renato R. 48, 1033, 1855<br />

Pinkston, Christina 2233<br />

Pinna, Antonio Daniele 1751<br />

Pinotti, Rachel 2170<br />

Pinzani, Massimo 337, 413, 517, 764,<br />

1373, 1388, 1407, 1410, 1445, 1517<br />

Pio, Jose R. 87<br />

Piotrowski, Joy I. 1121<br />

Piovesan, Sara 775<br />

Piratvisuth, Teerha 1631, 1662<br />

Pirillo, Martina 1888<br />

Pirmohamed, Munir 225<br />

Pirola, Carlos J. 2216, 2217<br />

Pisano, Giuseppina 393<br />

Pisano, Maia Belen 1752<br />

Pisaturo, Mariantonietta 1543<br />

Piscaglia, Fabio 355, 602<br />

Pischke, Sven 523, 1803<br />

Pitrone, Alessia 288<br />

Pizarro, Margarita 969<br />

Plankey, Michael 1444<br />

Plat, Jogchum 922, 961<br />

Platt, Blake 71, 843<br />

Platt, Heather L. 40, 210, 380, 438,<br />

729<br />

Plompen, Elisabeth P. 1590<br />

Plotnik, Julia N. 626, 1790, 1798,<br />

1807<br />

Plumeier, Iris 1002<br />

Poca, Maria 734<br />

Pocha, Christine 16, 1884<br />

Pockros, Paul J. 94, 108, 251, 609,<br />

1039, 1057<br />

Poddar, Minakshi 255<br />

Podesser, Bruno K. 1525<br />

Podevin, Philippe 30<br />

Podlaha, Ondrej 1651, 1668, 2052<br />

Podsadecki, Thomas 1039, 1086, 1161<br />

Poerzgen, Peter 1121<br />

Pogemiller, Hope 572<br />

Pohl, Sabine 1519<br />

Pohnert, Georg 808<br />

Poiteau, Lila 999, 1833<br />

Poizot-Martin, Isabelle 568, 1093<br />

Pokora-Pachowicz, Agnieszka 1582<br />

Pokorski, Izabella 1343, 1450<br />

Pol, Stanislas 1194, 1808, 1847, 1860<br />

Polak, Wojciech G. 68<br />

Polepally, Akshanth R. 1051, 1066<br />

Pollard, Katherine A. 2121<br />

Pollicino, Teresa 1127, 1654, 1898<br />

Polo, Miriam 1913<br />

Polotsky, Vsevolod 1422<br />

Poluektova, Larisa Y. 589<br />

Polvani, Simone 949, 1963<br />

Polychronidis, Georgios 1227<br />

Pomfret, Elizabeth A. 71, 838, 843,<br />

847<br />

Pomper, Martin 115<br />

Pompili, Maurizio 432<br />

Pons, Fernando 378, 1511<br />

Ponsioen, Cyriel Y. 73, 74, 76, 604,<br />

623, 634<br />

Ponzoni, Mirco 44<br />

Poole, Lauren G. 587, 2142<br />

Poon, Art 1832<br />

Poon, Kok Siong 1812<br />

Poongkuran, Mugilan 371<br />

Poonia, Bhawna 1029<br />

Poordad, Fred 39, 41, 247, 248, 250,<br />

706, 716, 1051, 1065, 1086, 1433,<br />

2257, 2260<br />

Poovorawan, Kittiyod 416, 555, 1631<br />

Poovorawan, Yong 2036, 2063<br />

Pop, Oltin T. 1763<br />

Pope, Amanda 1716<br />

Popov, Yury 826, 1367, 1368, 1379<br />

Porat-Shliom, Natalie 20<br />

Porayko, Michael K. 279<br />

Porcella, Rita 309<br />

Porcherot, Marine 1652<br />

Pornthisarn, Bubpha 1631, 1662<br />

Porsche, Cara 938, 959<br />

Port, Kerstin 1131<br />

Portal, Andrew J. 564<br />

Porte, Robert J. 68, 1258<br />

Porter, John R. 195, 814<br />

Posa, Alessandro 432<br />

Pose, Elisa 266<br />

Posner, Amanda J. 1736<br />

Posthouwer, Dirk 1613<br />

Posuwan, Nawarat 2063<br />

Potenza, Nicoletta 431<br />

Poterucha, John J. 836, 1263<br />

Potosky, Darryn R. 1101, 1215<br />

Potze, Wilma 2249<br />

Poulsen, Kyle L. 1339, 1346<br />

Poupon, Raoul 73, 634, 642<br />

Pouriki, Sophia 1489<br />

Pourmand, Kamron 1826<br />

Pouteau, Michele 2044, 2045<br />

Povero, Davide 57, 969, 1371, 1971<br />

Powell, Elizabeth E. 784, 1362<br />

Poynard, Thierry 786, 1510<br />

Pradat, Pierre 568, 1093, 1162, 2003<br />

Prado, Veronica 1307<br />

Prall, Stacy 552<br />

Pratschke, Johann 276, 1222<br />

Pratte, Katherine 174<br />

Precoma, Dalton B. 2248<br />

Preisinger, Christian 1924<br />

Premkumar, Madhumita 1344, 2078<br />

Premoli, Caterina 1747<br />

Premont, Richard T. 66, 936, 1386<br />

Preston, Ta-Wanda 1864, 1880<br />

Pretorius, Carel J. 784<br />

Price, Angie 92<br />

Price, Jackie 2168<br />

Price, Patricia 2185<br />

Prieto, Martin 583, 596, 1230, 1251,<br />

2032, 2171<br />

Prins, Maria 1846<br />

Prip-Buus, Carina 887<br />

Pritilata, Rout 284<br />

Protopopas, George 1181<br />

Protzer, Ulrike 1684


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1349A<br />

Prough, Russell A. 666, 2233<br />

Provenzale, Dawn T. 540<br />

Provenzano, Angela 1410<br />

Puente, Angela 1511<br />

Puerto, Marta 1295, 1520<br />

Pugasab, Alongkorn 1999<br />

Pugatch, David 710<br />

Pugliese, Pascal 568, 1093<br />

Pullicino, Richard 2251<br />

Püngel, Tobias 1756<br />

Pungpapong, Surakit 1038, 1124,<br />

1236<br />

Punpanich, Dollapas 2089<br />

Puntes, Víctor 659<br />

Puoti, Massimo 1802, 1844<br />

Purcell, Robert H. 1646<br />

Puri, Puneet 52, 240, 905, 1309, 1316,<br />

1340, 1472, 2163, 2235<br />

Purnak, Tugrul 308<br />

Pursell, Natalie W. 510, 2116<br />

Pustavoitau, Aliaksei 1255<br />

Puszyk, William M. 1967<br />

Putignano, Antonella 1769, 1780, 1781<br />

Pyenson, Bruce 1069<br />

Pyrsopoulos, Nikolaos 295, 1181, 1827<br />

Qamar, Amir A. 412<br />

Qaqish, Roula B. 708, 714<br />

Qi, Xin 651<br />

Qi, Xingshun 754<br />

Qi, Xun 1998, 2002<br />

QI, Lingxia 2107<br />

Qin, Shengying 225<br />

Qin, Xian-Yang 657<br />

Qiu, Chunhui 353<br />

Qiu, Xinjie 919<br />

Qu, Jun 332<br />

Qu, Xiaowang 1001<br />

Qu, Xiumei 2123<br />

Quaglia, Alberto 1733, 1978<br />

Qualls, Charles 1916<br />

Quan, David J. 1560<br />

Quaranta, Maria Giovanna 1153,<br />

1802, 1844<br />

Quer, Josep 1021, 1674<br />

Quertinmont, Eric 178<br />

Questel, Franck 1102<br />

Quinn, Gemma 39<br />

Quinn, Matthew A. 1507<br />

Quintana, Albert 1295<br />

Quintero-Platt, Geraldine del Carmen<br />

1341, 1342<br />

Quintini, Cristiano 840, 1271<br />

Quispe, Wilber 515<br />

Qureshi, Huma 1872<br />

Qureshi, Kamran 1122<br />

Rabatin, Abigail 1142<br />

Rachakonda, Vikrant 1496, 1531<br />

Racila, Andrei 575<br />

Radaeva, Svetlana 1332<br />

Radenne, Sylvie 3, 95, 1158<br />

Radford, Peggy 1716<br />

Radhakrishnan, Kavita 1560<br />

Radreau, Pauline 1652<br />

Radu, Claudia M. 351<br />

Raffa, Giuseppina 1654, 1898<br />

Raffetti, Elena 1568<br />

Raftopoulos, Spiro C. 1264<br />

Raghunathan, Sahana 330<br />

Rahimi, Robert S. 1210<br />

Rahman, Adeeb 988, 1289<br />

Rai, Nitish 259<br />

Rai, Rahul 2245<br />

Raimondo, Giovanni 288, 1127, 1476,<br />

1654, 1802, 1844, 1898<br />

Rainteau, Dominique 816<br />

Rajagopalan, Ramakrishnan 1693<br />

Rajanayagam, Jeremy K. 1710, 1728<br />

Rajesh, S. 1340<br />

Rajwanshi, vivek 993<br />

Rakela, Jorge 211, 1761<br />

Rakhmanaliev, Elian 1812<br />

Rallier, Sandrine 1833<br />

Ramachandran, Priya 867, 1698<br />

Ramakrishna, Gayatri 327, 695, 1005<br />

Ramalho, Fernando 150<br />

Ramanathan, Meera 1145, 1241, 1262<br />

Ramani, Komal 594<br />

Rametta, Raffaela 2187<br />

Ramirez, Gilberto 1574<br />

Ramirez-Vega, Ruben 1199<br />

Ramji, Alnoor 722<br />

Ramos, Hilario 35<br />

Rana, Abbas 369<br />

Rana, Khurram 716, 726<br />

Ranabhat, Ganesh 80<br />

Ranchal, Isidora 1300<br />

Rao, Anuradha 19<br />

Rao, Huiying 782, 1436, 1820, 1850,<br />

1877<br />

Rao, Kiran V. 1181<br />

Rao, Swati 1122<br />

Rao, Vinaya 1233<br />

Raouf, Ahmed A. 1853<br />

Raoul, Jean-Luc 394<br />

Rapaccini, Gian Ludovico 355, 432<br />

Raseni, Alan 2098<br />

Rashid, Asif 130, 682, 2118<br />

Rasineni, Karuna 1310<br />

Raskopf, Esther 80<br />

Rastellini, Cristiana 1991<br />

Rastogi, Archana 327, 695, 1483,<br />

1762<br />

Ratziu, Vlad 105, 162, 737, 739, 786,<br />

1428, 1435, 2145, 2164, 2239<br />

Rau, Monika 2195<br />

Rauch, Andri 1575, 1818<br />

Raun, Loren 1991<br />

Rautou, Pierre-Emmanuel 81, 145, 265,<br />

772, 887, 1288, 2148<br />

Ravendhran, Natarajan 715<br />

Ravenelle, Francois 1333<br />

Ravindran, Karthik 420<br />

Rawal, Bhupendra 1245<br />

Rawat, Dinesh 1697, 1725<br />

Rawls, John F. 2188<br />

Ray, Kristen 1831<br />

Rayar, Michel 487<br />

Raybon, Joseph J. 728<br />

Raychaudhuri, Pradip 1988<br />

Rayhill, Stephen C. 1211<br />

Raymond, Daniel 1839<br />

Raymond, Eric 417<br />

Raymond, Valérie-Ann 179<br />

Raza, Roshan 136<br />

Razavi, Homie 1818, 1842, 1859,<br />

1872<br />

Razavi-Shearer, Devin M. 1872<br />

Razumilava, Nataliya 1994<br />

Re, Viviana 1752<br />

Reau, Nancy 205, 271, 522, 539, 544,<br />

553, 1065, 1813<br />

Rechling, Christian 1137<br />

Reddy, K. Gautham 271, 539, 544,<br />

553<br />

Reddy, K. Rajender 82, 217, 570, 712,<br />

743, 1034, 1039, 1045, 1049, 1057,<br />

1130, 1134, 1442, 1759, 1768<br />

Reddy, Mettu S. 859, 867<br />

Reddy, Sheela S. 1452<br />

Reding, Raymond 1734<br />

Redman, Rebecca 708, 714<br />

Reebye, Vikash 118<br />

Reesink, Hendrik W. 208, 251, 1633,<br />

1995, 2005, 2255, 2263<br />

Reeves, Helen 425<br />

Reggiani, Paolo 1747<br />

Regula, Jaroslaw 609<br />

Rehermann, Barbara 1650<br />

Rehman, Saad 1223<br />

Reiberger, Thomas 149, 270, 475, 561,<br />

732, 755, 771, 1204, 1466, 1525<br />

Reich, David J. 202<br />

Reichman, Trevor W. 1242<br />

Reid, Danielle 2143<br />

Reig, Ana 629<br />

Reig, Anna 603, 620<br />

Reig, Maria Elisa 378<br />

Reiling, Janske 818, 975<br />

Reiller, Brigitte 1796<br />

Reilly, Briget 536<br />

Reimundo, Pilar 1674<br />

Rein, David B. 156, 1792, 1839<br />

Reinoso, Humberto 318<br />

Reisch, Thomas 705<br />

Reisinger, Florian 2128<br />

Reissing, Johanna 828<br />

Rela, Mohamed 859, 867, 1698<br />

Ren, Hong 1044, 1647, 2048<br />

Ren, Jinma 774<br />

Ren, Junping 984<br />

Ren, Suping 33, 2064<br />

Ren, Wanhua 516<br />

Ren, Xiaowei 1316<br />

Rendon, Paloma 378<br />

Renelus, Benjamin D. 790<br />

Rengasamy, Asokan 1692, 1699<br />

Renner, Eberhard L. 11, 204, 542<br />

Rennison, Julie H. 518, 651, 667<br />

Renou, Christophe 1072<br />

Rentero Alfonso, Carles 64<br />

Renz, John F. 1229<br />

Requena, Silvia 1180<br />

Resche-Rigon, Matthieu 998, 1601<br />

Rescigno, Maria 893<br />

Reuben, Adrian 1914<br />

Reul, Winfried 961<br />

Reverter Segura, Enric 982<br />

Revill, Peter A. 1551, 1557, 1563<br />

Rewisha, Eman A. 1862, 1879<br />

Rey, David 1093<br />

Rey, Jorge 2106<br />

Reyes, Ella F. 71, 843, 847<br />

Reyes, Jorge 878, 1211<br />

Reynaert, Hendrik 1091<br />

Reynolds, Catherine 530<br />

Reynolds, Justin A. 1111, 1155, 1164<br />

Rezvani, Milad 98<br />

Rhame, Frank S. 1034<br />

Rhee, David B. 515


1350A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Rhodes, Kimberly 1099, 1857<br />

Riachi, Ghassan 264<br />

Riad, Joseph 1076<br />

Ribas, Vicent 899, 1320<br />

Ribeiro, Roberto M. 2221<br />

Ribera, Jordi 65, 659<br />

Riccardi, Laura 432<br />

Rice, Kenner 1374<br />

Rice, Thomas M. 1793<br />

Richards, Lisa M. 913, 1431<br />

Richards, Tara D. 216<br />

Richardson, Crit T. 279<br />

Richardson, Jason R. 919, 934<br />

Richardson, Peter 90, 156, 244<br />

Richen, Xavier 1796<br />

Richou, Carine 1036, 1187<br />

Riedl, Florian 1525<br />

Rieger, Armin 561, 1204<br />

Rieger, Megan 1945<br />

Riese, Peggy 1762<br />

Rietdijk, Svend 208<br />

Rife, Kelsey 1154<br />

Rigamonti, Cristina 1743<br />

Riggio, Oliviero 2067<br />

Rigou, Geraldine 974<br />

Rijckborst, Vincent 1585<br />

Rijnders, Bart J. 1007<br />

Rikner, Leif 810<br />

Rimassa, Lorenza 602<br />

Rimland, David 438, 1488<br />

Rimola, Jordi 860<br />

Rinaldo, Piero 2<br />

Rinaudo, Jo Ann 371<br />

Rincón, Diego 1503<br />

Rinella, Mary E. 1215, 1226, 1235<br />

Ringers, Jan 68, 1238<br />

Rinker, Franziska 2026<br />

Rinninella, Emanuele 432<br />

Riordan, Stephen 106, 627<br />

Rios, Miguel 734<br />

Ríos-Torres, Silvia L. 1471<br />

Ripoll, Cristina 1503, 1519, 1520<br />

Riquelme, Arnoldo 2210, 2231<br />

Rispoli, Rossella 1978<br />

Ritzenthaler, Jeffrey D. 2142<br />

Rius, Bibiana 2087<br />

Riva, Alessandra 1751<br />

Riva, Antonio 599, 2143<br />

Rivas-Estilla, Ana Maria G. 1022<br />

Riveiro-Barciela, Mar 1021, 1674<br />

Rivera, Elenita M. 218<br />

Rivino, Laura 167<br />

Rizvi, Bisharah S. 2144<br />

Rizwan, Maira 2100, 2114<br />

Rizza, Giorgia 1<br />

Rizzetto, Mario 1, 1802, 1844, 2209<br />

Ro, Weonsang S. 504<br />

Roa, Jhoanna 1094<br />

Roach, Jonathan 1692, 1699<br />

Roayaie, Sasan 851<br />

Robaeys, Geert 1613, 1842<br />

Robbins, Kristen N. 1715<br />

Robertazzi, Suzanne 1101<br />

Roberts, Benjamin R. 1291<br />

Roberts, Dean W. 1914<br />

Roberts, Jackson L. 1407<br />

Roberts, John P. 4, 50, 53, 72, 414,<br />

1219, 1273, 1560<br />

Roberts, Lewis R. 121, 127, 129, 169,<br />

381, 396, 397, 415, 851, 1124,<br />

1497, 1965, 1966<br />

Roberts, Mark S. 104, 151<br />

Roberts, Sheree 1298<br />

Roberts, Stuart K. 106, 249, 442, 483,<br />

627, 1077, 2262<br />

Roberts, Teri 1826<br />

Robertson, Avril A. 55<br />

Robertson, Michael 40, 42, 210, 251,<br />

700, 701, 703, 707, 712, 715, 717<br />

Robic, Marie Angele 145<br />

Robinson, Benjamin 337, 1388<br />

Robinson, Joseph L. 1479<br />

Robles, Camila 2210<br />

Robles Diaz, Mercedes 583, 595, 596<br />

Robson, Richard A. 1128<br />

Robson, Simon C. 608, 826, 1222,<br />

1394, 2227, 2228<br />

Rocha, Chiara 1742<br />

Rocha, Clarissa S. 63<br />

Roche, Bruno 1237<br />

Rockey, Don C. 737, 739, 1478, 1922,<br />

1931<br />

Rockstroh, Jürgen K. 37, 210, 252, 700,<br />

703, 1058, 1081, 1156<br />

Rodell, Tim 2015, 2052<br />

Roder, Heinrich 1900<br />

Roder, Joanna 1900<br />

Rodes, Juan 342<br />

Rodgers, Joel B. 1804<br />

Rodrigo-Torres, Daniel 1327<br />

Rodrigues, Cecília M. 176, 832, 897,<br />

940, 960<br />

Rodrigues, Livia 240, 2221<br />

Rodrigues, Pedro M. 176, 832, 897,<br />

940, 960<br />

Rodrigues, Teresa 150<br />

Rodriguez, Carla V. 1811, 1871<br />

Rodriguez, Christophe 1700<br />

Rodriguez, Eduardo A. 1206<br />

Rodríguez, Manuel 378, 2032<br />

Rodriguez de Miguel, Cristina 2084<br />

Rodriguez-Agudo, Daniel 804<br />

Rodriguez-Barradas, Maria C. 1488<br />

Rodriguez-Canales, Jaime 166<br />

Rodríguez-Córdova, Paola A. 1471<br />

Rodriguez-Frias, Francisco 1021, 1674<br />

Rodriguez-Garcia, Jose Luis 1508<br />

Rodriguez-Lope, Carlos 378<br />

Rodriguez-Torres, Maribel 718<br />

Rodt, Thomas 444<br />

Roeker, Lindsey 1263<br />

Rogal, Shari S. 1753<br />

Rogers, Jason 1151<br />

Roggendorf, Michael 31, 1021<br />

Rogler, Charles E. 515<br />

Rohatgi, Sarika 1718<br />

Rohel, Alexandra 3<br />

Rohr, Cristian O. 2217<br />

Rohrbach, Jeff 536<br />

Rojas, Angela 1863<br />

Rojas de Santiago, Pamela 969, 971<br />

Rojas Luengas, Vanessa 1012, 1297<br />

Rokosh, S. Rae 1282<br />

Rollin, Francois 1835<br />

Romagnoli, Renato 1, 1740<br />

Romain, Mélissa 887<br />

Roman, Eva 583, 1509<br />

Roman, Jesse 2142<br />

Roman, Sonia 1855<br />

Romano, Antonietta 148, 266, 292,<br />

775<br />

Rombouts, Krista 337, 517, 908, 1373,<br />

1388, 1407, 1410, 1445, 1517<br />

Romero, Miriam 1028<br />

Romero-Gomez, Manuel 105, 162,<br />

1134, 1863, 2145, 2164, 2171<br />

Romero-Marrero, Carlos J. 1894<br />

Rondon, Juan 2257<br />

Rong, Guanghua 354<br />

Ronot, Maxime 417<br />

Rooge, Sheetalnath B. 690, 694, 1483,<br />

1659<br />

Roos, Ruud A. 1667<br />

Roosblad, Jimmy 1846<br />

Rorive, Sandrine 178<br />

Rosa, Isabelle 132, 206, 567, 1072,<br />

2082<br />

Rosa, Jose Luis 1418<br />

Rosa, Julia S. 2075<br />

Rosales-Zabal, Jose Miguel 1863<br />

Rosato, Stefano 1802, 1844<br />

Rose, Christopher F. 51, 1481, 1520,<br />

1523<br />

Rosen, Charles B. 836<br />

Rosen, Hugo R. 28, 503, 938, 959,<br />

1430, 1706, 1708, 2085<br />

Rosenau, Jens 200<br />

Rosenberg, Arielle R. 30, 997<br />

Rosenberg, Gillian 1557, 1563<br />

Rosenberg, William M. 249, 776, 793,<br />

1455, 1530, 1856, 1904<br />

Rosenthal, Philip 133, 1694, 1730,<br />

1736<br />

Rosi, Silvia 148, 292<br />

Rosmorduc, Olivier 132, 417<br />

Ross, David 128, 540<br />

Rossaro, Lorenzo 1041<br />

Rosselli, Matteo 113<br />

Rossi, Francesca 1837<br />

Rossi, Giorgio 1743, 1751<br />

Rossi, John 118<br />

Rossi, Ric 2099<br />

Rossi, Simona 530, 535<br />

Rossi, Stephen 106, 198, 627, 1986<br />

Rossignol, Emilie 3<br />

Rosso, Chiara 2209<br />

Rosso, Natalia 1339, 1346, 2098<br />

Rota, Matteo 571, 1453<br />

Rota, Monica 1453<br />

Roth, Christopher G. 1906<br />

Roth, David 700, 703, 1054, 1148<br />

Roth, Nitzan 1350<br />

Rothenberg, Marc E. 887<br />

Rothstein, Kenneth D. 202<br />

Rothweiler, Sonja 826<br />

Rotllan, Noemi 65<br />

Rotman, Yaron 218, 2224<br />

Rouach, Hélène 30<br />

Roudot, Alice 105, 162, 974, 2145<br />

Roudot-Thoraval, Françoise 362, 421,<br />

449, 567, 1232, 1808, 1847<br />

Rougemont, Anne-Laure 1735<br />

Rougier, Hayette 1676, 1677<br />

Rouhani, Farshid 138<br />

Roura-Poch, Pere 2084<br />

Rousseau, Annick 1162<br />

Rousseau, Benoit 693<br />

Rouveau, Nicolas 1833


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1351A<br />

Roux, Olivier 265, 772, 1781, 2148<br />

Rowe, Ian A. 574, 1321<br />

Rowell, Richard 1478<br />

Roychowdhury, Sanjoy 2132<br />

Roytman, Marina 1121<br />

Roytrakul, Sittiruk 1999<br />

Rødgaard-Hansen, Sidsel 2071<br />

Ruane, Peter J. 205, 248, 716, 1065,<br />

1076<br />

Rubbia-Brandt, Laura 2174<br />

Rubenstein, Kevin 1811, 1871<br />

Rubin, Angel 1230, 1251<br />

Rucci, Paola 983<br />

Rude, Eric J. 1801<br />

Rude, Mary K. 2091<br />

Rudler, Marika 145, 744, 1510<br />

Rudnick, Sean R. 1501<br />

Rugge, Massimo 1889<br />

Ruhaak, L. Renee 606<br />

Rui, Zhang 1812<br />

Ruiz, Alex 595<br />

Ruiz, Alicia 1674<br />

Ruiz, Isaac 1123, 1269<br />

Ruiz de Galarreta, Marina 1382<br />

Ruiz-Cabello, Francisco 583<br />

Ruiz-Gaspa, Silvia 615<br />

Ruiz-Margáin, Astrid 828, 1471, 1508<br />

Rule, Jody A. 1761, 1776<br />

Runge, Jurgen H. 2152<br />

Runkana, Ashok 668<br />

Ruocco, Giancarlo 165<br />

Ruping, Kathleen 1181<br />

Rupp, Christian 618<br />

Rupp, Daniel 992<br />

Rupp, Loralee B. 15, 525, 1578, 1789,<br />

1799, 1800<br />

Ruscica, Massimiliano 2187<br />

Rusie, Erica 1141<br />

Russ, Kirk B. 1220, 1253<br />

Russell, Stephen J. 2<br />

Russell, Stuart D. 1255<br />

Russo, Aniello 431<br />

Russo, Francesco P. 1270, 1802, 1844,<br />

1889<br />

Russo, Mark W. 17, 597, 1023<br />

Rustgi, Vinod K. 1478, 1741<br />

Ryan, Karen K. 916<br />

Ryan, Marno C. 442, 483<br />

Ryan, Michael 219<br />

Ryan, Robert 1163, 1179<br />

Rychlicki, Chiara 21<br />

Ryoo, Minjung 594<br />

Ryska, Miroslav 721<br />

Ryu, Ho Sang 391, 433, 1504<br />

Sa-Cunha, Antonio 487<br />

Saab, Sammy 201, 1079, 1087, 1139,<br />

1143, 1144, 1152, 1462, 1512,<br />

1538, 1745<br />

Saad, Jennifer 1329, 2073<br />

Saad Hashem, Mohammed 1163<br />

Saag, Michael 210, 1804<br />

Saba, Francesca 2209<br />

Saberi, Behnam 1350<br />

Sabino, Ester C. 2221<br />

Saborowski, Anna 2128<br />

Sacco, Rodolfo 355<br />

Saccomanno, Stefania 21<br />

Sacerdoti, David 767<br />

Sachdeva, Sanjeev 1473<br />

Sack, Ulrich 261<br />

Sada, Yvonne 357<br />

Sadagoparamanujam, V.M. 2112<br />

Sadana, Prabodh 1338<br />

Sadikova, Ekaterina 1788<br />

Sadiq, Javaid 1234<br />

Sadiq, Omar 420<br />

Sadowsky, H. Steven 1235<br />

Sadozai, Hassan 1012, 1297<br />

Saeb-Parsy, Kourosh 258<br />

Saeed, Ali 612<br />

Saeki, Akira 408, 783, 1600<br />

Sætrom, Pål 118<br />

Safadi, Rifaat 660, 1532, 1878<br />

Safaeinili, Niloufar 563, 849<br />

Safer, Ricky 630<br />

Safran, Michal 1685<br />

Safwan, Mohamed 867, 1698<br />

Sagnelli, Caterina 1543<br />

Sagnelli, Evangelista 431, 1543, 1837<br />

Saha, Banishree 109, 496, 1016,<br />

1318, 2127<br />

Sahin, Hacer 963<br />

Sahney, Amrish 103, 146, 147, 262,<br />

2070, 2092<br />

Sahota, Amandeep K. 1119, 1135,<br />

1171, 1569, 1619<br />

Sahraoui, Salah 1621<br />

Sahu, Smiti S. 436<br />

Saibara, Toshiji 2156<br />

Saich, Andrew 537<br />

Saigal, Sanjiv 2245<br />

Saikia, Paramananda 2121, 2132<br />

Saini, Rajnish 260, 753<br />

Saito, Hidetsugu 181, 569, 2131,<br />

2140, 2234<br />

Saito, Keigo 509<br />

Saito, Masaki 1975<br />

Saito, Yoshinobu 97, 231, 453, 669,<br />

1082, 1357, 1661, 1985, 2130, 2136<br />

Saitoh, Satoshi 1062, 1191, 2020,<br />

2241<br />

Saitoh, Shinji 135<br />

Saitta, Carlo 288, 1476, 1654, 1898<br />

Saji, Yukiko 1201<br />

Sajwani, Karim 1128, 2265<br />

Sakae, Haruka 799, 937, 1838<br />

Sakai, Akira 551<br />

Sakai, Hiroshi 869<br />

Sakai, Takao 1390<br />

Sakai, Yoshio 325, 947, 1348, 1395,<br />

1653, 1968<br />

Sakai, Yoshiyuki 466, 770, 781, 1533,<br />

1669<br />

Sakaida, Isao 329, 760, 1447, 1464,<br />

1492<br />

Sakakibara, Yuko 1584<br />

Sakamori, Ryotaro 26, 97, 231, 453,<br />

669, 1082, 1197, 1357, 1635, 1636,<br />

1661, 1809, 1982, 1985, 2130, 2136<br />

Sakamoto, Minoru 1565, 1673<br />

Sakamoto, Naoya 287, 1632, 1672,<br />

1946<br />

Sakane, Sadatsugu 97, 1985<br />

Sakaue, Takahiko 688<br />

Sakisaka, Shotaro 385, 410<br />

Sako, Katsumi 456<br />

Sakuma, Tetsushi 1627<br />

Sala, Margarita 378<br />

Salama, Khaled M. 2215<br />

Salamat, Amjad 1872<br />

Salas, Eduardo 35<br />

Salazar-Gonzalez, Rosa-Maria 916<br />

Salehitezangi, Nazanin 563, 849<br />

Salerno, Debora 165<br />

Salgia, Reena 9<br />

Salhab, Ahmad 660, 1532<br />

Saliba, Faouzi 145, 1237, 1269<br />

Salizzoni, Mauro 1<br />

Sällberg, Matti 992<br />

Salloum, Shadi 25, 1015<br />

Salmeron, Javier 2171<br />

Salmon, Isabelle 178<br />

Salotti, Joanie 913<br />

Salustro, Claudia 309<br />

Salvadori, Ricardo 1495<br />

Salvatore, Lucia 432<br />

Salyana, Muhammad Atif 1282<br />

Salzer, Hannah 618<br />

Samaan, Maggie H. 840<br />

Samadi Kochaksaraei, Golasa 1576<br />

Samala, Niharika 2144<br />

Sambrotta, Melissa 1721<br />

Samet, Jeffrey 1488<br />

Samir, Waleed 1076<br />

Samonakis, Dimitrios N. 2094<br />

Samp, Jennifer C. 1089<br />

Sampaziotis, Fotios 258<br />

Samsonov, Mikhail 721<br />

Samstein, Benjamin 369, 838, 1454<br />

Samuel, Didier 421, 449, 487, 1045,<br />

1049, 1084, 1130, 1237, 1269,<br />

1782, 2082<br />

Samyn, Marianne 1711<br />

San Juan, Fernando 1251<br />

Sanabria, Judith 595<br />

Sanai, Faisal M. 1179, 1607<br />

Sanaka, Sirish 1122<br />

Sanchez, Ana M. 652<br />

Sanchez, Concepcion 1281<br />

Sanchez, Jorge 2166<br />

Sanchez, William 50, 53<br />

Sanchez, Yuri 1042, 1068, 1069,<br />

1087, 1143, 1144, 1462<br />

Sánchez, Araceli G. 1028<br />

Sanchez Ciceron, Adriana 595<br />

Sánchez-Pacheco, Aurora 1028<br />

Sanchez-Perez, Maria Jose 1341, 1342<br />

Sancho-Bru, Pau 1327, 2148<br />

Sander, Beate 350<br />

Sandlers, Yana 330<br />

Sandoval, Daniela 2210<br />

Sandoval, Darleen A. 916<br />

Sandrasegaran, Kumar E. 2150<br />

Sanford, Ukina 78<br />

Sangiovanni, Angelo 229, 393<br />

Sangro, Bruno 344<br />

Sanguankeo, Anawin 964<br />

Sanpajit, Theeranun 1631, 1662<br />

Sansom, Owen J. 678<br />

Santana, Rubia A. 1855<br />

Santangello, Diana 536<br />

Santantonio, Teresa A. 1802, 1844,<br />

2003<br />

Santolaria-Fernandez, Francisco 1342<br />

Santopaolo, Francesco 203<br />

Santoro, Michael 1467, 1478<br />

Santos, Jorge M. 1921<br />

Santos, Jose Luis 2210<br />

Santos, Miriam 473<br />

Santrampurwala, Nishreen 818, 975


1352A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Sanyal, Arun J. 52, 105, 107, 157,<br />

160, 162, 175, 239, 278, 737, 739,<br />

807, 905, 1309, 1316, 1332, 1428,<br />

1472, 1521, 2145, 2147, 2153,<br />

2154, 2163, 2208, 2220, 2235,<br />

2239, 2249<br />

Sanz, Ignacio 77<br />

Sapir, Tamar 1141<br />

Sapisochin, Gonzalo 53<br />

Saponaro, Chiara 2209<br />

Saracino, Giovanna 1210<br />

Saraf, Neeraj 2245<br />

Sarangi, Aditya N. 1475<br />

Sarangi, Vivekananda 1965<br />

Sarchielli, Erica 882, 902<br />

Sari, Sinan 1719<br />

Sarin, Shiv K. 83, 103, 146, 147, 262,<br />

327, 581, 690, 694, 695, 745, 954,<br />

1005, 1268, 1328, 1340, 1344,<br />

1483, 1659, 1663, 1762, 2070,<br />

2078, 2088, 2092, 2203<br />

Sarkar, Avik 750<br />

Sarkar, Monika 4, 1444, 2252<br />

Sarmati, Loredana 222<br />

Sarrazin, Christoph 76, 980, 1055,<br />

1078, 1179, 1559, 1797, 1861<br />

Sarrecchia, Cesare 222<br />

Sartor, Ryan B. 1521<br />

Saruwatari, Junji 1912<br />

Sasadeusz, Joe 1083, 1161<br />

Sasaki, Motoko 347, 614, 1901<br />

Sasaki, Reina 1207, 1628<br />

Sasaki, Ryo 760, 1464, 1492<br />

Sasaki, Soichiro 2119<br />

Sasaki, Yutaka 513, 1912, 1987<br />

Sasazuki, Takehiko 1632<br />

Sass, David A. 1276, 1452<br />

Sastre Turrión, Beatriz 1287<br />

Sastry, Jagannadha 1030<br />

Satake, Shintaro 1369<br />

Satapathy, Sanjaya K. 200, 1212,<br />

1233<br />

Satishchandran, Abhishek 496, 512,<br />

1318, 1330, 1958, 2127<br />

Sato, Fumiyuki 287<br />

Sato, Ken 477<br />

Sato, Masaya 461<br />

Sato, Mitsuaki 1565<br />

Sato, Shuichi 2199<br />

Sato, Shunsuke 1851<br />

Sato, Takahiro 1116<br />

Sato, Toshifumi 1771<br />

Sato, Tsutomu 1984<br />

Sato, Yasunori 347, 614, 1901<br />

Sato, Yasushi 1984<br />

Sato, Matsubara, Misako 517<br />

Satoh, Hiroaki 551<br />

Satoh, Takeaki 1159<br />

Satoskar, Rohit 1101<br />

Sauer, Peter 618<br />

Sauer, Vanessa 684, 2103<br />

Sauerbruch, Tilman 80<br />

Saunders, Maria 523<br />

Savard, Christopher 923<br />

Saviano, Antonio 432<br />

Savvidou, Savvoula 2012<br />

Sawada, Koji 927<br />

Sawadpanich, Kookwan 416<br />

Sawara, Kei 1528<br />

Sawhney, Rohit 1482<br />

Sawinski, Deirdre 871<br />

Sawle, Ashley 225<br />

Sax, Paul 1034<br />

Saxena, Akriti P. 1592<br />

Saxena, Neeraj K. 1911<br />

Saxena, Priyanka 1483, 2078<br />

Saxena, Romil 436<br />

Saxena, Utsav H. 510, 2116<br />

Saxena, Varun 471, 1825<br />

Saxinger, Lynora 641<br />

Sayiner, Mehmet 153, 892, 1186,<br />

2173<br />

Scaglione, Steven J. 532, 559<br />

Scalone, Luciana 571<br />

Scanlon, Paul D. 1470<br />

Scantlebury, Petra 1455<br />

Scarpi, Emanuela 435<br />

Scartozzi, Mario 435<br />

Scatton, Olivier 417<br />

Scetbun, Jérémy 887<br />

Schaaf, Sebastian 43<br />

Schaap, Frank G. 604<br />

Schaapman, Jelte 2069<br />

Schaefer, Esperance A. 991, 1015,<br />

1630, 1644<br />

Schaeffer, David 1413<br />

Schaffalitzky de Muckadell, Ove B. 780<br />

Schaffer, Randolph L. 861, 872<br />

Schafmayer, Clemens 617<br />

Scharnagl, Hubert 1726<br />

Scharpf, Danielle T. 1260<br />

Scharschmidt, Bruce F. 1467, 1478,<br />

1485<br />

Schattenberg, Jörn M. 1949, 2195<br />

Schaub, Johanna R. 680<br />

Scheenstra, Rene 1728<br />

Scheimann, Ann O. 280, 294<br />

Scheiner, Bernhard 270, 561, 1204<br />

Schemmer, Peter 1227<br />

Schewe, Knud 1060, 1081, 1156<br />

Schiano, Thomas D. 200, 308, 704,<br />

870, 988, 1095, 1185, 1228, 1250,<br />

1289, 1337, 1742, 2170<br />

Schiavon, Leonardo L. 2075<br />

Schiefke, Ingolf 739<br />

Schierwagen, Robert 80, 961<br />

Schiff, Eugene R. 706, 785, 1017<br />

Schild, Hans-Joerg 1353<br />

Schillie, Sarah F. 1546<br />

Schilsky, Michael L. 211<br />

Schinazi, Raymond F. 1544<br />

Schinkel, Janke 208<br />

Schinoni, M I. 595<br />

Schiöth, Helgi B. 519<br />

Schippers, Angela 963<br />

Schlaak, Joerg F. 672<br />

Schlag, Michael 1179<br />

Schlansky, Barry 10, 69, 172<br />

Schlattjan, Martin 1779<br />

Schlegel, Andrea 499<br />

Schleicher, Michael 1886<br />

Schleser, Franziska 808<br />

Schlicht, Erik M. 607, 610<br />

Schluckebier, Dominique 1735<br />

Schmeltzer, Paul A. 17, 1023<br />

Schmelzle, Moritz 276, 1222, 1394<br />

Schmerboeck, Bianca 2068<br />

Schmid, Patrick 1818<br />

Schmid, Peter 31<br />

Schmidt, Hartmut H. 684, 2103<br />

Schmidt, Mark A. 15, 525, 1578,<br />

1789, 1800<br />

Schmidt, Monica 14, 533, 2093<br />

Schmidt, Robin H. 2142<br />

Schmidt, Warren N. 1100<br />

Schmidt, Wolfgang E. 2206<br />

Schmitt, Carolyn 1184<br />

Schmitt, Timothy 352, 565, 1258,<br />

1814, 1815<br />

Schmotzer, Amy R. 1244<br />

Schmutz, Guenther 1060, 1081, 1156<br />

Schnabl, Bernd 59, 111, 1319<br />

Schnee, Matthieu 264, 1072<br />

Schneider, Kai M. 100<br />

Schneier, Amanda 1783, 2170<br />

Schnell, Gretja 705<br />

Schnickel, Gabriel 1274<br />

Schnitzler, Paul 1803<br />

Schober, Andreas 1861<br />

Schoenbachler, Ben 1574, 1787<br />

Schölzel, Katrin 1373<br />

Schork, Nicholas 1431<br />

Schott, Eckart 37, 252<br />

Schott, Micah B. 1310<br />

Schramm, Christoph 76<br />

Schreder, Alina 2128<br />

Schreiber, Jonas 2079<br />

Schreiter, Thomas 1779<br />

Schroeder, Barbara 1310<br />

Schrum, Laura W. 1389<br />

Schrumpf, Elisabeth 258, 819, 833<br />

Schuch, Anita 1666<br />

Schuck, Peter 1646<br />

Schuetze, Marcel 1060, 1081<br />

Schulze, Ryan J. 1310<br />

Schumacher, Justin D. 919, 934<br />

Schuppan, Detlef 43, 119, 183, 950,<br />

955, 1353, 1368, 1376, 1379, 1385,<br />

1432, 1434, 1437<br />

Schuschke, Dale 926<br />

Schuster, Catherine 1019<br />

Schütt, jutta 2128<br />

Schwab, Robert 43<br />

Schwabe, Christian 713<br />

Schwabe, Robert F. 98, 1957<br />

Schwabl, Philipp 149, 270, 475, 561,<br />

732, 755, 771, 1204, 1525<br />

Schwartz, Myron E. 253, 254, 380,<br />

457, 870, 1397, 1742<br />

Schwartz, Robert E. 229, 679, 1680<br />

Schwartz Sagi, Liat 1920, 1929<br />

Schwarz, Kathleen B. 133, 1555, 1694,<br />

1727, 1729<br />

Schwarzer, Remy 149, 475, 732, 755,<br />

771, 1466<br />

Schwasinger, Emma 1718<br />

Schweitzer, Nora 444<br />

Schwimmer, Jeffrey B. 159<br />

Scian, Romina 2217<br />

Sciarrone, Salvatore Stefano 1889<br />

Scioscia, Rosetta 309, 1153<br />

Sclair, Seth N. 785<br />

Scoazec, Giovanna 1123<br />

Scopigno, Tullio 958<br />

Scott, Gregory 1839<br />

Scott, Melanie 656<br />

Scuteri, Alessandra 1888<br />

Seaberg, Eric C. 1444<br />

Seal, John 1242<br />

Seay, Toni 1220


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1353A<br />

Sebagh, Mylene 1782<br />

Secchi, Maria Francesca 299<br />

Secci, Luca 309<br />

Seebauer, Caroline T. 59<br />

Seegers, Barbara 1078<br />

Seeley, Randy J. 916<br />

Seetharam, Anil B. 1043, 1109, 1145,<br />

1224, 1241, 1262<br />

Segato, E. 268<br />

Segev, Dorry L. 852, 1727, 1729,<br />

1749<br />

Segovia-Miranda, Fabian 617<br />

Seidel, Raphael A. 808<br />

Seike, Masataka 2031, 2267<br />

Seipt, Clara C. 1571<br />

Seki, Akihiro 325, 947<br />

Seki, Ekihiro 1327, 1345<br />

Seki, Michiharu 382<br />

Seki, Shuhji 324<br />

Sekimoto, Tadashi 1490, 1516<br />

Seko, Yuya 886, 888, 2202, 2230<br />

Selby, Warwick 1243<br />

Sellge, Gernot 100<br />

Selvapatt, Nowlan 2047<br />

Selwyn Samraj, Felcy Pavithra 655<br />

Selzner, Markus 11<br />

Selzner, Nazia 11, 204, 1012, 1297<br />

Semela, David 221, 1818, 1884<br />

Semple, Scott I. 260<br />

Sen, Bijoya 327<br />

Sendi, Hossein 1023<br />

Sendino, Oriol 2084<br />

Sengupta, Tanya P. 36<br />

Senju, Takeshi 1157<br />

Senkerikova, Renata 787<br />

Sennett, Karen 1455<br />

Senzolo, Marco 268, 351, 1270, 1889<br />

Seo, Hyung-Il 1025<br />

Seo, Jin Seog 424<br />

Seo, Satoru 402, 1213<br />

Seo, Wonhyo 1358, 1414<br />

Seo, Yeon Seok 391, 405, 433, 736,<br />

1438, 1504, 1605<br />

Sepp-Lorenzino, Laura 36<br />

Serejo, Fatima 150, 1876<br />

Serfaty, Lawrence 105, 162, 251, 701,<br />

777, 1676, 2145<br />

Serilmez, Murat 478<br />

Serji, Badr 1782<br />

Seror, Olivier 421, 449<br />

Serper, Marina 122, 1608<br />

Serra, Giancarlo 309, 1153<br />

Serrano, Jose 225, 1930<br />

Sersté, Thomas 1842<br />

Sese, Pilar 603, 620<br />

Setchell, Kenneth D. 916, 1695<br />

Seth, Punit 507<br />

Seth, Ratanesh K. 962, 2133, 2137<br />

Seto, Wai-Kay 126, 1545, 1550<br />

Setoyama, Hiroko 513, 1912, 1987<br />

Setshedi, Mashiko 1667<br />

Setsu, Toru 310<br />

Setze, Carolyn 1084<br />

Severin, Thomas 260, 753<br />

Sewell, Justin L. 306<br />

Seyedkazemi, Star 1041<br />

Sezaki, Hitomi 1062, 1191, 2020,<br />

2241<br />

Sha, Xiaoying 1541<br />

Shaaban, Mohamed 1271<br />

Shackel, Nicholas A. 1243, 1264<br />

Shafi, Muhammad I. 1625<br />

Shafizadeh, Tracy 1359<br />

Shah, Anee 2116<br />

Shah, Apurva 752<br />

Shah, Hemant 348, 542, 628, 1556<br />

Shah, Malay B. 854<br />

Shah, Naina 1482<br />

Shah, Neeral L. 761<br />

Shah, Niraj J. 1037, 1074, 1075<br />

Shah, Omer 1512<br />

Shah, Puja M. 1501<br />

Shah, Rohan R. 330<br />

Shah, Samir R. 578<br />

Shah, Shawn 580, 1329, 2073<br />

Shah, Shimul 873<br />

Shah, Sohela 78<br />

Shah, Vijay 1315, 1316, 1332, 1355,<br />

1406, 1947<br />

Shahinian, Vahakn 395<br />

Shahoumian, Troy 93<br />

Shaikh, Haroon A. 521<br />

Shaikh, Obaid S. 1041, 1753<br />

Shajari, Shiva 1398, 1416<br />

Shakado, Satoshi 385, 410<br />

Shaked, Abraham 5, 853, 868<br />

Shaker, Mohamed E. 1209<br />

Shaleh, Hassan M. 1966<br />

Shalev, Aryeh 2224<br />

Shan, Hong 1755<br />

Shang, Jia 1465, 2023<br />

Shanker, Mihir 784<br />

Shanmugam, Naresh P. 867<br />

Shanmukhappa, Shiva K. 823, 1695<br />

Shannon, Adam 842<br />

Shao, Benjamin 1469<br />

Shao, Jian-ying 1647<br />

Shao, Qing 1536, 1537, 2008<br />

Shao, Tuo 1305<br />

Shapiro, David 75, 155, 317, 546,<br />

562, 609, 625, 628, 644, 645, 1457,<br />

1458, 1915, 2153, 2154<br />

Sharif, Khalid 1234<br />

Sharma, Anuj 1212<br />

Sharma, Barjesh C. 1473<br />

Sharma, Dinesh 489, 490<br />

Sharma, Navneet 259<br />

Sharma, Shvetank 327, 581, 1344<br />

Sharma, Suraj 174, 348<br />

Sharma, Vivek 1753<br />

Sharma, Yogeshwar 515, 1764, 1917<br />

Sharpe, Matthew R. 585<br />

Sharpton, Suzanne R. 852<br />

Sharr, William 70, 1746<br />

Shasthry, Saggere M. 581<br />

Shasthry, Varsha 1328, 1340, 2203<br />

Shatzel, Joseph 580, 1329, 2073<br />

Shaughnessy, Melissa 210, 729<br />

Shaw, David 1083<br />

Shaw, Kenneth 993<br />

Shaw, Richard E. 2104<br />

Shaw, Stanley 2194<br />

Shawler, Todd 2123<br />

Shay, Jerry W. 682<br />

Shayakhmetov, Dmitry 12<br />

Shearn, Colin T. 921<br />

Sheen, I-Shyan 1828<br />

Sheeron, Shawn 609<br />

Sheiko, Melissa A. 1706, 1708, 1709<br />

Sheinbaum, Alan J. 488<br />

Shen, Gong 2265<br />

Shen, Jianliang 919<br />

Shen, Ming 12<br />

Shen, Xi-Zhong 1964<br />

Sheng, Jifang 2048<br />

Sheng, Lili 508<br />

Shepherd, Thomas C. 1445<br />

Sheppard, Dean 1368<br />

Sheps, Jonathan 824<br />

Sherker, Averell H. 133, 213, 1694,<br />

1922<br />

Sherman, David I. 791, 800<br />

Sherman, Morris 174, 348, 542, 851<br />

Shetty, Kirti 2118<br />

Shetty, Shishir 1296<br />

Shi, Aichao 1660<br />

Shi, Hang 933<br />

Shi, Hong 131<br />

Shi, Jiaxiao 1788<br />

Shi, Tiffany 823<br />

Shi, Xiang-de 498, 1951<br />

Shi, Ying 1562<br />

Shiba, Shunsuke 181, 569, 2131,<br />

2140, 2234<br />

Shibata, Hidetaka 1671<br />

Shibata, Noreene 2109<br />

Shibata, Toshiya 402<br />

Shibatou, Toshihiko 456<br />

Shibolet, Oren 40, 331<br />

Shiboski, Stephen 1793<br />

Shiffman, Mitchell L. 94, 106, 205, 609,<br />

616, 627, 737, 1428, 1813<br />

Shigefuku, Ryuta 1446<br />

Shigekawa, Minoru 26, 231, 453,<br />

1636, 1809, 1982<br />

Shigematsu, Hirohisa 2031<br />

Shih, Chao-Jen 2040<br />

Shih, Mason 1920, 1929<br />

Shiha, Gamal E. 708, 1076<br />

Shiina, Shuichiro 461<br />

Shim, Ju Hyun 446, 448<br />

Shima, Toshihide 2199<br />

Shimada, Mitsuo 1983<br />

Shimada, Noritomo 1097<br />

Shimada, Shingo 866<br />

Shimakami, Tetsuro 495, 990, 994,<br />

1000, 1010, 1348, 1653<br />

Shimazaki, Tomoe 1672<br />

Shimizu, Akira 869<br />

Shimizu, Masahito 657<br />

Shimizu, Toshiaki 1696<br />

Shimoda, Michiko 606<br />

Shimoda, Shinji 640, 2031<br />

Shimomura, Mayuka 1642<br />

Shimomura, Yasuyuki 2236<br />

Shimono, Yoshihiro 466, 770, 781,<br />

1533, 1669<br />

Shimosegawa, Tooru 1026, 1648,<br />

1943, 2167<br />

Shin, Eui-Cheol 1025, 1650<br />

Shin, Ji-Hyun 2118<br />

Shin, Mi-Kyung 1422<br />

Shin, Seung Kak 1896<br />

Shinkai, Kazuma 1584<br />

Shinkai, Masato 1690<br />

Shinkai, Noboru 1006, 1573, 1599,<br />

1954<br />

Shinoda, Masahiro 336<br />

Shinohara, Fumi 472, 1893<br />

Shiode, Yuto 669, 1082


1354A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Shiota, Goshi 657<br />

Shiozawa, Hirokazu 2177<br />

Shirabe, Ken 1983<br />

Shirafkan, Ali 1991<br />

Shiraishi, Koichi 314, 383<br />

Shirasaki, Takayoshi 495, 990, 994,<br />

1000, 1010, 1014, 1377, 1395, 1653<br />

Shirasawa, Hiroshi 1628<br />

Shiratsuki, Shogo 760, 1464, 1492<br />

Shiri-Sverdlov, Ronit 922<br />

Shirota, Tomoki 869<br />

Shivakumar, Pranavkumar 676, 1698<br />

Shlomai, Amir 875<br />

Shneider, Benjamin 133, 1694<br />

Sho, Takuya 287<br />

Shoji, Hirotaka 476, 1643<br />

Shomura, Masako 314, 383<br />

Shoreibah, Mohamed G. 735, 1174<br />

Shorvon, Philip J. 791, 800<br />

Shoukry, Naglaa H. 1380<br />

Showalter, Victor C. 653<br />

Shrestha, Roshan 200<br />

Shringarpure, Reshma 239, 2153, 2154<br />

Shrivastav, Manoj V. 859<br />

Shrivastava, Shikha 1013, 1029<br />

Shu, Sally 238<br />

Shubham, Smriti 690, 694, 1483<br />

Shui, Xue 510, 2116<br />

Shukla, Vivek 2118<br />

Shulga Morskaya, Svetlana 36<br />

Shulman, Nancy 722, 1039, 1051,<br />

1065, 1084, 1107<br />

Shuster, Diana L. 1088<br />

Si, Won Keun 389<br />

Si Ahmed, Si Nafa 264<br />

Sia, Daniela 253, 254<br />

Siakavellas, Spyros I. 2012<br />

Sibulesky, Lena 1221<br />

Siciliano, Massimo 432<br />

Siddiq, Masood 1872<br />

Siddiqi, Arif 1872<br />

Siddique, Asma 248, 250, 1065<br />

Siddiquee, Zakir 2116<br />

Siddiqui, Aleem 968<br />

Siddiqui, Hamda 954<br />

Siddiqui, Mohammad S. 52, 905,<br />

1309, 1472, 2163, 2220, 2235, 2249<br />

Siddiqui, Shaheryar 1513<br />

Sidharthan, Sreetha 220<br />

Siebert, Uwe 1089<br />

Sieghart, Wolfgang 270, 475, 732<br />

Sievers, Tyson 2006<br />

Sievert, William 945, 996<br />

Sigel, Keith M. 1870<br />

Signoriello, Giuseppe 1543<br />

Signorovitch, James 711<br />

Sikaroodi, Masoumeh 905, 1463, 1527<br />

Sileanu, Florentina E. 275<br />

Siler, Scott Q. 224, 228<br />

Silk, Rachel 220<br />

Sill, Bruce E. 1456, 1461<br />

Silva, Ismael D. 240<br />

Silva, Karla 78<br />

Silva, Lilian M. 84, 2189<br />

Silva, Luciana C. 2200<br />

Silva, Marco 1540<br />

Silva, Raquel S. 473<br />

Silva, Renato F. 469<br />

Silva, Telma E. 2075<br />

Silvain, Christine 3, 206, 1072, 1187,<br />

2082<br />

Silveira, Marina G. 1154<br />

Silvera, Ana 1281<br />

Silverman, Earl 1713<br />

Silverman, Gary A. 193, 197<br />

Simão, Adélia 473<br />

Simão, André L. 176, 832, 897, 960<br />

Simerabet, Hayat 816<br />

Simioni, Paolo 351, 767<br />

Simmons, Julia 823, 1695<br />

Simon, Christian O. 1597<br />

Simon, Karl-Georg 1060, 1078, 1081,<br />

1156<br />

Simon, Krzysztof 2001<br />

Simon, Miguel A. 2032<br />

Simon-Talero, Macarena 1509<br />

Simons-Petrusa, Brenna 1790, 1807<br />

Simpson, Mary Ann 71, 412, 843, 847<br />

Simpson, Pippa 1914<br />

Sims, Zayani 92<br />

Sin, Sui-Ling 1746<br />

Sinakos, Emmanouil 1126<br />

Sinclair, Marie 842, 1264<br />

Sindhi, Rakesh 763<br />

Sinegre, Thomas 1505<br />

Singal, Amit G. 371, 418, 548<br />

Singal, Ashwani 1220, 1249, 1250,<br />

1253, 1257, 2111<br />

Singal, Ashwani K. 281, 735, 1316,<br />

2100, 2114<br />

Singanayagam, Arjuna 740<br />

Singh, Ankur 1475<br />

Singh, Avishek K. 1659<br />

Singh, Baljinder 259<br />

Singh, Dharmvir 330, 518, 651, 667,<br />

668, 671<br />

Singh, Gurdeep 1817<br />

Singh, Harsimran D. 1626<br />

Singh, Shyam 695<br />

Singh, Tavankit 2182<br />

Singh, Virendra 259<br />

Singhal, Pooja 286<br />

Singhvi, Ajay 1235<br />

Singla, Manish B. 285, 1866<br />

Sinha, Rekha 1040<br />

Sinn, Dong Hyun 422, 423, 485, 1591,<br />

1960, 2049<br />

Sinnige, Marjan J. 1633, 2046, 2263<br />

Siow, Deanna L. 587, 2142<br />

Sirica, Alphonse E. 1993<br />

Sirlin, Claude B. 45, 913, 1431, 2150,<br />

2201<br />

Sitnik, Roberta 1855<br />

Sjogren, Maria 285, 1866<br />

Skaro, Anton I. 71, 563, 843, 847, 849<br />

Skibba, Kathryn 563, 849<br />

Skinner, Brianna 2126<br />

Skinner, Narelle A. 996<br />

Sklan, Ella H. 331, 672<br />

Skoien, Richard 1161<br />

Skorupski, Sharon 1254<br />

Slagle, Jason M. 529<br />

Smaïl, Allaoua 554<br />

Smart, Laura 1260<br />

Smedile, Antonina 1740, 2209<br />

Smets, Françoise 1734<br />

Smith, Abigail 838<br />

Smith, Alexander 557<br />

Smith, Bryce 1792<br />

Smith, Coleman I. 1101<br />

Smith, Davey M. 1852<br />

Smith, David B. 993<br />

Smith, Donna 1048<br />

Smith, Graham R. 1381<br />

Smith, Heidi L. 704<br />

Smith, June 530<br />

Smith, Mary A. 60<br />

Smith, Nathaniel 877, 1534<br />

Smith, Robert 11<br />

Smith, Robert E. 552<br />

Smith, Victoria 632, 1367, 1379<br />

Smits, Loek 2152<br />

Smuts, Heidi 1667<br />

Smyrk, Thomas C. 169<br />

Snead, Nicholas M. 2062<br />

Sninsky, John J. 131<br />

Snowball, Mary 125, 1798, 1807<br />

So, Samuel K. 497, 500<br />

So-Armah, Kaku 1488<br />

Soares, Filipa 258<br />

Soares e Silva, Pedro E. 2075<br />

Sobajima, Tomoaki 906<br />

Sobesky, Rodolphe 1237<br />

Sobhonslidsuk, Abhasnee 1999<br />

Sobral Dias, Margarida 150<br />

Söderberg Löfdal, Karin 116<br />

Soffredini, Roberta 2011<br />

Sogni, Philippe 373, 1194<br />

Sohail, Hifsa 1234<br />

Sohail, Muhammad 968<br />

Sohda, Tetsuro 385, 410<br />

Sohn, Joo Hyun 736<br />

Sohn, Won 485, 1415<br />

Soin, Arvinder 2245<br />

Sokal, Etienne M. 1728, 1734<br />

Sokol, Ronald J. 133, 1687, 1688,<br />

1694, 1714, 1715, 2134<br />

Sokolov, Eugene 653<br />

Solà, Elsa 266<br />

Solari, Sandra 1217<br />

Soldevila-Pico, Consuelo 217<br />

Solé, Cristina C. 266<br />

Solid, Craig 1867, 1883<br />

Soliman, Reham 1076<br />

Solis, Nancy 969<br />

Solís-Herruzo, José A. 896, 939, 1758<br />

Solis-Munoz, Pablo 896, 939, 1758<br />

Sollik, Lisa 1002<br />

Soltan, Mervat M. 1890<br />

Soltys, Kyle A. 763<br />

Somani, Vaibhav S. 752<br />

Somasundar, Ponnandai 372<br />

Sommadossi, Jean-Pierre 2266<br />

Sommer, Lisa 980, 1559<br />

Son, Byoung Kwan 951<br />

Sonderup, Mark W. 1667<br />

Song, Do Seon 1615<br />

Song, Guisheng 102<br />

Song, Jinlin 711<br />

Song, Kyoungsub 686, 1096<br />

Song, Min-Ae 1553<br />

Song, Ming 926, 1303<br />

Song, Sharon 579<br />

Song, Wei 1554<br />

Song, Won-Min 364, 2174<br />

Song, Xiaoling 2151<br />

Song, Yong 47<br />

Sonmez, Seval 2057<br />

Sonnenberg, Amnon 1791


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1355A<br />

Sonneveld, Milan J. 2001, 2013<br />

Sonoda, Yuuki 408, 783<br />

Sood, Siddharth 311, 483, 1225<br />

Sood, Vikrant 1697, 1725<br />

Soofi, Madiha E. 1843<br />

Sookoian, Silvia 2216, 2217<br />

Soonthornworasiri, Ngamphol 416, 555<br />

Sophonsritsuk, Areepan 1999<br />

Sorbo, Maria Chiara 222<br />

Sorda, Juan A. 2106<br />

Sorensen, Lisa G. 133<br />

Soriano, German 583, 596, 1509<br />

Soriano, Vincent 1180, 1881, 1908<br />

Sorin, Yushi 1191, 2020, 2241<br />

Soro-Arnaiz, Ines 930<br />

Soroka, Carol J. 23, 805, 829<br />

Soubrane, Olivier 417, 480<br />

Soulier, Alexandre 999, 1123, 1833<br />

Sousa, Pedro 661<br />

Souza, Caroline B. 2072, 2076<br />

Soza, Alejandro 1161, 1217, 2210<br />

Sozbilen, Murat 1247<br />

Sørensen, Michael 768<br />

Spaan, Michelle 1007<br />

Spadidea, Panagiota 2094, 2095<br />

Spadoni, Ilaria 893<br />

Spaggiari, Mario 850, 1271<br />

Spahr, Laurent 2082, 2174<br />

Sparkenbaugh, Erica 81<br />

Spaulding, Anne C. 1835<br />

Spearman, C. W. 1667<br />

Speiser, Jaime L. 1761<br />

Spengler, Ulrich 37, 252, 1008, 1058,<br />

1287, 2077<br />

Sperl, Jan 787<br />

Spicak, Julius 787<br />

Spiezia, Luca 351, 767<br />

Spikes, Leslie 290<br />

Spindelboeck, Walter 2068<br />

Spinella, Rosaria 1482<br />

Spinner, Nancy B. 1693<br />

Spino, Cathie 1694<br />

Spirli, Carlo 811<br />

Spivey, James 200<br />

Splith, Katrin 276<br />

Spoletini, Gabriele 1388<br />

Sponholz, Christoph 808<br />

Sporea, Ioan 802, 1443, 2004<br />

Spradling, Philip R. 15, 125, 525,<br />

1789, 1798, 1800<br />

Sprague, Andrew G. 36<br />

Sprinzl, Martin F. 376, 1559<br />

Squadrito, Giovanni 288, 1476<br />

Squires, Robert H. 1714<br />

Sripariwuth, Ekawee 1631, 1662<br />

Srishord, Alexandrea 534<br />

Srishord, Indie 153<br />

Srishord, Manirath 575<br />

Srivastava, Ankur 793, 1455, 1530<br />

Srivastava, Siddharth 1473<br />

Srivastava, Sudhir 371<br />

Sroczynski, Gaby 1089<br />

St Clair, Kristina 1175<br />

St-Laurent Thibault, Catherine 1456,<br />

1461<br />

St. Pierre, Timothy G. 798<br />

Stadlbauer, Vanessa 2068<br />

Staels, Bart 105, 162, 887, 928, 974,<br />

2145<br />

Stahlschmidt, Fabio L. 2248<br />

Stål, Per 116, 158<br />

Stambul, Beatrice 1796<br />

Stamm, Luisa M. 38, 96, 713, 718,<br />

1034, 1128, 1133, 2265<br />

Stanco, Marialuisa 148, 266, 292<br />

Standeford, Holly A. 190, 817, 834,<br />

835<br />

Standish, Richard A. 798<br />

Stanzione, Maria 1837<br />

Starace, Mario 431, 1543<br />

Stark, George R. 1025<br />

Starkel, Peter 1091<br />

Stättermayer, Albert 297, 732, 1070,<br />

1137<br />

Stauber, Rudolf E. 1092, 1885, 2068,<br />

2069<br />

Staufer, Katharina 1092<br />

Staugaard, Benjamin 792<br />

Stawicka, Agnieszka 1502<br />

Stedman, Catherine A. 38, 219, 249<br />

Stefanescu, Horia 741<br />

Stefano, José Tadeu 240, 2221<br />

Steffens, Hermann 1170, 1200<br />

Steinacher, Daniel 977<br />

Steinberg, Joel L. 1491<br />

Steindl-Munda, Petra E. 1070, 1137<br />

Steiner, Sebastian 561, 1204<br />

Steinmann, Eike 1002<br />

Stella, Jacquelline 684<br />

Stelma, Femke 208, 1633, 1995, 2263<br />

Stepanova, Maria 18, 315, 316, 534,<br />

1035, 1186, 1460, 2205<br />

Stephen, Guy R. 202<br />

Stephenne, Xavier 1734<br />

Stephens, C. 225, 583, 595, 596<br />

Stepien, Magdalena 170<br />

Sterling, Richard K. 52, 1309, 1472,<br />

1840<br />

Stern, Nicholas 344<br />

Stern, Rafael 732, 1070, 1092, 1137,<br />

1885<br />

Stern, Sarah 420<br />

Steuerwald, Nury 1023<br />

Stevens, James L. 884<br />

Stewart, Stephen 1103, 1321<br />

Stewart, Thomas 94<br />

Sticca, Antonietta 148, 292<br />

Sticova, Eva 787<br />

Stieger, Bruno 806<br />

Stiegler, Philipp 2068<br />

Stiles, Bangyan L. 1945<br />

Stine, Jonathan G. 1258, 1501<br />

Stirling, Kathryn 1298<br />

Stites, Shana D. 535<br />

Stivala, Alicia 1095, 1151<br />

Stockley, Robert 2105<br />

Stoeckle, Martin 1818<br />

Stoehr, Albrecht 1176, 1205<br />

Stoica, Teodora 1114<br />

Stojakovic, Tatjana 977, 1726<br />

Stoker, Jaap 2152<br />

Stokes, Caroline S. 2086, 2253<br />

Stokes, Scott 1839<br />

Stokkeland, Knut 116<br />

Stoll, Barbara 194, 1689<br />

Stolz, Andrew 225, 1350, 1923, 1930,<br />

1931, 1932, 1933<br />

Stolz, Donna B. 197<br />

Stone, Jack 1794<br />

Storr, Rachel 2116<br />

Strachan, Mark W. 2168<br />

Straley, Stephanie 1825<br />

Strano, Sabrina 958<br />

Strassburg, Christian P. 76, 80, 961,<br />

1008, 1287, 1927, 2077, 2108<br />

Strasser, Michael P. 1092<br />

Strasser, Simone I. 609, 1077, 1243<br />

Strassl, Robert 561<br />

Straub, Becky 1063<br />

Strautnieks, S. 1721<br />

Stravitz, R. Todd 52, 213, 1240, 1309,<br />

1472, 1761, 1763, 1768, 1914<br />

Strazzabosco, Mario 571, 743, 811,<br />

1453, 1950<br />

Streetz, Konrad L. 963<br />

Streilein, Robert 2169<br />

Streinu-cercel, Adrian 1586<br />

Stremmel, Wolfgang 618, 1411<br />

Strnad, Pavel 831, 2115<br />

Strobel, Bastian 1525<br />

Stroble, Carol 606<br />

Stroehlein, John R. 130, 682<br />

Stuart, Alan 2117<br />

Stuart, Keith E. 412<br />

Stukenborg, George R. 1501<br />

Sturdevant, Dan 1029<br />

Sturm, Ekkehard 1728<br />

Sturm, Nathalie 1439<br />

Stypinski, Daria 730, 731<br />

Su, Chien-Wei 365, 409, 439, 751,<br />

757, 759, 1639, 2050<br />

Su, Christopher 1606<br />

Su, Danmei 904<br />

Su, Feng 1211<br />

Su, Grace L. 418<br />

Su, Jie 1411<br />

Su, Lishan 1620<br />

Su, Shicheng 1423<br />

Su, Tung-Hung 1996<br />

Su, Wei-Wen 1996<br />

Su, Wen-Pang 2040<br />

Su, Xiaoping 130, 682<br />

Su, Ying-Hsiu 1554<br />

Su, Zhuang 2035<br />

Suarez, Emilio 2032<br />

Suarez, Yajaira 65<br />

Suazo, Jose 2210<br />

Subajima, Tomoaki 1642<br />

Subramaniam, Nathan 975<br />

Subramanian, Mani 241, 616, 624,<br />

632, 737, 739, 746, 777, 1428,<br />

1435, 1442, 1551, 1557, 1563,<br />

1651, 2004, 2014, 2015, 2025,<br />

2037, 2043, 2052, 2059, 2149, 2239<br />

Subramanian, Ram M. 82, 570<br />

Subramanian, Savitha 923<br />

Suchy, Frederick J. 809, 1688<br />

Suda, Goki 287, 1672, 1946<br />

Suda, Jo 2135<br />

Suda, Takahiro 26, 1809, 1982<br />

Suda, Tsuyoshi 699<br />

Suddle, Abid 646, 1113, 1744<br />

Sudo, Tomoko M. 1378<br />

Sue, Paul K. 560<br />

Suehiro, Mitsuhiko 2030<br />

Suemizu, Hiroshi 1635<br />

Suemura, Shigeki 1635, 1809, 1982<br />

Sueoka, Hideaki 1420<br />

Sugawara, Kayoko 1059, 1064, 1474,<br />

1486


1356A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Sugihara, Takaaki 1499<br />

Sugimoto, Katsutoshi 1047<br />

Sugimoto, Kazushi 1020<br />

Sugimoto, Rie 1159<br />

Sugimoto, Ryosuke 584, 1500<br />

Sugimoto, Satoru 901<br />

Sugiura, Tokio 135<br />

Sugiyama, Kazuo 569<br />

Sugiyama, Masaya 476, 1138, 1614,<br />

1632, 1643, 1658, 1673, 1795,<br />

2158, 2193<br />

Sugiyama, Nao 1003<br />

Suguness, Arvind 1831<br />

Suh, Dong Jin 446<br />

Suh, Jeong Ill 2038<br />

Suh, Kyung-Suk 400, 474, 853, 1277<br />

Suh, Sang Jun 391, 405, 1897<br />

Suh, Suk-Won 853, 1278<br />

Suhy, David 2258<br />

Suk, Fat-Moon 1996<br />

Suk, Ki Tae 736, 1515, 1957<br />

Sukeepaisarnjaroen, Wattana 1631,<br />

1662<br />

Sukriti, Sukriti 327, 1344, 2078<br />

Suksawatamnuay, Sirinporn 1631, 1662<br />

Sulfab, Maisoun 1359<br />

Sulkowski, Mark S. 91, 94, 205, 210,<br />

219, 250, 700, 702, 703, 712, 716,<br />

1035, 1039, 1057, 1749<br />

Sullivan, Danielle 1161, 1198<br />

Sultan, Khaleel H. 1147, 1188<br />

Sultan, Maya 1685<br />

Sultanik, Philippe 1194<br />

Sumazaki, Ryo 1690<br />

Sumida, Yoshio 886, 888, 901, 1980,<br />

2156, 2158, 2202, 2207, 2230<br />

Sumiyoshi, Hideaki 1369<br />

Sumpter, Colin 113<br />

Sun, Yameng 374<br />

Sun, Bing 1149<br />

Sun, Chao 2125<br />

Sun, Haibo 1149<br />

Sun, Jing 1580, 1620<br />

Sun, Mengxi 186<br />

Sun, Qian 656, 1335<br />

Sun, Runbin 919, 934<br />

Sun, Xiaochen 229, 254, 364, 1392,<br />

1940, 2174<br />

Sun, Xiaofeng 1394<br />

Sun, Ying 77, 639<br />

Sun, Ying S. 945<br />

Sunagozaka, Hajime 1348<br />

Sundaram, Shikha 137, 1709, 1715,<br />

1737, 1738<br />

Sundaram, Vinay 566, 1279, 1479,<br />

1512, 2083<br />

Sung, Pil Soo 1025<br />

Superbia, Marcel 1484, 1494, 1495<br />

Supper, Paul 1525<br />

Surana, Pallavi 779<br />

Suriguga, Suriguga 1427, 1434<br />

Susser, Simone 980, 1055, 1559<br />

Sussman, Norman L. 9, 1761<br />

Suto, Hiroyuki 944, 1682<br />

Sutton, Angela 362<br />

Sutton, Harry 134<br />

Suzuki, Ayako 540, 596<br />

Suzuki, Fumitaka 707, 1062, 1191,<br />

2020, 2241<br />

Suzuki, Harukazu 657<br />

Suzuki, Hitoshi 551<br />

Suzuki, Kazuyuki 1528<br />

Suzuki, Kunio 2213<br />

Suzuki, Michihiro 1446<br />

Suzuki, Mitsuyoshi 1696<br />

Suzuki, Shoko 1115, 1192<br />

Suzuki, Tetsuro 1393<br />

Suzuki, Yasuaki 2207<br />

Suzuki, Yoshiyuki 1062, 1191, 2020,<br />

2241<br />

Suzuki, Yuhei 477<br />

Suzuki, Yuichiro 1565<br />

Suzumura, Kazuhiro 1420<br />

Svarovskaia, Evguenia S. 91, 219, 249,<br />

713, 1168<br />

Svegliati Baroni, Gianluca 21, 355<br />

Sverdlov, Deanna 1368, 1379<br />

Svicher, Valentina 203<br />

Swain, Mark 1290<br />

Swain, Telisha 1313<br />

Swallow, Elyse 711<br />

Swaney, James 141<br />

Swanson, Herbert 313, 537, 546, 562<br />

Swanson, William M. 1117, 1132<br />

Swearingen, Christopher J. 1914<br />

Swearingen, Dennis 725<br />

Sweeney, William E. 1718<br />

Swenson, Eugene S. 702, 706, 716,<br />

726<br />

Swet, Jacob H. 653<br />

Swiderska, Aleksandra 1502<br />

Swiderska-Syn, Marzena 66, 1386,<br />

1715, 2143<br />

Swierczewska, Magdalena 115<br />

Sydor, Svenja 1779<br />

Sykes, Megan 2129<br />

Symonds, Bill 1094<br />

Symons, Julian A. 993<br />

Syn, Wing-Kin 66, 163, 1978, 2143<br />

Sypsa, Vana 2012<br />

Szabo, Gyongyi 109, 180, 496, 512,<br />

1016, 1318, 1330, 1958, 2127, 2198<br />

Szabo, Shelagh M. 1139, 1152<br />

Szanda, Gergo 1374<br />

Szenborn, Leszek 2043<br />

T. Dogan, Omer 2219<br />

Tabak, Fehmi 241, 1586, 2001, 2013,<br />

2014<br />

Taber, David J. 1272<br />

Tabernero, David 1021, 1674<br />

Tabibian, James H. 605, 643<br />

Tabrizian, Parissa 851<br />

Tachi, Yoshihiko 1834<br />

Tachlytski, Irina 1685<br />

Tack, Kenneth 2257, 2259, 2260<br />

Tacke, Frank 100, 961, 1756<br />

Tada, Seiya 1159<br />

Tada, Toshifumi 344<br />

Taddei, Tamar H. 16, 743, 1899<br />

Tadrous, Paul 791, 800<br />

Tafarel, Jean R. 2248<br />

Tagawa, Kazumi 382<br />

Taguchi, Y-h 429<br />

Tahata, Yuki 1195, 1197, 1201, 2213<br />

Tai, Andrew W. 27<br />

Tai, Chi-San 1705<br />

Tai, Ming-Hong 1942<br />

Taibi, Chiara 1751<br />

Tajima, Kazuki 336<br />

Tajiri, Kazuto 1599<br />

Tak, Won Young 363, 403, 426, 749,<br />

2025, 2028, 2038<br />

Takabatake, Hisashi 325<br />

Takabatake, Riuta 1377, 1395, 1968<br />

Takada, Hitomi 1115<br />

Takagi, Hitoshi 477<br />

Takaguchi, Koichi 1097<br />

Takahashi, Atsushi 551<br />

Takahashi, Kazuaki 1778<br />

Takahashi, Kazuhiro 1274, 2031<br />

Takahashi, Kazuto 944<br />

Takahashi, Masahiko 2177<br />

Takahashi, Takeshi 1635<br />

Takahashi, Tetsuyuki 1981<br />

Takahira, Sachiko 314<br />

Takai, Satoshi 1648<br />

Takaki, Akinobu 339, 2236<br />

Takaki, Yugo 1724<br />

Takakura, Kazuki 2074<br />

Takamatsu, Shinji 906, 1642, 2156<br />

Takamatsu, Yuki 1655<br />

Takami, Taro 329, 760, 1447, 1464,<br />

1492<br />

Takami Lageborn, Christine 116<br />

Takamura, Masaaki 310, 2165<br />

Takamura, Toshinari 144, 990, 1010,<br />

1348, 1377, 1968<br />

Takase, Masaru 820<br />

Takashi, Niizeki 455<br />

Takashima, Tomoyuki 466, 770, 781,<br />

1533, 1669, 1975<br />

Takata, Kazuhide 385, 410<br />

Takata, Ryo 466, 770, 781, 1533,<br />

1669<br />

Takatori, Hajime 699<br />

Takayama, Kazuo 1690<br />

Takayama, Tadatoshi 407<br />

Takebe, Takanori 676<br />

Takeda, Kazuyoshi 2119<br />

Takegoshi, Kai 144, 1377, 1395, 1968<br />

Takehara, Tetsuo 26, 97, 227, 231,<br />

453, 669, 906, 1082, 1195, 1197,<br />

1201, 1357, 1635, 1636, 1642,<br />

1661, 1809, 1974, 1982, 1985,<br />

2130, 2136, 2156, 2213<br />

Takei, Hajime 804, 1696<br />

Takei, Yoshiyuki 584, 1020, 1500<br />

Takeiri, Masatoshi 1918<br />

Takemoto, Kenji 1918<br />

Takemura, Shigekazu 429, 1748<br />

Takeshita, Eiji 2146<br />

Taketani, Hiroyoshi 886, 888, 901,<br />

2230<br />

Taketomi, Akinobu 476, 866, 1632<br />

Takeuchi, Takahito 1724<br />

Takeuchi, Yasuto 2236<br />

Takeyama, Yasuaki 385, 410<br />

Takigawa, Atsuo 26, 453, 1809, 1982<br />

Takikawa, Hajime 1757<br />

Takikawa, Yasuhiro 1528<br />

Takimoto, Rishu 1984<br />

Takkenberg, Bart 1633, 1995, 2005,<br />

2046<br />

Takyar, Varun K. 779, 1498<br />

Talal, Andrew H. 332, 1823, 1841<br />

Talaty, Jennifer E. 725<br />

Taleb, Eman 1118<br />

Taliani, Gloria 222, 983, 1802, 1844<br />

Tallis, Caroline 1077<br />

Talmat, Nabila 1847


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1357A<br />

Talwani, Rohit 251, 715<br />

Tam, Edward 1051<br />

Tamai, Tsutomu 799, 937, 1421, 1838<br />

Tamaki, Nobuharu 1115<br />

Tamberi, Stefano 435<br />

Tamburrino, Domenico 1388<br />

Tamè, Maria Rosa 1751<br />

Tameda, Masahiko 1020<br />

Tamhane, Ashutosh 1804<br />

Tamko-Mella, Ghislaine Flore 1833<br />

Tamori, Akihiro 429, 1125, 1748,<br />

1824, 1983<br />

Tamura, Yasuaki 1375, 1378<br />

Tan, Amy 870<br />

Tan, Bingyan 454<br />

Tan, Brandon 916<br />

Tan, Chee-Kiat 451<br />

Tan, Corey 819<br />

Tan, Damien 167<br />

Tan, Hua 993<br />

Tan, Poh Seng 364<br />

Tan, Youwen 2048<br />

Tana, Michele M. 306<br />

Tanabe, Kazutaka 1213, 1918<br />

Tanabe, Kenneth K. 957, 1938<br />

Tanabe, Minoru 406<br />

Tanahashi, Toshihito 429<br />

Tanaka, Eiji 1632, 1673<br />

Tanaka, Hajime 613<br />

Tanaka, Hiroki 927<br />

Tanaka, Hironori 1975<br />

Tanaka, Kathryn 1333<br />

Tanaka, Motohiko 513, 1912, 1987<br />

Tanaka, Sadao 456<br />

Tanaka, Saiyu 2156, 2202, 2230<br />

Tanaka, Satoshi 97, 231, 453, 669,<br />

1082, 1357, 1661, 1985, 2130, 2136<br />

Tanaka, Shingo 1984<br />

Tanaka, Shinji 406, 1980<br />

Tanaka, Shogo 1748<br />

Tanaka, Takashi 385, 410<br />

Tanaka, Takuji 144, 1377, 1395, 1968<br />

Tanaka, Yasuhito 1006, 1018, 1026,<br />

1097, 1138, 1448, 1573, 1599,<br />

1618, 1640, 1655, 1983, 2063<br />

Tanaka, Yasunobu 1433<br />

Tanaka, Yasuo 461<br />

Tanaka, Yujiro 1018<br />

Tanaka, Yuki 1159<br />

Tandoi, Francesco 1, 1740<br />

Tandon, Puneeta 82, 570, 742<br />

Tanfin, Zahra 816<br />

Tang, Alister L. 1121<br />

Tang, Hong 1562<br />

Tang, Weiliang 1611<br />

Tang, Yilang 183, 1353<br />

Tange, Kazuhiro 524<br />

Tangkijvanich, Pisit 2036, 2063<br />

Taniai, Makiko 1326<br />

Taniguchi, Eitaro 233<br />

Tanikawa, Ken 1690<br />

Tanikawa, Tomohiro 2030<br />

Tanikella, Rajasekhar 1513<br />

Taniki, Nobuhito 181, 569, 2131,<br />

2140, 2234<br />

Tanimoto, Hironori 2031<br />

Tanné, Florence 264<br />

Tanno, Federico C. 595<br />

Tanoue, Yasushi 505<br />

Tantipanichtheerakul, Supot 1631, 1662<br />

Tanwandee, Tawesak 467, 1631, 1662<br />

Tanwar, Sudeep 793, 1455, 1530<br />

Tao, Junyan 1948<br />

Tao, Shiqi 1865<br />

Tao, Sijia 1544<br />

Tapper, Elliot B. 566, 1222, 1477,<br />

2196, 2197, 2227<br />

Tariciotti, Laura 203<br />

Tarocchi, Mirko 949, 1963<br />

Tarr, Phillip 2091<br />

Tartaglione, Erica 803<br />

Tartof, Sara Y. 1135<br />

Tashiro, Shigeki 1159<br />

Tashiro, Taku 1584<br />

Tate, Janet 1488<br />

Tateishi, Ryosuke 461<br />

Tateno, Chise 1150, 1670, 1679, 2120<br />

Tateyama, Masakuni 1912, 1987<br />

Tatosian, Daniel A. 707<br />

Tatsumi, Kohei 81<br />

Tatsumi, Tomohide 26, 97, 231, 453,<br />

669, 1082, 1195, 1197, 1201, 1357,<br />

1635, 1636, 1661, 1809, 1982,<br />

1985, 2130, 2136, 2213<br />

Taura, Kojiro 402, 1213, 1918<br />

Taura, Naota 952, 1671<br />

Tawara, Seiichi 26, 453, 1635, 1809,<br />

1982<br />

Tawdrous, Monika 2068<br />

Taylor, Amy E. 1695<br />

Taylor, Avril 1794<br />

Taylor, James 718<br />

Taylor, Ryan 1291, 1761, 1814, 1815<br />

Taylor, Sandra L. 606<br />

Taylor, Simeon 141<br />

Taylor, Stephanie 1309<br />

Taylor, Thom 1048<br />

Taylor, William R. 169<br />

Tayyab, Ghias Un Nabi 1872<br />

Tchaikovskaya, Tatyana 515<br />

Te, Helen S. 271, 539, 544, 553, 579,<br />

1101, 1229<br />

Teachenor, Olga 1146, 1199<br />

Teckman, Jeffrey 121, 1694<br />

Tedgui, Alain 887<br />

Telese, Andrea 337, 1388<br />

Tellatin, S 268<br />

Teltsch, Dana Y. 1175<br />

Tempesti, Sara 949, 1963<br />

Temple, Luisa 78<br />

Temple, Sarah 357<br />

Ten Have, Gabriella A. 671, 1523<br />

ten Kate, Fiebo J. 1590<br />

Teng, Kun-Yu 493<br />

Teng, Ruei-Dun 1639<br />

Tenti, Elena 435<br />

Teo, Chong-Gee 2126<br />

Teoh, Narci C. 55, 230<br />

Teperino, Raffaele 1386<br />

Teperman, Lewis W. 200<br />

Terai, Shuji 310, 1447, 2165<br />

Teranishi, Yuga 592, 1125<br />

Terao, Naoko 906<br />

Terashima, Takeshi 370, 392, 1653<br />

Terra, Carlos 84, 2189<br />

Terrault, Norah 94, 200, 236, 239,<br />

471, 1034, 1084, 1106, 1259, 1560,<br />

1598, 1825, 2153, 2154, 2161,<br />

2220, 2252<br />

Terrin, Norma 1592<br />

Teruki, Miyake 524<br />

Teshale, Eyasu H. 15, 525, 1789, 1800<br />

Testro, Adam 1225<br />

Teta-Bissett, Monica 494, 511<br />

Teterina, Anastasia 2238, 2254<br />

Teti, Elisabetta 222<br />

Teuber, Gerlinde 1170, 1200, 1861<br />

Teuber, Peter 278<br />

Tevar, Amit D. 216<br />

Thabut, Dominique 263, 744, 1510<br />

Thacker, Leroy 82, 570, 1463<br />

Than, Tin A. 590<br />

Thanapirom, Kessarin 360, 1631, 1662<br />

Thandassery, Ragesh B. 1843<br />

Thapaliya, Samjhana 651, 667, 668,<br />

671, 1308<br />

Tharinger, Hugo 2009<br />

Tharwa, El-Sayed S. 1862, 1879<br />

Thasler, Wolfgang E. 520, 1684<br />

Theise, Neil D. 681, 685, 2214<br />

Therapondos, George 200, 1242<br />

Therneau, Terry M. 6, 397<br />

Thevenon, Sylvie 1162<br />

Thevenot, Thierry 145, 264, 554, 1036,<br />

1187, 2082<br />

Thi, Emily P. 2007, 2034, 2062<br />

Thibault, Vincent 786, 1808<br />

Thiele, Maja 773, 780, 789, 792,<br />

1440<br />

Thimme, Robert 1666, 1803<br />

Thirumurthy, Harsha 1063<br />

Thogaripally, Suneetha 1804<br />

Thoma, Christian 2183<br />

Thomas, Emmanuel 785, 1017<br />

Thomas, Howard 1551<br />

Thomas, Peter 1969<br />

Thomas, Richard C. 2226<br />

Thomes, Paul G. 1323, 1347, 1389<br />

Thomopoulos, Konstatinos 2094, 2095<br />

Thompson, Alex J. 106, 249, 442, 627,<br />

703, 996, 1077, 1551, 1557, 1563,<br />

1651, 2037, 2262<br />

Thompson, Richard J. 1721, 2105<br />

Thomsen, Karen Louise 908, 1300<br />

Thomson, Angus W. 338<br />

Thomson, Mary 576<br />

Thongsawat, Satawat 1631, 1662<br />

Thonig, Antje 1519<br />

Thorburn, Douglas 73, 76, 413, 634,<br />

1216<br />

Thorgeirsson, Snorri S. 677<br />

Thuluvath, Paul J. 82, 570, 739<br />

Thung, Swan N. 254, 506, 870, 1337<br />

Thurner, Lorenz 463<br />

Thursfield, Vicky 442<br />

Thuy, Tuong Thi Van 191<br />

Tian, Dean 328<br />

Tian, Hui 198, 1986<br />

Tian, Jijing 1285<br />

Tian, Lei 1396<br />

Tice, Ashley B. 166<br />

Tieleman, Madelon 68<br />

Tien, Phyllis 2252<br />

Tierney, Amber 1101, 1184<br />

Tiirikainen, Maarit 1553<br />

Tijeras-Raballand, Annemilaï 417<br />

Tikhanovich, Irina 1291<br />

Tikhonova, Natalia 721<br />

Tillman, Bryan 216<br />

Tillman, Holly 213, 1761, 1768


1358A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Tillmann, Hans L. 540, 1922, 1931<br />

Timm, Joerg 672<br />

Ting Zhang, Ting 1939<br />

Tiniakos, Dina 652<br />

Tinti, Francesca 983<br />

Tiribelli, Claudio 1339, 1346, 2098<br />

Tiro, Jasmin A. 548<br />

Tisdale, John F. 1498<br />

Tisone, Giuseppe 203, 1751<br />

Titos, Esther 2087<br />

Tizzard, Sarah 1704<br />

Tobita, Hiroshi 2199<br />

Toda, Nobuo 382, 2240<br />

Todus, Robin 1172<br />

Toffanin, Sara 253<br />

Togawa, Takao 135<br />

Togayachi, Akira 1448<br />

Tokoro, Masanori 2267<br />

Tokumoto, Yoshio 788, 1293, 2146<br />

Tokunaga, Katsushi 1138, 1632<br />

Tokushige, Katsutoshi 1326, 2207<br />

Toledano, Ronen 2224<br />

Tolenaars, Dagmar 604, 827<br />

Toli, Barbara 1126<br />

Tomasi, Maria Lauda 594<br />

Tomasiewicz, Krzysztof 1582<br />

Tominaga, Kentaro 310<br />

Tomita, Kyoko 895<br />

Tomiyama, Takashi 613, 639<br />

Tomkoetter, Lena 1282<br />

Tona, Francesco 268<br />

Tonascia, James 160, 2150, 2220<br />

Tonello, Silvia 775<br />

Tong, Myron 1165<br />

Tong, William 1571<br />

Tonnu-Mihara, Ivy 1110<br />

Tonon, Marta 148, 292, 767, 775<br />

Tontini, Gian Eugenio 1506<br />

Topal, Baki 171<br />

Topic, Nina 538<br />

Torbeyns, Anne 702<br />

Tordjmann, Thierry 816<br />

Toribio, Wilma 1870<br />

Toriguchi, Kan 1918<br />

Torii, Nobuyuki 1326<br />

Torimura, Takuji 233, 304, 399, 455,<br />

681, 688, 918, 1819, 1990, 2158<br />

Tornai, David 2090<br />

Tornai, Istvan 2090<br />

Tornai, Tamas 2090<br />

Torok, Natalie J. 1285, 2179<br />

Torre, Aldo 1471, 1508, 2223<br />

Torrens, Maria 1509<br />

Torres, Dawn M. 285, 1866, 2239<br />

Torres, Harrys A. 1030, 1160<br />

Torres, Javiera 969<br />

Torres Hernandez, Alejandro 1282<br />

Torres-Capelli, Mar 930<br />

Torres-Gonzalez, Edilson 2142<br />

Torzilli, Guido 602<br />

Toshimori, Akiko 524<br />

Toshiro, Takezaki 1421<br />

Tosti, Maria Elena 1802, 1844<br />

Tosti Guerra, Cristina 1410<br />

Tosun, Selma 1622, 1623<br />

Toutouza, Marina G. 1489<br />

Towner, William J. 205<br />

Townsend, Mary L. 1142<br />

Townshend-Bulson, Lisa J. 1790, 1798,<br />

1807<br />

Toyoda, Hidenori 344<br />

Toyota, Joji 1116, 2065<br />

Tozun, Nurdan 558, 1822<br />

Trabucchi, Michele 987<br />

Trabut, Jean-Baptiste 1102<br />

Tracy, Russel 2228<br />

Traetow, Daniel 1260<br />

Trakroo, Sushrut 1122<br />

Tran, Albert 3<br />

Tran, Chris 36<br />

Tran, Lien 269<br />

Tran, Patricia 316<br />

Tran, Tram T. 249, 1479<br />

Tran, Viet Chi 1845<br />

Tran Minh, Margherita 1239<br />

Trapero, Maria 1511<br />

Trauner, Michael 76, 149, 270, 297,<br />

475, 561, 609, 662, 732, 755, 771,<br />

977, 1070, 1092, 1137, 1204, 1466,<br />

1525, 2164<br />

Trautwein, Christian 100, 828, 831,<br />

961, 963, 1045, 1756, 1924, 2115<br />

Trawick, Bobby N. 1209<br />

Traxinger, Brianna 1692, 1699<br />

Trebicka, Jonel 80, 773, 961<br />

Treeprasertsuk, Sombat 360, 416, 555<br />

Trehanpati, Nirupma 327, 695, 1344,<br />

1659, 1663, 1762, 2088<br />

Tremblay, Mélanie 51, 1523<br />

Trembling, Paul M. 793, 1530<br />

Trenell, Michael I. 2183<br />

Trepo, Christian 1162<br />

Trépo, Eric 178, 2079<br />

Treviño, Ana 1180<br />

Trevisani, Franco 355, 356<br />

Triantos, Christos K. 2094, 2095<br />

Trinacty, Connie M. 15, 525, 1578,<br />

1789, 1799, 1800<br />

Trindade, Alexandre 897<br />

Trindade, Miguel 1876<br />

Trinh, Huy N. 123, 207, 368, 1182,<br />

1549, 1561, 2015, 2025, 2052<br />

Trinh, My My 2262<br />

Trinh, Roger 722, 1051<br />

Trinh, Thu Le 1967<br />

Tripathi, Dinesh M. 1518<br />

Tripathi, Rakesh 705<br />

Tripodi, Gianluca 1654<br />

Tripon, Simona 744, 1510<br />

Trivedi, Palak J. 73, 76, 634<br />

Trooskin, Stacey B. 1864, 1880<br />

Trujillo, Ruby 1121<br />

Trunecka, Pavel 787<br />

Trychta, Kathleen 323<br />

Tsai, Ming-Chao 1942<br />

Tsai, Naoky CS C. 1046, 1108, 1121,<br />

1553<br />

Tsai, Shannon 195<br />

Tsai, Yi-Shan 2264<br />

Tsang, Owen Tak Yin 2025<br />

Tsang, Rebecca 532, 559<br />

Tschernig, Thomas 1299, 1385<br />

Tschudy-Seney, Benjamin 685<br />

Tschuor, Christoph 665<br />

Tse, Yee-Kit 349<br />

Tseng, Kuo-Chih 1996<br />

Tseng, Po-Lin 1873, 1942<br />

Tseng, Tai-Chung 1996<br />

Tsianos, Epameinondas 801<br />

Tsianou, Zoi 801<br />

Tsien, Cynthia 667<br />

Tsipouras, Markos 801<br />

Tsiriga, Athanasia 1489<br />

Tsochatzis, Emmanuel 413, 764, 1455<br />

Tsubouchi, Eiji 2146<br />

Tsubouchi, Hirohito 937, 1421, 1757<br />

Tsuchida, Takuma 1392<br />

Tsuchiura, Takayo 1632<br />

Tsuchiya, Hiroyuki 1336<br />

Tsuchiya, Kaoru 1115<br />

Tsuds, Yasuhiro 966<br />

Tsuge, Masataka 472, 663, 986, 1056,<br />

1150, 1627, 1645, 1649, 1670,<br />

1681, 1895, 1944, 2120<br />

Tsujii, Masahiko 1197, 2213<br />

Tsukamoto, Hidekazu 112, 682, 904,<br />

1936, 2118<br />

Tsunematsu, Seiji 287, 1946<br />

Tsuneyama, Koichi 1981<br />

Tsung, Allan 184<br />

Tsunoda, Tomoyuki 164<br />

Tsuruga, Yosuke 866<br />

Tsuruya, Kota 314, 383, 697, 2177<br />

Tsutsumi, Susumu 1006, 1599<br />

Tsutsumi, Takeya 912<br />

Tsuzaki, Ryuichiro 339<br />

Tsuzura, Hironori 1851<br />

Tucker, Edwin 1920, 1929<br />

Tucker, Joseph D. 577<br />

Tujios, Shannan R. 290, 1766<br />

Tuma, Dean J. 589, 1323, 1347, 1969,<br />

2139<br />

Tumas, Daniel 1359<br />

Tummala, Swetha 201, 1745<br />

Tunnicliffe, Elizabeth M. 2190<br />

Tunõn, María Jesús 1398<br />

Tur-Kaspa, Ran 875<br />

Turan, Ilker 246, 1247, 1739<br />

Turcios, Lilia 673<br />

Turmelle, Yumirle P. 133, 1694<br />

Turner, Samuel 89, 1090, 1852<br />

Turner, Scott M. 1425<br />

Turnquist, Heth R. 184<br />

Turquet, Jeid 2105<br />

Turturro, Francesco 1160<br />

Tuzun, Ali 2033<br />

Tyndall, Mark 86, 543<br />

Tzallas, Alexandros T. 801<br />

Tzarnas, Stephanie 1864, 1880<br />

U, Rebecca Y. 177<br />

Uchida, Daisuke 2236<br />

Uchida, Sawako K. 429, 1125, 1824<br />

Uchida, Shinjiro 783, 1600<br />

Uchida, Takuro 663, 986, 1150, 1649,<br />

1670, 1681, 1858, 1895, 1944, 2120<br />

Uchida, Yoshihito 1059, 1064, 1577,<br />

1757<br />

Uchikoshi, Manabu 1302<br />

Uchino, Koji 461<br />

Uchiyama, Akira 226, 917, 1771<br />

Uddin, Muazzam 1872<br />

Udo de Haes, Joanna 208, 2263<br />

Udoh, Uduak S. 1313<br />

Udompap, Prowpanga 88, 1208, 2155<br />

Ueda, Keiji 1642<br />

Uehara, Daisuke 477<br />

Uehara, Hisanori 1981<br />

Uehara, Takahide 524<br />

Ueland, Joseph 709<br />

Uemoto, Shinji 402, 1213, 1918, 1976


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1359A<br />

Uemura, Hayato 1064<br />

Ueno, Takato 688, 918<br />

Uesaka, Katsuhiko 820<br />

Ueyama, Misuzu 794, 1795, 2158,<br />

2193<br />

Ulivi, Paola 435<br />

Ullah, Naeem 1275<br />

Ullmer, Christoph 816<br />

Um, Soon Ho 391, 433, 736, 1438,<br />

1504, 1605<br />

Umbro, Ilaria 983<br />

Umeda, Rumiko 2177<br />

Umemura, Atsushi 232, 886, 888, 901,<br />

1980<br />

Umemura, Machiko 1672<br />

Umeno, Yumi 1990<br />

Umetsu, Teruyuki 1026, 1648, 1943,<br />

2167<br />

Ungerer, Jacobus 784<br />

Ungerleider, Nathan 686<br />

Unler, Gülhan K. 2206<br />

Unser, Ariel 52, 1463, 1472, 1491,<br />

1527<br />

Upala, Sikarin 964<br />

Uprichard, Susan L. 982<br />

Urabe, Ayako 1195, 1201, 2213<br />

Urata, Noriyo 2030<br />

Urban, Thomas J. 225, 1930, 1933<br />

Urbanek, Petr 1487, 1678<br />

Urbani, Luca 337, 1388<br />

Uribe, Claudia L. 1050<br />

Uriel, Alison J. 1112<br />

Urraro, Teresa 1190<br />

Urrunaga, Nathalie H. 1911<br />

Urrutia, Raul A. 692<br />

Ursic-Bedoya, José 816<br />

Urushihara, Naoko 1378<br />

Uschner, Frank E. 80, 961<br />

Uslan, Mustafa I. 846<br />

Usui, Akihiro 1993<br />

Usui, Shingo 569<br />

Uto, Hirofumi 799, 937, 1421, 1838<br />

Utrankar, Amol 1223<br />

Utsumi, Masashi 339<br />

Utsumi, Teruo 1322, 1526<br />

Utsunomiya, Hiroki 2146<br />

Utsunomiya, Tohru 1983<br />

Uyama, Naoki 188, 1420<br />

Vadalà, Domenica 288<br />

Vaddady, Pavan 731<br />

Vadnal, Prashant 118<br />

Vaillant, Andrew 31<br />

Vaishnav, Joban 1255<br />

Valantin, Marc-Antoine 568, 1093<br />

Valdeolivas, Tatiana 2072, 2076<br />

Valderama, Adriana 16<br />

Vale, Ana 1540<br />

Vale, Luke 636, 1457<br />

Valenti, Luca 973, 2187<br />

Valentin, Nelson 282<br />

Valla, Dominique C. 772, 1288, 2148<br />

Vallet-Pichard, Anaïs 95, 1158, 1194<br />

Vallier, Ludovic 258<br />

Valveri, Vincenza 2011<br />

Van Allen, Jessica 1770<br />

van Bömmel, Alena 245<br />

van Bömmel, Florian 245, 1559, 1594,<br />

1638, 2003, 2012<br />

van Buuren, Henk R. 68, 73, 302, 601,<br />

634<br />

van Campenhout, Margo J. 1585, 1586<br />

van de Garde, Martijn D. 2122<br />

van de Graaf, Stan F. 637<br />

van de Laar, Thijs J. 1846<br />

van den Berg, Aad P. 68, 302<br />

Van Den Berg, Machteld 167<br />

van den Boom, Rivka 1238<br />

Van Der Graaff, Denise 976<br />

van der Helm, Jannie 1846<br />

van der Mark, Vincent 492<br />

van der Meer, Adriaan J. 350, 445,<br />

1166, 1590, 2255<br />

Van Der Merwe, Schalk 171<br />

van der Net, Guido 2122<br />

Van der Ploeg, Lex H. 143<br />

van der Poorten, David 441<br />

van der Ree, Meike H. 208, 2005,<br />

2255, 2263<br />

van der Reijden, Johan 2069<br />

van der Valk, Marc 37, 208, 252, 1058<br />

Van der Ven, Peter F. 80<br />

van Dijk, Remco 492<br />

Van Dooren, Gino 1163, 1179<br />

van Dort, Karel A. 1633, 2005<br />

van Erpecum, Karel J. 609<br />

Van Eygen, Veerle 39<br />

van Gaal, Luc 928<br />

van Gorp, Patrick 922<br />

van Gulik, Thomas 637<br />

van Hoek, Bart 68, 302, 1049, 1238,<br />

2069<br />

Van Hul, Noemi 678<br />

Van Itallie, Christina 20<br />

van Leeuwen, Ester M. 1633<br />

van Liempd, Sebastiaan 953<br />

Van Natta, Mark L. 107, 157, 236,<br />

2150<br />

Van Nieuwkerk, Carin M. 302, 1995<br />

van Nuenen, Adrianus C. 2255<br />

van Oord, Gertine 1007, 1586<br />

Van Remoortere, Peter 39<br />

van Scherpenzeel, Monique 2102<br />

van Seggelen, Wouter 1112, 1852<br />

van Silfhout, Joanne J. 302<br />

Van Steenbergen, Werner 171<br />

Van Steenkiste, Christophe 1091<br />

Van Vlierberghe, Hans 1049, 1067<br />

van Vliet, Andre 208, 2263<br />

Van Wagoner, David R. 518, 651, 667<br />

Van Wilgenburg, Bonnie 1298<br />

Vanatta, Jason 1212, 1233<br />

Vanderhoff, Aaron M. 1151<br />

Vandevoorde, Ann 39<br />

Vanlemmens, Claire 1187<br />

Vannelli, Gabriella Barbara 882, 902<br />

Vanslembrouck, Ragna 171<br />

Vanveuren, Charlotte 2222<br />

VanWagner, Lisa B. 563, 849, 1235<br />

Vanwolleghem, Thomas 797, 1637,<br />

2122<br />

Vaquero, Javier 1295, 1520<br />

Varaut, Anne 1123<br />

Varela, Maria 378<br />

Varela, Marta 953, 2244<br />

Vargas, Hugo E. 82, 248, 250, 570,<br />

1038, 1106, 1124<br />

Vargas, Juan 78<br />

Vargas, Victor 609, 2069<br />

Varghese, Rony 2162, 2186<br />

Varier, Raghu 1702<br />

Varma, Sharat 1728, 1734<br />

Varshney, Aditi 1659<br />

Varunok, Peter 1065<br />

Vasconcelos, Ana Clara 2237<br />

Vasilescu, Alexandra 280<br />

Vasilieva, Larisa E. 1489, 2096<br />

Vasudevan, Madavan 695, 1659<br />

Vatcheva, Kristina 1424<br />

Vatsalya, Vatsalya 1334<br />

Vavassori, Sara 393<br />

Vega, Maricruz 597, 1932<br />

Veidal, Sanne S. 978<br />

Veitsman, Ella 1032<br />

Velame, Daniela 2237<br />

Veldt, Bart J. 445, 1166<br />

Vella, Stefano 1153, 1802, 1844<br />

Vellicce, Alejandra 2106<br />

Vellozzi, Claudia 1574, 1787, 1792<br />

Vellucci, Vincent 709<br />

Velosa, José F. 150, 1876<br />

Veloso, Julio M. 1256<br />

Venditti, Mario 2067<br />

Venkat, Nandini 2116<br />

Venkatesan, Chapy 575<br />

Venkatesh, Anu 364, 2174<br />

Venter, Julie 190, 621, 817, 834, 835,<br />

929, 1317, 1349<br />

Ventura-Cots, Meritxell 380, 438, 457,<br />

1509<br />

Venugopal, Senthil K. 1664<br />

Vergani, Diego 300, 608, 1704<br />

Vergniol, Julien 450, 1439, 2181<br />

Verhaag, Esther M. 591<br />

Verheij, Joanne 637, 1995, 2046,<br />

2152<br />

Verhey, Elke 2002<br />

Verheyen, Fons 922<br />

Verkade, Henkjan J. 1728<br />

Verlinden, Wim 797<br />

Verma, Manisha 530, 535, 1932<br />

Verma, Neha 1451<br />

Verma, Suman 646, 1098, 1113,<br />

1117, 1132, 1173, 1216, 1588, 2054<br />

Verma, Vikas K. 1315, 1406<br />

Verma, Yogita 1268<br />

Vermehren, Johannes 1055, 1078,<br />

1797<br />

Verna, Elizabeth C. 200, 272<br />

Vernaz, Nathalie 1539<br />

Verne, Julia 564<br />

Vernet, Guy 1833<br />

Verrier, Eloi R. 1019<br />

Verrijken, An 928<br />

Verslype, Chris 171<br />

Verspaget, Hein W. 1238, 2069<br />

Verucchi, Gabriella 1802, 1844<br />

Vetter, Diana 1779<br />

Vianna, Rodrigo M. 1054<br />

Vibert, Eric 417, 421, 449, 487, 1237,<br />

1269<br />

Vickerman, Peter 1794, 1830<br />

Victor, David W. 200, 1215<br />

Victor, Livia B. 2072, 2076<br />

Vidaud, Michel 998, 1601<br />

Vieira, Maria Clara 2237<br />

Vierling, John M. 200, 624, 701, 706,<br />

712, 715, 1086, 1478<br />

Vieth, Michael 1506<br />

Viganò, Mauro 2011<br />

Vigil de Melo, Silvana 2075


1360A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Vignozzi, Linda 882, 902<br />

Vij, Mukul 1698<br />

Vijayan, Balasubramaniam 1104<br />

Vijayan, Santosh 183<br />

Vijaykumar, T. R. 284<br />

Vijayvergiya, Rajesh 259<br />

Vijgen, Leen 39<br />

Vila, Marta 1674<br />

Vila, Sergi 326<br />

Vilar, Graca 1859<br />

Vilarinho, Silvia 1719<br />

Vilca-Melendez, Hector 1775<br />

Vilchez, Regis A. 1086, 1107<br />

Vilchez, Valery 673, 854, 1267, 2081<br />

Villa, Erica 1802, 1844<br />

Villa, Rosa 8<br />

Villadsen, Gerda E. 768<br />

Villani, Ambra 811<br />

Villanueva, Augusto 254<br />

Villanueva, Càndid 734<br />

Villard, Jean 1735<br />

Villaret, Maxime 997<br />

Villela-Nogueira, Cristiane A. 273, 289,<br />

762<br />

Villeneuve, Jean-Pierre 1997<br />

Vilstrup, Hendrik V. 301, 768, 2071<br />

Vimalesvaran, Sunitha 1733<br />

Vinaixa, Carmen 1230, 1251<br />

Vincent, Catherine 609<br />

Vincent, Robert 160, 905<br />

Vinci, Maria 1802, 1844<br />

Vinel, Jean-Pierre 263<br />

Viner, Kendra 1810<br />

Vinikoor, Michael J. 1575<br />

Vinken, Mathieu 586<br />

Violette, Shelia 1368<br />

Vion, Anne-Clémence 887<br />

Virabhak, Suchin 1087, 1143<br />

Virdone, Roberto 355<br />

Visconti, Luca 288<br />

Vismara, Marta 1950<br />

Vispo, Eugenia 438<br />

Visvanathan, Kumar 996, 1225, 2262<br />

Viswanathan, Preeti 515, 1764, 1917<br />

Vitale, Alessandro 351, 356<br />

Vitale-Cross, Lynn 20<br />

Vitalis, Zsuzsanna 2090<br />

Vitalone, Matthew J. 341, 2123<br />

Vitek, Libor 1487<br />

Vitin, Alexander A. 878<br />

Vittal, Anusha 352, 1814, 1815<br />

Vivarelli, Marco 1751<br />

Viveiros, Kathleen 1592<br />

Vivianne, Gnemmi 2222<br />

Vivoli, Elisa 973<br />

Vlachaki, Efthimia 1126<br />

Vlachogiannakos, Jiannis 2012<br />

Vlaic, Sebastian 954<br />

Vogel, Arndt 344, 444, 2128<br />

Vogel, Robert 33, 2064<br />

Vogel, Wolfgang 1092, 1885<br />

Voidonikolas, Georgios 1479<br />

Voigt, Michael D. 1215<br />

Voitenleitner, Christian 35<br />

Volk, Neil 580, 1329, 2073<br />

Von Drygalski, Annette 1034<br />

von Felden, Johann 1803<br />

Von Roenn, Natasha 1172<br />

Vonderscher, Jacky 1652<br />

Vonghia, Luisa 797, 928<br />

Vongsuvanh, Roslyn 441<br />

Vos, Miriam B. 2232<br />

Voss, Jordan 565<br />

Voutila, Jon 118<br />

Vradelis, Stylianos 1900<br />

Vrang, Niels 978<br />

Vreden, Stephen G. 1846<br />

Vu, Vinh D. 207, 368, 386, 458, 1182,<br />

1561<br />

Vue, Jenny 1231<br />

Vue, Padade M. 1688<br />

Vukotic, Ranka 1888<br />

Vuppalanchi, Raj 1903, 2220<br />

Vutien, Philip 1668, 1805<br />

Vyas, Ashish 1663, 2088<br />

Vyas, Tanmay S. 1328, 2203<br />

Waasdorp Hurtado, Christine 1706,<br />

1708<br />

Wada, Fumitaka 688<br />

Wada, Nozomu 2236<br />

Wade, James 52, 1491<br />

Wagner, Martin 662<br />

Wagner, Norbert 963<br />

Wahl, Janice 40, 42, 210, 251, 700,<br />

701, 703, 712, 715<br />

Wahlang, Banrida 666<br />

Wahlin, Staffan 308<br />

Waintraub, Daniel J. 1095<br />

Wainwright, Helen 1667<br />

Waisbourd-Zinman, Orith 195<br />

Waisman, Ari 183, 1353, 1949<br />

Waitzberg, Dan 240<br />

Wakai, Toshifumi 2165<br />

Wakayama, Kenji 866<br />

Wakayama, Yuko 569, 2131, 2234<br />

Waked, Imam 344, 708, 733, 1163,<br />

1853<br />

Wakefield, Dorothy 1469<br />

Wakimoto, Yukio 1573, 1599, 1618<br />

Wakita, Takaji 985, 1003, 1679<br />

Wakui, Yuta 1648, 1943, 2167<br />

Walczak, Robert 974<br />

Walenbergh, Sofie 922<br />

Walewski, Jose L. 972<br />

Walia, Sandeep 1254<br />

Waljee, Akbar K. 418<br />

Walker, Alex J. 1140<br />

Walker, David R. 1050, 1139, 1152,<br />

1175<br />

Walker, Lucy J. 611<br />

Walker, Susan A. 1309<br />

Wallace, Paul K. 332<br />

Walsh, Christopher 1034<br />

Walsh, Nick 577<br />

Walsh, Renae 1551<br />

Wamhoff, Brian R. 175, 664, 915,<br />

1916<br />

Wan, Jinghong 1391<br />

Wan, Ying 190, 514, 817, 834, 835,<br />

929, 1317, 1349<br />

Wan, Yu-Jui Yvonne 63, 508, 649, 674<br />

Wan, Zhihong 1609<br />

Wandeler, Gilles 1575<br />

Wands, Jack R. 256, 1962<br />

Wang, Chong 2018<br />

Wang, Alice Y. 1087, 1143, 1144,<br />

1462, 2184<br />

Wang, Amber W. 511<br />

Wang, Bruce M. 683, 2111<br />

Wang, Cheng 1536, 1537, 2008<br />

Wang, Chia-Chi 1996<br />

Wang, Chih-Chi 1942<br />

Wang, Chuan 2018<br />

Wang, Chunping 354<br />

Wang, Clay 1936<br />

Wang, Connie W. 844, 1219, 1468<br />

Wang, Deli 1039<br />

Wang, Dongsheng 1149<br />

Wang, Fu-Sheng 2179<br />

Wang, Gary 993<br />

Wang, Gui-Qiang 1564, 1602, 2023<br />

Wang, Guo-Ying 419, 440<br />

Wang, Hongliang 27<br />

Wang, Hongtao 194<br />

Wang, Hongwu 909<br />

Wang, Hongyan 1656<br />

Wang, Horng-Yuan 1996<br />

Wang, Hua 110<br />

Wang, Hui 1611<br />

Wang, Jiang-Bin 1465<br />

Wang, Jianhong 754<br />

Wang, Jiao-Jing 54<br />

Wang, Jiaohong 1952<br />

Wang, Jiayou 879, 1324, 1338<br />

Wang, Jie 34<br />

Wang, Jing 205, 613, 1034, 1541,<br />

1579, 1617, 1773, 2151<br />

Wang, Jing-Houng 1873, 2010<br />

Wang, Jingyun 1580, 1620<br />

Wang, Jiuping 1773<br />

Wang, Jiyao 131, 1465, 1586, 2002<br />

Wang, Junfeng 74, 623<br />

Wang, Kasper S. 133<br />

Wang, Ke 1772<br />

Wang, Lai Mun 2190<br />

Wang, Lan S. 1095<br />

Wang, Lan Sun 2214<br />

Wang, Li 1336, 1865<br />

Wang, Li-Yu 1593, 1604<br />

Wang, Lin 1360<br />

Wang, Lirui 59, 111<br />

Wang, Mengjun 371<br />

Wang, Mo-li 1149, 1193<br />

Wang, Peipei 1423<br />

Wang, Ping 189, 1404<br />

Wang, Qi 1184<br />

Wang, Qianfan 36<br />

Wang, Qianyi 374<br />

Wang, Qiaomin 1537<br />

Wang, Qixia 308<br />

Wang, Reena 720, 728<br />

Wang, Renxue 824<br />

Wang, Rong-qi 638<br />

Wang, Ruisi 1315, 1355<br />

Wang, Shaogui 1325<br />

Wang, Shaoqing 1966<br />

Wang, Shu-Chi 2264<br />

Wang, Sihyung 696<br />

Wang, Stanley 248, 250, 719<br />

Wang, Tao 943<br />

Wang, Ting 1528<br />

Wang, Tingting 1537<br />

Wang, Weimin 510, 2116<br />

Wang, Wen-Jun 321<br />

Wang, Wenjun 468<br />

Wang, Xiang 1521<br />

Wang, Xiaofeng 1656<br />

Wang, Xiaohong 1865<br />

Wang, Xiaojie 104, 151<br />

Wang, Xiaojing 909, 1660


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1361A<br />

Wang, Xiaomei 1149, 1193, 1580,<br />

1603, 1620, 1869<br />

Wang, Xiaoming 374<br />

Wang, Xiaoxin 139, 142, 907<br />

Wang, Xiaoyu 183, 955, 1353, 1376<br />

Wang, Xintao 333<br />

Wang, Xue-Yan 1973<br />

Wang, Xuemei 130<br />

Wang, Xuyang 168<br />

Wang, Yan 193, 197<br />

Wang, Yang 704, 717<br />

Wang, Yaqi 909<br />

Wang, Yen-Po 738, 751<br />

Wang, Yifeng 140<br />

Wang, Yin-Chen 2050<br />

Wang, Ying 686<br />

Wang, Yongping 919<br />

Wang, Yongqing 22<br />

Wang, Yu 374, 898, 1966, 1973<br />

Wang, Yucai 295<br />

Wang, Yudong 1536, 1537, 2008<br />

Wang, Yuming 209, 1865, 2024<br />

Wang, Yunwu 328<br />

Wang, Zhengyu 754<br />

Wangensteen, Kirk J. 494, 511<br />

Wani, Hamid 1188<br />

Wanichanuwat, Jirachaya 1999<br />

Ward, Stephen C. 870, 1337<br />

Warner, Alex 1455<br />

Warner, Nikole L. 2142<br />

Washburn, W. Kenneth 1900<br />

Washington, Shenna 946<br />

Wasitthankasem, Rujipat 2036<br />

Wasser, Martin N.J.M. 1238<br />

Wat, Cynthia 245, 2026<br />

Watabe, Takeshi 1795<br />

Watanabe, Hiroshi 1912<br />

Watanabe, Mamoru 164, 406, 1018,<br />

1027<br />

Watanabe, Sumio 226, 794, 917,<br />

1771, 2158, 2193<br />

Watanabe, Takako 406, 1018, 1027<br />

Watanabe, Takao 788, 1293<br />

Watanabe, Takehisa 513, 1912, 1987<br />

Watanabe, Tsunamasa 1006, 1446,<br />

1655<br />

Watashi, Koichi 1679<br />

Watkins, Jennifer 2081<br />

Watkins, Paul B. 96, 224, 225, 228,<br />

1930, 1932, 1933<br />

Watkins, Steve 1359<br />

Watt, Gordon P. 1424<br />

Watt, Jennifer 129<br />

Watt, Kymberly D. 836, 1038, 1124<br />

Wattacheril, Julia J. 2166<br />

Watts, Micah M. 521<br />

Watts, Michelle 1172<br />

Waziry, Reem K. 1868<br />

Weatherburn, Peter 1830<br />

Webb, Glynn 523<br />

Webb, Gwilym 631<br />

Weber, Achim 46, 1402<br />

Weber, Bernd 1170, 1200<br />

Weber, Susan C. 359, 386, 388, 458,<br />

1806, 1816, 1848<br />

Wedd, Joel P. 2085<br />

Wedemeyer, Heiner 445, 1002, 1086,<br />

1107, 1131, 1166, 1572, 1597,<br />

1797, 1803, 1861, 2003, 2026<br />

Wee, Aileen -. 1436<br />

Weerachayaphorn, Jittima 661<br />

Weersma, Rinse K. 76<br />

Wehrkamp, Cody J. 60<br />

Wei, Lai 782, 1436, 1820, 1850,<br />

1877, 1973<br />

Wei, Lan 957, 1938<br />

Wei, Liang 341<br />

Wei, Wei 497<br />

Wei, Xiaoli 114, 926<br />

Weick, Alexander 1254<br />

Weigert, Roberto 20<br />

Weijer, Sebastiaan 1995<br />

Weil, Delphine 1036, 1187<br />

Weiland, Ola 37, 252, 1058<br />

Weimer, Liliana E. 1802, 1844<br />

Weinberg, Ethan 1823<br />

Weiner, Joshua 2129<br />

Weinger, Matthew B. 529<br />

Weinman, Edward J. 829<br />

Weinman, Steven A. 352, 1291, 1311,<br />

1814, 1815<br />

Weinmann, Arndt J. 376<br />

Weinreb, Paul 1368<br />

Weinstein, Ali A. 18, 316<br />

Weinstein, Jeffrey S. 1146, 1169<br />

Weir, Graeme 753<br />

Weismuller, Tobias J. 76<br />

Weiss, Emmanuel 265, 1288, 1391<br />

Weiss, Jeffrey J. 1151, 1870<br />

Weiss, Karl Heinz 618<br />

Weiss, Nicolas 1510<br />

Weiss, Thomas S. 954<br />

Wek, Ronald C. 320, 884<br />

Weksbergc, Rosanna 682<br />

Welker, Martin W. 1055<br />

Welle, Stephen L. 671<br />

Weller, Shaun 1310<br />

Wells, James 676<br />

Wells, Malcolm M. 2170<br />

Wells, Rebecca G. 195, 814, 1366<br />

Weltman, Martin 106, 627, 1343,<br />

1450<br />

Welzel, Tania M. 37, 252, 1045,<br />

1057, 1058, 1106, 1130, 1860, 2003<br />

Wen, Li 305<br />

Wen, Simin 2018<br />

Wen, Wan-Hsin 1701<br />

Wendon, Julia 212, 740, 1758, 1763,<br />

1775<br />

Weng, Cindy 2112<br />

Weng, Shih-Yen 183, 955, 1353, 1376<br />

Werba, Gregor 1282<br />

Werby, Jasper 462<br />

Werner, Christoph R. 1183<br />

Werner, Mårten 308<br />

West, Sara 552<br />

Westbrook, Rachel 764, 1216<br />

Westlin, William 957, 1938<br />

Weston, Chris J. 1296<br />

Wheatley, Daniel J. 1321<br />

Wheeler, David A. 129<br />

White, Duncan 536<br />

White, Jon 130, 682, 2118<br />

White, Melanie 52, 1463, 1472, 1491,<br />

1527<br />

Whitener, Melissa 1814, 1815<br />

Whitington, Peter F. 159, 1694<br />

Wiedtke, Ellen 98<br />

Wiegand, Steffen B. 2026<br />

Wiegel, Joshua 1101<br />

Wiesner, Russell H. 1249<br />

Wievel, Emily 1839<br />

Wigg, Alan J. 106, 627, 1077, 1264<br />

Wiggins, Shanna M. 136<br />

Wijarnpreecha, Karn 964<br />

Wilairatana, Polrat 555<br />

Wilder, Julius M. 1887<br />

Wildhaber, Barbara E. 1735<br />

Wildner, Letícia M. 2075<br />

Wilkey, Daniel W. 587<br />

Wilkins, Benjamin J. 814<br />

Wilkins, Kenneth J. 779, 1717, 2144<br />

Willacy, Oliver 337, 1388<br />

Willandt, Barbara 171<br />

Willars, Christopher 212, 1758, 1775<br />

Willebrords, Joost 586<br />

Willems, Bernard E. 1045, 1997<br />

Willemse, Sophie 208, 1995, 2046,<br />

2263<br />

Willemssen, Francois 1238<br />

Willenbring, Holger 98, 680<br />

Willett, Michael S. 2247<br />

Williams, Bianca 1314<br />

Williams, Michael J. 678<br />

Williams, Roger 163, 599, 1978, 2143<br />

Williamson, Kate D. 74, 622<br />

Williamson, Rachel M. 2168<br />

Willis, Sarah 1114<br />

Willms, Arnulf G. 43<br />

Willy, Jeffrey A. 320, 884<br />

Wilson, Antoinette 527<br />

Wilson, Caroline 44<br />

Wilson, Eleanor 92, 220, 578<br />

Wilson, Laura A. 107<br />

Wiltberger, Georg 1222<br />

Win, Sanda 590<br />

Wind-Rotolo, Megan 28<br />

Winkler, Anja 1927, 2108<br />

Winokur, Allison 2175<br />

Winwood, Paul J. 1413<br />

Wires, Emily 323<br />

Wise, Eric S. 1223<br />

Witek, James 1040<br />

Witt, David J. 1053<br />

Witters, Peter 1691<br />

Witthoeft, Thomas 1170, 1200<br />

Wöbse, Michael 1572<br />

Wochner, Katharina 1525<br />

Wockner, Leesa F. 784<br />

Woerns, Marcus A. 1949<br />

Wohl, David 1063<br />

Wolak, Arik 2224<br />

Wolfe, Lynne A. 1722<br />

Wolff, Rodrigo 1217<br />

Wolford, Dennis 725<br />

Wolkoff, Allan W. 321, 333, 943<br />

Wolter, Franziska 1287<br />

Won, Je Hwan 379<br />

Wong, Amanda A. 1110<br />

Wong, April 1536, 1537<br />

Wong, Danny 126, 1545, 1550<br />

Wong, David K. 243, 348, 542, 1556,<br />

1998<br />

Wong, Florence 82, 278, 570<br />

Wong, Grace 367, 459, 2181<br />

Wong, Grace LH 235, 238, 349, 1548,<br />

1612<br />

Wong, Guan Wee 312<br />

Wong, John B. 1592<br />

Wong, Kelly A. 2257


1362A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Wong, Kristin N. 388<br />

Wong, Linda L. 1553<br />

Wong, Ophelia H. 47<br />

Wong, Philip 1045<br />

Wong, Robert J. 7, 199, 201, 237,<br />

306, 534, 877, 1252, 1534, 1549,<br />

1745, 1750, 2159<br />

Wong, She-Yan 1261, 1276, 1452<br />

Wong, Tiffany C. 1746, 2000<br />

Wong, Vincent W. 235, 238, 349,<br />

367, 459, 1548, 1612, 2181<br />

Wong, William W. 350<br />

Woo, Hyun Young 749, 876, 2028,<br />

2038<br />

Woo, Jacky 2052<br />

Woodard, Colin 1625<br />

Wooddell, Christine I. 32, 2022<br />

Woods, Donna 71, 843, 847<br />

Woods, Ryan 543<br />

Woody, Angela 1095, 1870<br />

Woolbright, Benjamin 585<br />

Woolson, Kathy L. 523<br />

Woolwine-Cunningham, Yvonne J. 332<br />

Workowski, Kimberly 1034<br />

Wörns, Marcus 376<br />

Wrba, Friedrich 297<br />

Wree, Alexander 55, 969, 1371,<br />

1522, 1971, 2185<br />

Wright, Elizabeth C. 218<br />

Wright, Matthew C. 44<br />

Wright, Melanie 2147, 2208<br />

Wu, Bechien U. 404<br />

Wu, Catherine H. 223<br />

Wu, Cheng-Kun 1873<br />

Wu, Christina 1121<br />

Wu, Chuanghong 1814, 1815<br />

Wu, Chun-Ying 345, 1547, 1589<br />

Wu, David 223<br />

Wu, Dehua 1939<br />

Wu, Di 1656, 1683<br />

Wu, Diana Y. 1625<br />

Wu, Elizabeth 1850, 1877<br />

Wu, George 223, 2014<br />

Wu, Guihui 904<br />

Wu, Hao 19<br />

Wu, Heng 102<br />

Wu, Hui-Lin 2027<br />

Wu, Jaw-Ching 759, 1639, 1665<br />

Wu, Jia-Feng 1701, 1705<br />

Wu, Jian 366, 935<br />

Wu, Jiashin 879, 1324, 1338<br />

Wu, Jingtao 710<br />

Wu, Li-Chen 654<br />

Wu, Min 2058<br />

Wu, Ming-Heng 1400<br />

Wu, Ming-Shiang 345, 1547<br />

Wu, Nan 190, 782, 813, 817, 834,<br />

835, 929, 1317, 1349<br />

Wu, Pengfei 904<br />

Wu, Quanfeng 989, 1967<br />

Wu, Quanxin 2024<br />

Wu, Raymond 1936<br />

Wu, Ruihong 1149, 1193, 1603,<br />

1620, 1869<br />

Wu, Shuang 1628<br />

Wu, Ting 909<br />

Wu, Tong 185, 501, 506, 686, 1096<br />

Wu, Vanessa 1536, 1537, 2008<br />

Wu, Wen-rui 498, 1951<br />

Wu, Xiaoning 374<br />

Wu, Xing 2018<br />

Wu, Yan 1222<br />

Wu, Yang 54, 1343, 1450<br />

Wu, Yujing 131, 1405<br />

Wu, Zeguang 1660<br />

Wulff, Jacob 161<br />

Wursthorn, Karsten 1176, 1205<br />

Wyatt, Brooke 1151<br />

Wyles, David L. 248, 250, 716<br />

Wymer, Mark 153, 2212<br />

Wymore, Erin 71, 843, 847<br />

Wyrzykiewicz, Tadeusz 36<br />

Wysocki, Jacek 2043<br />

Xavier, Torras 734<br />

Xavier Brito, Leonor 150<br />

Xia, Jielai 468<br />

Xia, Yuchen 1684<br />

Xiang, Junxi 698<br />

Xiang, Xiaogang 1611<br />

Xiang, Xiaoyu 1947<br />

Xiao, Gary S. 202<br />

Xiao, Li 1441<br />

Xiao, Xiaoyan 305<br />

Xie, Guanhua 66, 936, 1386<br />

Xie, Jun 496<br />

Xie, Pei-yi 1755<br />

Xie, Qing 1586, 1611, 2002, 2013,<br />

2023, 2048<br />

Xie, Wangang 1084, 1086<br />

Xie, Xing-Wang 1973<br />

Xie, Yan 54, 911<br />

Xie, Yao 2048<br />

Xie, Yi 2048<br />

Xin, ShaoJie 1609, 1675<br />

Xing, Jian 15, 1789, 1800<br />

Xu, Aimin 898<br />

Xu, Che 1910<br />

Xu, Cheng 1865<br />

Xu, Dongping 1595, 1675<br />

Xu, Fujie 15, 525, 1789, 1800<br />

Xu, Gang 366<br />

Xu, Hong 54<br />

Xu, Hongqin 1149, 1580, 1620, 1869<br />

Xu, Huilei 36<br />

Xu, Jason J. 2083<br />

Xu, Jian-Ming 1465<br />

Xu, Jianfang 1583<br />

Xu, Jun 186, 187, 890, 904<br />

Xu, Lei-bo 498, 1951<br />

Xu, Ming-Jiang 110, 334, 903<br />

Xu, Xiaoyuan 2066<br />

Xu, Yang 2066<br />

Xu, Yongcai 67<br />

Xu, Yuqin 2048<br />

Xu, Zhihui 1595<br />

Xuan, Lei 548<br />

Xue, Bingzhong 933<br />

Xue, Yuhua 1979<br />

Yaari, Shaul 1878<br />

Yab, Tracy C. 169<br />

Yada, Akito 188, 1420<br />

Yada, Masayoshi 1157<br />

Yadav, Ajay K. 1664<br />

Yagi, Hiroshi 336<br />

Yagi, Takahito 339<br />

Yakushijin, Takayuki 453, 1195, 1197,<br />

1201, 2213<br />

Yalamanchili, Rachana 1185, 1202<br />

Yalinaycirak, Meltem 932<br />

Yamada, Akira 1197<br />

Yamada, Gotaro 2030<br />

Yamada, Kazutoshi 370<br />

Yamada, Makoto 2156<br />

Yamada, Masanobu 477<br />

Yamada, Ryoko 1195, 1197, 1201,<br />

2213<br />

Yamada, Takuya 1584<br />

Yamada, Yukinori 1197<br />

Yamagiwa, Satoshi 310, 2165<br />

Yamagiwa, Yoko 794, 1138, 1795<br />

Yamaguchi, Akihiro 181, 569, 2131,<br />

2140, 2234<br />

Yamaguchi, Kanji 886, 888, 901,<br />

1980, 2230<br />

Yamaguchi, Masakatsu 1116<br />

Yamai, Takuo 1661<br />

Yamamoto, Akiko 906, 2156<br />

Yamamoto, Gen 1213<br />

Yamamoto, Ken 1632<br />

Yamamoto, Kuniko 1326<br />

Yamamoto, Naoki 329, 1447<br />

Yamamoto, Takashi 1627<br />

Yamamoto, Yasunori 2146<br />

Yamanaka, Junichi 1420<br />

Yamane, Keiko 482<br />

Yamasaki, Chihiro 1679<br />

Yamasaki, Kazumi 408, 783, 1600<br />

Yamasato, Florencia 2106<br />

Yamashina, Shunhei 226, 917, 1771<br />

Yamashita, Naoki 1159<br />

Yamashita, Shunichi 551<br />

Yamashita, Taro 144, 699, 1000,<br />

1348, 1395, 1653, 1953, 1968, 1970<br />

Yamashita, Tatsuya 392, 1348, 1653<br />

Yamato, Masatoshi 325, 947<br />

Yamauchi, Akira 1937<br />

Yamazoe, Taiji 513<br />

Yan, Fengxia 790<br />

Yan, Hongqing 1193<br />

Yan, Irene K. 335, 358, 384<br />

Yan, Libo 1562<br />

Yan, Lunan 851<br />

Yan, Taotao 1541, 1579, 1617, 1773<br />

Yan, Wei 328<br />

Yan, Weiming 1656, 1660, 1683<br />

Yan, Xiaohong 239, 2153, 2154<br />

Yan, Zehui 1865<br />

Yanagawa, Takayo 1369<br />

Yanagi, Ami 1679<br />

Yanagita, Kimihiko 2031<br />

Yang, Zhihong 1336<br />

Yang, Guo-Xiang 77, 602, 613, 639,<br />

1285<br />

Yang, Heping 1952<br />

Yang, Hong 198, 1986<br />

Yang, Hushan 371<br />

Yang, Hwai-I 207, 1561, 1604, 1828<br />

Yang, Hyun 452, 1882<br />

Yang, Jenny C. 219, 718, 1080, 1105,<br />

1120, 1133, 1168<br />

Yang, Jenny D. 1549<br />

Yang, Jiaqi 361<br />

Yang, Jijin 851<br />

Yang, Jin-Hui 1465<br />

Yang, Jing 22, 1521<br />

Yang, JinMo 1615<br />

Yang, Jr-Shiuan 2116<br />

Yang, Ju Dong 127, 381, 415, 1124<br />

Yang, Ju-Il 1960<br />

Yang, Kai-ting 1945


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1363A<br />

Yang, Lei 245, 2026<br />

Yang, Lifei 698<br />

Yang, Lin 1396<br />

Yang, Ling 1441<br />

Yang, Luoluo 2135<br />

Yang, Man 468<br />

Yang, Michael 549<br />

Yang, Ming 1436, 1850, 1877<br />

Yang, Rong 709<br />

Yang, Ruifeng 1820<br />

Yang, Sera 1972<br />

Yang, Sheng-Shun 1996<br />

Yang, Sien-Sing 2050<br />

Yang, Su-Jau 87<br />

Yang, Sun 516<br />

Yang, Tzeng-Huey 1547<br />

Yang, Victor 1015<br />

Yang, Wenli 79<br />

Yang, Wucai 1536<br />

Yang, Xiaoyan 709<br />

Yang, Yang 419, 440<br />

Yang, Yin 38<br />

Yang, Ying 1306<br />

Yang, Ying-Ying 965<br />

Yang, Yongfeng 2048<br />

Yang, Yongping 354<br />

Yang, Yuan 1541, 1579, 1617, 1773<br />

Yang, Yuching 224, 228<br />

Yang, Zhen 516<br />

Yang, Zhiping 754<br />

Yanko, Adir 1032<br />

Yanny, Beshoy T. 1119<br />

Yano, Hirohisa 465, 1990<br />

Yao, Francis Y. 4, 50, 53, 414, 471<br />

Yao, Liying 1293<br />

Yao, Lu 185, 686<br />

Yao, Zhi Q. 984, 1004<br />

Yapali, Suna 1247<br />

Yaqoob, Usman 1355, 1406<br />

Yara, Sho-ichiro 970<br />

Yardeni, David 2224<br />

Yarmush, Martin L. 25<br />

Yarnykh, Vasily L. 803<br />

Yartel, Anthony K. 1792<br />

Yaseen Alsabbagh, Mohammed Eyad<br />

1122<br />

Yasuchika, Kentaro 402, 1213, 1976<br />

Yasuda, Katsutaro 1976<br />

Yasui, Kohichiroh 886, 888, 901,<br />

1980, 2230<br />

Yasui, Shin 1207<br />

Yasui, Yutaka 1115<br />

Yasukawa, Lee Ann 359, 386, 388,<br />

458, 1806, 1816, 1848<br />

Yasumura, Seiji 551<br />

Yasunaka, Tetsuya 2236<br />

Yates, Denise 260, 753<br />

Yates, Katherine P. 160, 2161<br />

Yatsuhashi, Hiroshi 408, 783, 1448,<br />

1600, 1632<br />

Yatsuzuka, Naoyoshi 707<br />

Yazdani, Hamza 184<br />

Yazdanpanah, Yazdan 1845<br />

Ye, Dewei 898<br />

Ye, Juan 366<br />

Ye, Wen 133<br />

Ye, Xin 2007, 2062<br />

Yeap, Sze Pheh 1264<br />

Yee, Brittany E. 2250<br />

Yee, Leland J. 241, 2014, 2025<br />

Yeh, Heidi 1732<br />

Yeh, Matthew M. 230, 923, 946<br />

Yeh, Ming-Lun 2264<br />

Yeh, Wendy W. 725, 730, 731<br />

Yen, Elizabeth H. 2149<br />

Yen, Yi-Hao 1873, 1942<br />

Yeo, Winnie 344<br />

Yeoman, Andrew D. 312<br />

Yeon, Jong Eun 391, 405, 1504, 1897,<br />

2017<br />

Yeung, Chun-Yan 1701<br />

Yi, Li-Na 1942<br />

Yi, Nam-Joon 400, 474, 1277<br />

Yi, Ruitian 1541<br />

Yi, Sooji 1554<br />

Yi, Wenjing 1004<br />

Yildirim, Beytullah 1613<br />

Yilmaz, Abdulkerim 2219<br />

Yilmaz, Funda 1739<br />

Yilmaz, Guldal 1719<br />

Yilmaz, Yusuf 2180<br />

Yim, Colina 348, 647<br />

Yim, Hyung Joon 391, 405, 1438,<br />

1504, 1897<br />

Yim, Sun Young 433, 1438, 1504<br />

Yimam, Kidist K. 76<br />

Yin, Xin 1703<br />

Yin, Chi-Biao 2023<br />

Yin, Hong 19<br />

Yin, Meng 1355<br />

Yin, Michael 380<br />

Yin, Paley 1587<br />

Yin, Philip 702, 706<br />

Yin, Wenwei 1647<br />

Yin, Xiaoyu 454<br />

Yin, Xinmin 114, 926<br />

Yin, Zhanxin 468, 754<br />

Ying, Jun 357<br />

Ying, Qilong 1283<br />

Ying, Wenbin 1433<br />

Yip, Benjamin 2250<br />

Yip, Marcus 29, 1852<br />

Yip, Terry C. 349<br />

Yoh, Kazunori 781, 1669<br />

Yokohama, Keisuke 966<br />

Yokomori, Hiroaki 1514<br />

Yokoo, Hideki 866<br />

Yokosuka, Osamu 1207, 1490, 1516,<br />

1628, 1632<br />

Yokota, Masaki 1159<br />

Yokote, Seiichiro 513<br />

Yokoyama, Keiji 385, 410<br />

Yokoyama, Takahide 869<br />

Yokoyama, Yoshinobu 1636, 1661<br />

Yoneda, Akihiro 1375, 1378<br />

Yoneda, Masato 785, 1017<br />

Yonem, Ozlem 2219<br />

Yonemoto, Koji 304<br />

Yong, Sherri 774<br />

Yoo, Byung Chul 423, 1960, 2049<br />

Yoo, Jeong-Ju 242, 400, 447, 470,<br />

474, 479, 1610, 2016, 2042, 2055<br />

Yoo, Suk Young 2151<br />

Yoo, Wonbeak 2151<br />

Yoo, Yang J. 405<br />

Yoo, Yang Jae 1897, 2017<br />

Yoon, Dukyong 2056<br />

Yoon, Jung-Hwan 242, 400, 447, 470,<br />

474, 479, 1610, 1907, 2016, 2042,<br />

2055<br />

Yoon, Ki Tae 876, 1542, 2028<br />

Yoon, Seung Kew 452, 502, 1025,<br />

1419, 1615, 1882<br />

Yoon, Yoo-Seok 857, 1902<br />

York, Samuel 778<br />

Yoshida, Atsushi 1254, 1274<br />

Yoshida, Eric M. 204, 1040, 1049<br />

Yoshida, Hitoshi 1302, 1941<br />

Yoshida, Kanako 1824<br />

Yoshida, Mariko 699, 1953<br />

Yoshida, Osamu 338, 788, 1293<br />

Yoshida, Shuhei 1379<br />

Yoshida, Yuichi 227, 906, 1195, 1197,<br />

1201, 1974, 2156, 2213<br />

Yoshihara, Mie 2267<br />

Yoshiji, Hitoshi 677, 689<br />

Yoshimaru, Youko 1912, 1987<br />

Yoshimi, Satoshi 1858<br />

Yoshimoto, Tsuyoshi 1159<br />

Yoshio, Sachiyo 476, 1643<br />

Yoshioka, Teppei 26, 1636, 1809<br />

Yoshioka, Wataru 296<br />

Yoshitoshi, Elena Y. 1976<br />

Yoshiza, Kai 1097<br />

Yoshizane, Yasumi 1679<br />

Yoshizato, Katsutoshi 191<br />

Yosry, Ayman 708<br />

Yotsuyanagi, Hiroshi 912<br />

Yotti, Raquel 1503<br />

You, Hong 374, 1404<br />

You, Kazunori 466, 770, 1533<br />

You, Min 879, 1324, 1338<br />

You, Qiang 1910<br />

You, San-Lin 1593, 1604<br />

You, Shaoli 1609, 1675<br />

Younes, Ziad 219, 1051, 1067<br />

Young, Elisa 106, 627<br />

Young, Kellie 237<br />

Younis, Rasha 1978, 2143<br />

Younossi, Issah 18<br />

Younossi, Zobair M. 18, 87, 153, 199,<br />

201, 315, 316, 534, 575, 877, 891,<br />

892, 956, 1035, 1079, 1177, 1186,<br />

1252, 1460, 1534, 1538, 1745,<br />

1750, 1788, 2159, 2173, 2192,<br />

2204, 2205, 2212<br />

Younoszai, Zahra 892, 2173, 2192<br />

Youssef, Amira S. 1821<br />

Yu, Amanda 86, 543<br />

Yu, Ami 377<br />

Yu, Dominic 413, 793, 1530<br />

Yu, Fei 709<br />

Yu, Ge 1149<br />

Yu, Haibin 1406<br />

Yu, Herbert 1553<br />

Yu, Lei 1211<br />

Yu, Lijia 996, 1225<br />

Yu, Ming-Lung 367, 398, 459, 1828,<br />

2264<br />

Yu, Su Jong 242, 400, 447, 470, 474,<br />

479, 1610, 1907, 2016, 2042, 2055<br />

Yu, Tunan 1962<br />

Yu, Xian-huan 498, 1951<br />

Yu, Yao 714<br />

Yu, Zhangsheng 1316<br />

Yu, Zu-Jiang 1616<br />

Yuan, Junying 832<br />

Yuan, Yong 711, 1456, 1461, 1828<br />

Yuasa, Korta 552<br />

Yudina, Tetyana 659


1364A AUTHOR INDEX HEPATOLOGY, October, 2015<br />

Yue, Shi 1283, 2138<br />

Yuen, Lilly 1551, 1557, 1563, 1566<br />

Yuen, Man-Fung 126, 1545, 1550,<br />

1746<br />

Yukawa, Toyokazu 2074<br />

Yuksel, Muhammed 305<br />

Yun, Tae Jung 1438<br />

Yurdaydin, Cihan 1572, 1678, 2004,<br />

2012, 2033<br />

Yusuf, Aasim 1872<br />

Zaaijer, Hans L. 1995<br />

Zabron, Abigail 212<br />

Zachou, Kalliopi 73<br />

Zaghla, Hassan E. 733<br />

Zaidan Dagli, Maria Lucia 586<br />

Zakharia, Kais 396, 1966<br />

Zamboni, Fausto 166, 1646<br />

Zamor, Philippe J. 17, 210, 715, 1023<br />

Zamora, Javier 745<br />

Zamora-Valdes, Daniel 836<br />

Zampino, Rosa 1837<br />

Zanetto, Alberto 351, 1270, 1889<br />

Zanieri, Francesca 949, 1963<br />

Zapata, E. 583<br />

Zapata, Homero A. 1281<br />

Zapata, Rodrigo L. 595<br />

Zarski, Jean-Pierre H. 1439<br />

Zatsiorsky, James V. 496<br />

Zbrzezniak, Justyna 1502<br />

Zehnter, Elmar 1170, 1200<br />

Zeilinger, Katrin 1921<br />

Zein, Nizar N. 840, 855, 856, 1265<br />

Zekri, Abdelrahman 1821<br />

Zemel, Michael 933<br />

Zenatti, Vanessa L. 2072, 2076<br />

Zendejas, Ivan 177<br />

Zeng, Changjun 685<br />

Zeng, Hong 498, 1951<br />

Zeng, Ni 1945<br />

Zeng, Qing-Lei 1616<br />

Zeng, Xianghua 1865<br />

Zeng, Zheng 1587<br />

Zeng, Zhenzhen 1634<br />

Zeng, Zhiping 131<br />

Zeremski, Marija 1841<br />

Zerial, Marino 617<br />

Zerkly, Salah 2082<br />

Zern, Mark 685<br />

Zeuzem, Stefan 37, 91, 94, 249, 252,<br />

445, 700, 701, 703, 714, 980, 1045,<br />

1055, 1057, 1058, 1086, 1106,<br />

1130, 1166, 1559, 1860, 1861,<br />

2001, 2013, 2257, 2260<br />

Zhai, Xiangjun 1583<br />

Zhan, Le 919, 934<br />

Zhang, Chen 1910<br />

Zhang, Yanqiong 2024<br />

Zhang, David 1405<br />

Zhang, David Y. 1392<br />

Zhang, Dazhi 1647<br />

Zhang, DingYu 1412<br />

Zhang, Dongyan 1939<br />

Zhang, Fang 218<br />

Zhang, Haiyun 1441<br />

Zhang, Hang 1301<br />

Zhang, Heng-Hui 1973<br />

Zhang, Hong 2058<br />

Zhang, Huaihong 209<br />

Zhang, Ji-Yuan 2179<br />

Zhang, Jian Q. 207, 359, 368, 388,<br />

1561, 1806, 1848<br />

Zhang, Jie 2261<br />

Zhang, Jingwen 1304, 1312, 1331<br />

Zhang, Jinjin 589<br />

Zhang, Jinqiang 686<br />

Zhang, Junlan 79<br />

Zhang, Kunsong 454<br />

Zhang, Lei 468<br />

Zhang, Li 904<br />

Zhang, Lihua 1305<br />

Zhang, Lumin 1397<br />

Zhang, Min 120, 1772<br />

Zhang, Ming-Xiang 2023, 2048<br />

Zhang, Nan 159<br />

Zhang, Ning-Ping 366, 935, 2002<br />

Zhang, Peixin 1004<br />

Zhang, Peng 418, 1149<br />

Zhang, Qin 1586<br />

Zhang, Qiongfang 1647<br />

Zhang, Renwen 2066<br />

Zhang, Rui 498, 1951<br />

Zhang, Shanshan 1948<br />

Zhang, Shengli 2035<br />

Zhang, Shulin 1579, 1617<br />

Zhang, Shuqin 209, 2048<br />

Zhang, Wei 782<br />

Zhang, Weici 77, 613, 619, 639<br />

Zhang, Wenliang 1335<br />

Zhang, Wujuan 916, 1695<br />

Zhang, Xiang 114, 926<br />

Zhang, Xiaogang 1773<br />

Zhang, Xiaoping 1660<br />

Zhang, Xiaxia 2066<br />

Zhang, Xiugen 2126<br />

Zhang, Yanling 685<br />

Zhang, Ying 984, 1004<br />

Zhang, Yizhou 663, 986, 1649, 1681,<br />

1944<br />

Zhang, Yu 1541, 2066<br />

Zhang, Yuanqing 131<br />

Zhang, Yuanya 1656<br />

Zhang, Yuguo 638<br />

Zhang, Yuxia 1336<br />

Zhang, Zheng Z. 54<br />

Zhang, Zhongyang 253, 254<br />

Zhang, Zhuoli 468<br />

Zhao, Changqing 359, 367, 386, 387,<br />

388, 428, 458, 459, 1182, 1806,<br />

1816, 1848<br />

Zhao, Chenyang 668<br />

Zhao, Cuiqing 1306<br />

Zhao, Enpeng 1333<br />

Zhao, Gang 828, 1924<br />

Zhao, Hong 1564, 1602, 1644<br />

Zhao, Huaying 1646<br />

Zhao, Jie 1311, 1814<br />

Zhao, Jun 1595, 1675<br />

Zhao, Lu-Lu 1701<br />

Zhao, Ming 613<br />

Zhao, Qiyi 120, 1423<br />

Zhao, Suxian 638<br />

Zhao, Weihan 710, 723<br />

Zhao, Wen Jing 209<br />

Zhao, Xiao 814<br />

Zhao, Xinyan 374<br />

Zhao, Yan 468<br />

Zhao, Yang-Jing 1973<br />

Zhao, Yi-Ming 366<br />

Zhao, Yingren 1541, 1579, 1617,<br />

1773<br />

Zhao, Yue 37, 252, 1058<br />

Zhao, Zhongxin 1396<br />

Zheng, Huanwei 2048<br />

Zheng, Hui 2166, 2194<br />

Zheng, Xinglong 698<br />

Zheng, Yubao 1772<br />

Zhong, Jin 1149<br />

Zhong, Li 496<br />

Zhong, Shuping 460<br />

Zhong, Wei 1335<br />

Zhou, Chan 778<br />

Zhou, Hao 675<br />

Zhou, Haoyue 919<br />

Zhou, Huiping 22, 804, 1521<br />

Zhou, Kali 577<br />

Zhou, Lili 255<br />

Zhou, Mei 198, 1986<br />

Zhou, Nannan 709<br />

Zhou, Tianhao 813, 1317, 1349<br />

Zhou, Tiaohao 929<br />

Zhou, Wei 2116, 2261<br />

Zhou, Xin-Min 1465<br />

Zhou, Xingliang 1283<br />

Zhou, Yueren 525, 1799<br />

Zhou, Yun 984<br />

Zhou, Zhanxiang 1335<br />

Zhou, Zhou 182, 903<br />

Zhu, Airu 904<br />

Zhu, Bao Shen 209<br />

Zhu, Bing 1609<br />

Zhu, Chuanlong 1644<br />

Zhu, Chuanwu 2048<br />

Zhu, Fusheng 2048<br />

Zhu, Ji-Min 1964<br />

Zhu, Liguo 1583<br />

Zhu, Lin 1660<br />

Zhu, Liqing 1441<br />

Zhu, Mansheng 498, 1951<br />

Zhu, Mingyue 1977<br />

Zhu, Qingjing 1441<br />

Zhu, Weiping 2048<br />

Zhu, Xuan 1465<br />

Zhu, Yanni 1041<br />

Zhu, Yong 120<br />

Zhuo, Luting 1651<br />

Zibbell, Jon E. 1839<br />

Zibert, Andree 684, 2103<br />

Ziegert, Szilvia 1002<br />

Zignego, Anna Linda 1190, 1802,<br />

1844<br />

Zimerman, Michal 679<br />

Zimmer, Sebastian 961<br />

Zimmer, Vincent 76, 463<br />

Zimmermann, Henning W. 828, 831<br />

Zimmermann, Tim 1432<br />

Zinski, Anne 1804<br />

Ziol, Marianne 132, 831<br />

Zipprich, Alexander 1519, 1520<br />

Zitsman, Jonah 2129<br />

Zitteli, Patricia M. 1033<br />

Ziv, Elad 78<br />

Zocco, Maria Assunta 432<br />

Zolfino, Teresa 309<br />

Zoli, Marco 355<br />

Zoller, Heinz M. 1092, 1885<br />

Zollner, Gernot 2068<br />

Zopf, Steffen 1506<br />

Zorzi, Daria 1991


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) AUTHOR INDEX 1365A<br />

Zou, Huaibin 209<br />

Zou, Yan 1583<br />

Zoubek, Miguel Eugenio 828, 1924<br />

Zoulim, Fabien 206, 1847, 2003<br />

Zozaya, Jose M. 2032<br />

Zuberi, Bader F. 1872<br />

Zubkova, Iryna 1546<br />

Zublena, Irène 1162<br />

Zuckerman, Valentina 653<br />

Zucman-Rossi, Jessica 132, 362<br />

Zürcher, Samuel 1575<br />

Zwahlen, Marcel 1575<br />

Zweers, Serge 604


1366A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

Category Index<br />

All numbers refer to the corresponding abstract.<br />

Acute Liver Failure and<br />

Artificial Liver Support<br />

AO1. Acute Liver Failure and<br />

Artificial Liver Support<br />

Arcuate artery resistive ind 1758<br />

Bioluminescence imaging of m 1755<br />

Clinical Features and Outcom 1757<br />

Does Methacetin Used in Brea 1763<br />

Dual Role of Epidermal Growt 1767<br />

ELAD VTL C3A Cells May Impac 1770<br />

End Stage Liver Disease Pati 7<br />

Enhancing hepatocyte functio 8<br />

High CXCR-1 and CXCR-2 Expre 1762<br />

Indeterminate Etiology of Ac 1761<br />

Liver Volumetry determined b 212<br />

Lower fibrinogen level predi 1765<br />

Mesenchymal stem cell using 1760<br />

Nitric Oxide induced hepatic 1773<br />

Noradrenaline dose levels pr 1781<br />

Recovery of severe criticall 1769<br />

Replicative Stress and Cell 1764<br />

Single Center Experience: Hy 1784<br />

Acetaminophen Parent Compoun 1777<br />

Acute Liver Failure and ECMO 1775<br />

Cathepsin L-deficiency enhan 1771<br />

Characterization of Cerebral 214<br />

Clinical Features and Outcom 211<br />

Does weight loss surgery pre 1766<br />

Dual CCR2/CCR5 Antagonist Ce 1756<br />

Early elevation in serum IFN 1778<br />

Expression of Krüppel-like 1779<br />

Fulminant Hepatitis B Acute 1759<br />

Heat Stroke Leading to Acute 1768<br />

Hemophagocytic lymphohistioc 1783<br />

Histology predicts the need 1782<br />

Leukocyte Cell Derived Chemo 1774<br />

Liver specific organ failure 1780<br />

Safety and Tolerability of O 213<br />

Serial Procalcitonin Measure 1776<br />

The Receptor Interacting Pro 1772<br />

Biliary Physiology,<br />

Transport, Cholangiocyte<br />

Biology, and<br />

Experimental Cholestasis<br />

BO1. Transport, Bilirubin,<br />

Cholesterol, Lipids, and Bile Salts<br />

Bile acid induced cholestati 805<br />

Higher order metabolites of 808<br />

NF-kB is critically importan 809<br />

Proteomics analysis of rat c 806<br />

TGR5 agonists improve pancre 807<br />

The Ileal Bile Acid Transpor 810<br />

A Novel Mitochondrial Pathwa 804<br />

Administration of an Ileal A 19<br />

LKB1 regulates hepatocellula 20<br />

BO2. Cholangiocyte Biology<br />

Inhibition of Adenylyl Cycla 811<br />

Natural Killer T cells aggra 819<br />

Nlrp3 inflammasome activatio 21<br />

Pre-treatment with TIMP-3 pr 818<br />

Taurocholate Activates YAP v 22<br />

YAP Reduces Liver Parenchyma 813<br />

Cholestatic-induced biliary 812<br />

Glutathione depletion is cri 814<br />

Intraductal papillary neopla 820<br />

Knockout of the Secretin/Sec 817<br />

The bile acid Ursodeoxycholi 815<br />

The bile-acid receptor TGR5 816<br />

BO3. Experimental Cholestasis<br />

Anti-inflammatory and Antifi 198<br />

Keratin 23 represents a nove 831<br />

Loss of Caspase 8 in liver p 828<br />

RIP3-dependent signalling co 832<br />

Role of Cellular Inhibitor o 825<br />

Tetrahydroxylated Bile Acids 824<br />

CD39/ENTPD1 is protective in 826<br />

FXR-β-catenin complex and b 24<br />

Feedback regulation of autot 827<br />

Gut microbiota contributes t 833<br />

Hepatic CD4+ lymphocyte resp 823<br />

Knockdown of the melatonin r 835<br />

Knockdown of α7-Nicotinic R 830<br />

NHERF-1 assembles macromolec 829<br />

Prolonged usage of H1 or H2 822<br />

SRT1720 protects against cho 23<br />

The secretin receptor antago 834<br />

Wnt/β-catenin cooperates wi 821<br />

Cell and Molecular<br />

Biology<br />

CO1. Cell Structure and Function<br />

Acute alcohol binge causes s 334<br />

CD45+ fraction in murine adi 325<br />

Cataplerosis of α ketogluta 330<br />

Cross-talk between autophagy 326<br />

Genetic Induction of Metabol 331<br />

Hepatocyte Isolation from Li 332<br />

Histone 3 lysine 9 trimethyl 327<br />

Interaction with collagen ma 329<br />

Small Anti-Warburg Molecules 319<br />

UPR Regulation of Autophagy 320<br />

Annexin A6 is necessary for 64<br />

Characterization of Microtub 333<br />

Hepatocyte surface localizat 321<br />

Longitudinal monitoring of h 323<br />

Loss of adherens junctions p 255<br />

Mouse CD11b+Kupffer cells re 324<br />

Netrin-1 Promotes Liver Canc 328<br />

Tissue-Specific Subcellular 322<br />

CO2. Signal Transduction and<br />

Nuclear Receptors<br />

Akt-mediated FoxO1 inhibitio 65<br />

Aspartate β-hydroxylase mod 256<br />

Cerium oxide nanoparticle tr 659<br />

Cyclic AMP increases canalic 661<br />

Effects of Targeting Wnt/β- 673<br />

FGF21 facilitates normal liv 649<br />

HMGB1 interacts with AIM2 to 656<br />

Hedgehog regulates Yes-assoc 66<br />

Hepatic XBP1 deletion promot 257<br />

Hyperammonemia activates a l 668<br />

Hyperammonemia impairs skele 671<br />

IRAKM-Mincle axis contribute 675<br />

Lysophosphatidic acid recept 653<br />

MYCN is A Novel Biomarker fo 657<br />

Metabolomics-Targeted Pertur 672<br />

Orchestration of BH3-only pr 669<br />

RA and butyrate induce liver 674<br />

Regulation of the Hepatic Dr 655<br />

Restoration of Hepatic Physi 664<br />

Skeletal muscle hyperammonem 667<br />

Skeletal muscle mitochondria 651<br />

Thr505 RelA phosphorylation 652<br />

Circulating extracellular ve 109<br />

Effects of Ursodeoxycholic-a 662<br />

Farnesoid X Receptor (FXR) A 658<br />

HNF4α reprogramming during 650<br />

Hepatitis B virus X protein 654<br />

Involvement of Endoplasmic R 663<br />

NK cells from homozygous NLG 660<br />

Organopollutant Exposures Fu 666<br />

Sustained Inhibition of Live 670<br />

The regulation of liver volu 665<br />

CO3. Gene Expression and Therapy<br />

A proof of concept nonclinic 505<br />

Chenodeoxycholic acid induce 520<br />

Chimeric hepatitis E virus-l 502<br />

Defective mTOR-TFEB signalin 501<br />

Dynamic shift of microbiota 63<br />

Evaluation of Glyoxylate and 510<br />

Identification of FGF2 as a 517<br />

Impaired skeletal muscle mit 518<br />

Is Disulfiram a Potential Th 516<br />

MicroRNA-192 regulates the e 519<br />

Molecular heterogeneity in m 253<br />

Organic Cation Transporter-1 506<br />

Shifting gut microbiota and 508<br />

Stabilin-mediated endocytosi 507<br />

The Roles of Kras Isoforms i 504<br />

Therapeutic Potential in Tox 515<br />

YAP Promotes Cyclin D1 Expre 514<br />

Cell-Type-Specific Expressio 511<br />

Disruption of the HNF1α bin 492<br />

High-throughput sequencing t 503<br />

Identifying novel therapeuti 500<br />

In vivo distribution of exos 512<br />

Integrated proteomics analys 513<br />

Methylation-associated silen 498<br />

MicroRNA-122 (miR-122) Regul 493<br />

Mixed hepatocellular-cholang 254<br />

Molecular mechanisms underly 499<br />

Niclosamide ethanolamine inh 497<br />

Overexpression Screening to 494


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) CATEGORY INDEX 1367A<br />

Peretinoin, an Acyclic Retin 495<br />

Pre-operative plasma glypica 509<br />

Therapeutic overexpression o 496<br />

CO4. Stem Cell Biology<br />

CD34+ Liver Cancer Stem Cell 685<br />

Decellularized Spleen Matrix 698<br />

Detection of putative cancer 695<br />

Disruption of TGF-β-regulat 682<br />

Ex vivo-expansion of human C 688<br />

Expression pattern of the st 687<br />

FAP disease modelling using 684<br />

Interleukin-17 mediates live 693<br />

Niche specific insulin signa 691<br />

Sonic Hedgehog-Containing Ex 692<br />

TGF-β inhibits lncRNA H19 e 686<br />

Wnt/beta-catenin signaling a 699<br />

Attenuation of liver fibrosi 689<br />

Bone Marrow Mesenchymal Stem 690<br />

Clonal analysis of the human 101<br />

DNMT1 is essential for effec 677<br />

Growth Factor Mobilized CD34 694<br />

Hepatic maturation of induce 676<br />

Hepatocyte plasticity underl 680<br />

Human Induced Pluripotent St 258<br />

Identification of hepatic st 681<br />

Identification of mechanism 697<br />

Microbial-Derived Cues Drive 679<br />

Notch and Wnt control ductul 678<br />

Self-renewing diploid Axin2+ 683<br />

TSG-6 enhances autophagy in 696<br />

Health Services Research<br />

DO1. Health Care Delivery/Access/<br />

Quality<br />

Charges for Patients Hospita 533<br />

Communication, sequence and 529<br />

Development of a Model to Pr 14<br />

Did improvement of access to 524<br />

Diminished Health-Related Qu 522<br />

Evacuation after the Fukushi 551<br />

Hepatitis C Disease Burden i 104<br />

In Patients with Non-alcohol 575<br />

Learning the Liver: The Pilo 544<br />

National Estimates of Health 560<br />

Novel Algorithm to Predict H 580<br />

Outcome Indicators In Primar 571<br />

Outreach invitations improve 548<br />

Predicting survival in liver 555<br />

Rates of liver fibrosis asse 561<br />

Redesigning Outpatient Care 538<br />

Reduced Work Productivity (W 18<br />

Reducing Wait Times for Radi 542<br />

Successful management of pre 567<br />

The Prevalence and Mortality 534<br />

Thirty-day Readmission Follo 531<br />

Thirty-day Readmission for I 550<br />

Timing of Hepatitis B and C 543<br />

Wide variations in Albumin U 570<br />

Worse Outcomes Among Uninsur 563<br />

A simple educational 13<br />

A New Approach to Patient Ce 535<br />

A Prospective Study of a Pro 17<br />

A Regional Multidisciplinary 521<br />

A qualitative assessment of 547<br />

Abdominal Ultrasonography Gu 554<br />

Age, race and insurance stat 539<br />

Characteristics of Incident 562<br />

Comparable hepatitis C treat 541<br />

Deleterious effects of cirrh 566<br />

Depression and Alcohol Misus 525<br />

Disparities in the Quality o 553<br />

Distance to Transplant Cente 565<br />

Effect of HCV treatment outc 15<br />

Engagement in care of high r 549<br />

Epidemiology and survival in 532<br />

Epidemiology, treatment-patt 559<br />

Evaluation of DAA (direct-ac 527<br />

External Validity of Trials 574<br />

Geriatrics and cirrhosis: ch 528<br />

Heat Mapping Incident PBC Pa 546<br />

Hepatitis B Testing Prior to 540<br />

Hepatitis C Virus Sustained 573<br />

Hepatitis C care and treatme 568<br />

High Hepatitis E seroprevale 523<br />

High Rates of Hepatitis C Vi 545<br />

Interventions to optimize re 577<br />

Knowledge about hepatitis B 572<br />

Living Liver Donors: Dispari 556<br />

Modernizing the Child-Turcot 16<br />

Palliative Care Services are 557<br />

Physician versus Patient Per 537<br />

Primary Care Physician Persp 576<br />

QI in Practice: Reducing Rea 536<br />

Reduction of 30 day Readmiss 530<br />

Screening for poor prognosis 564<br />

Suboptimal Quality Of Preven 552<br />

Successful Implementation of 579<br />

The Burden of Alcohol Liver 526<br />

The loss of skeletal muscle 569<br />

Upper Limit Normal of Alanin 578<br />

“Normal” Serum ALT/AST L 558<br />

DO2. Cost-Effectiveness and<br />

Economics of Care<br />

Cost-effectiveness of Daclat 1461<br />

Hospital Based Initiative to 1452<br />

Superiority and Durability o 1460<br />

The Attention-to-Burden Inde 1451<br />

The Cost of Making Hepatitis 151<br />

Treating Hepatitis C in the 152<br />

A Trial-based Model of Liver 155<br />

A probabilistic decision mod 1455<br />

Bariatric surgery for the tr 154<br />

Cost-effectiveness of Sofosb 1453<br />

Hospitalisation for primary 1457<br />

Impact of Comorbidity on Inp 1459<br />

Impact of Daclatasvir-Sofosb 1456<br />

Living Donor Liver Transplan 1454<br />

Obeticholic Acid (OCA) has I 1458<br />

The cost-effectiveness of om 1462<br />

DO3. Health Economics and Cost<br />

Effectiveness of Viral Hepatitis<br />

Cost-effectiveness analysis 1536<br />

Cost-effectiveness analysis 1537<br />

Costs per Diagnosed Liver Di 156<br />

Hepatitis C (HCV) Infection 153<br />

Inappropriate Laboratory Uti 1535<br />

Rethinking pricing policy of 1539<br />

The Value of Cure Associated 1538<br />

All-oral direct acting antiv 1534<br />

Hepatitis B<br />

EO1. Virology, Pathogenesis, and<br />

Immunology<br />

An impact of cellular core-f 1642<br />

CRISPR/Cas9 “double”-nickase 1629<br />

Comprehensive analysis of th 1648<br />

Defective humoral immunity i 1663<br />

Deletion of APOBEC3B is Asso 1651<br />

Differential Serum Cytokine 1668<br />

Endogenous Antigen Presentat 1640<br />

Factors associated with decr 1677<br />

GORASP2 regulates HCV and HB 1630<br />

Genetic association study fa 1681<br />

Genome wide miRNA: mRNA Inte 1659<br />

Genome-wide scan of miRNA po 1649<br />

HBcrAg during nucleos(t)ide 1653<br />

Hepatitis B Virus Surface Pr 1656<br />

Humic matter inhibits HBV-in 1664<br />

IFN-α-mediated Base Excisio 1684<br />

Initial sites of hepadnaviru 1641<br />

Large Hepatitis Delta Antige 1665<br />

Long-term follow-up study of 1674<br />

Massive Intrahepatic Product 1646<br />

Modeling Hepatits B Virus in 1680<br />

Noninvasive markers of liver 1676<br />

Osteopontin is involved in c 163<br />

Precore and Basal Core Promo 1667<br />

Presence of naïve-like CD8+ 1666<br />

Proline142 deletion in the H 1639<br />

The mechanism of interferon 1644<br />

The Combination of KIR and H 1633<br />

The HBx-DLEU2 lncRNA complex 165<br />

The HLA-DPA1 rs3077 TT polym 1638<br />

The expansion of CD11b-CD27- 1647<br />

Viral Expression and Molecul 166<br />

A Novel Entecavir Resistance 1655<br />

Analysis of the effect on HB 1627<br />

Characterisation of the immu 167<br />

Comparison of detection rate 1675<br />

Difference of intracellular 1645<br />

Downregulation of TLR2 on CD 1683<br />

Dynamic Analysis Of Cellular 1672<br />

Effect of a novel synthetic 1652<br />

Effects of HLA-DPB1 genotype 1632<br />

Free episomal and integrated 1654<br />

Functional restoration of CD 1660<br />

Generation of human peripher 1635<br />

Genetic Variation in Vitamin 1631<br />

Geranylgeranylacetone exerts 1671<br />

Hepatitis B virus X protein 168<br />

Hepatitis B virus down regul 1685<br />

Hepatitis delta virus-induce 1650<br />

Human cytotoxic T lymphocyte 1670<br />

Human induced pluripotent st 164<br />

IFN-λ3 induction by nucleot 1673<br />

Impact of toll-like receptor 1658<br />

Indoleamine-2, 3-dioxygenase 1643<br />

Inhibition of epigenetic reg 1628<br />

Knockdown of NTCP reduces su 1661<br />

NFATc3 exerts anti-viral and 1634<br />

NK cell function is restored 1626<br />

NK cells play important role 1636


1368A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

No Detectable Resistance to 1678<br />

Novel robust in vitro hepati 1679<br />

Possible involvement of hepa 1669<br />

Serial Changes of Th1 and Th 1682<br />

The Association of vitamin D 1662<br />

The activating receptor NKP4 1657<br />

The role of HBV quasispecies 1686<br />

Unique Liver Transcriptomic 1637<br />

EO2. Diagnostics, Epidemiology,<br />

Prevention, and Natural History<br />

Aldehyde dehydrogenases-2 rs 1580<br />

Application of Highly Sensit 1573<br />

Assessing the impact of hepa 1546<br />

Baseline quantitative hepati 1604<br />

Clinical Features of Chronic 1587<br />

Clinical outcome of long-ter 1581<br />

Clinical scoring system for 1616<br />

Comparative Study for Anti-H 1599<br />

Conconrdance between transie 1621<br />

Different DNA methylation pa 1553<br />

Evaluation of HBV infection 1622<br />

Four years real-life experie 1544<br />

Genotype-Specific Prevalence 1559<br />

HBV-related cirrhosis in the 1578<br />

Hepatitis B Core-Related Ant 1585<br />

Hepatitis B Core-Related Ant 1586<br />

Hepatitis B virus infection 1543<br />

Hepcidin: A new clinical fac 1620<br />

High Rates of Surface Antige 1549<br />

Individualizing Hepatitis B 1560<br />

Liver fibrosis progression, 1548<br />

Long term clinical outcome i 1558<br />

Lower Incidence of Hepatocel 1561<br />

Mapping HBsAg epitope profil 1551<br />

Nationwide Prospective Surve 1577<br />

Objective Determination of H 121<br />

PreS mutations analyzed by d 1565<br />

Prevalence and Molecular Vir 1566<br />

Prevalence and serological f 1555<br />

Prevalence of Hepatitis B vi 1596<br />

Relationship between hepatoc 126<br />

Risk assessment of hepatocel 1605<br />

Seroprevalence and Clinical 1598<br />

Targetting HNF4α via Activa 1562<br />

The Predictive Values of Hep 246<br />

The Usefulness of CLIF-SOFA 1609<br />

The role of hepatitis B core 1550<br />

Utility of quantitative hepa 1576<br />

Angiopoetin-like protein 2 i 1564<br />

Anti-platelet therapy reduce 1610<br />

Assessment Of Treatment Resp 1624<br />

Assessment of liver steatosi 1582<br />

Barriers to obtaining approp 1608<br />

Baseline Hepatitis B core re 1572<br />

Biomarkers of Fibrosis in Lo 1607<br />

Changes in hepatitis B virus 1584<br />

Changes of Hepatitis B surfa 1615<br />

Characterization of Nucleic 1603<br />

Clinical Outcomes of Patient 1556<br />

Collaborative Public Health 1574<br />

Comparing antiviral efficacy 1542<br />

Comparison between spontaneo 1552<br />

Detection of genetic and epi 1554<br />

Does Cesarean section and fo 1619<br />

Elevated serum levels of WFA 1600<br />

End-of-therapy Serum Gradien 1547<br />

Evaluation of Chronic Hepati 1606<br />

Evaluation of Serum Cell Dea 1611<br />

Factors Associated with the 1545<br />

HBsAg transferring from moth 1579<br />

Health care disparity in del 1569<br />

Hepatic B flares in rheumato 1589<br />

Hepatitis B viral load in dr 1575<br />

Hepatocellular Carcinoma (HC 125<br />

Hepatocellular Carcinoma Inc 1540<br />

High-throughput and ultrasen 1614<br />

Identification of a non-inva 1601<br />

Inadequate Treatment of Chro 1625<br />

Increased Viral Diversity is 1557<br />

Influence of the ethnic stat 1613<br />

Interleukin-2 Receptor and T 1602<br />

Is booster beneficial after 1623<br />

Is there any rule for antivi 1541<br />

Kinetics of Serum Hepatitis 1567<br />

Length of antiviral therapy 122<br />

Longitudinal analysis of com 1595<br />

Performance Assessment of Co 1618<br />

Performance of the new cobas 1597<br />

Postpartum hepatic flare aft 1617<br />

Prediction of Clinical Outco 1591<br />

Prevention of perinatal tran 1583<br />

Quantitative Hepatitis B sur 1571<br />

Redefining the Natural Histo 1563<br />

Serum Alanine Aminotransfera 123<br />

Serum HBV RNA is an early ma 1594<br />

Serum Quantitative Levels of 1593<br />

Statin and the risk of hepat 1612<br />

The PAGE-B Score Accurately 1590<br />

The PAGE-B Score Stratifies 1568<br />

The Role of Surface Antibody 1592<br />

Ultrasound predicts steatosi 1570<br />

What is predicting post-preg 1588<br />

EO3. Therapeutics: New and<br />

Approved<br />

“Real Life” outcomes for Ten 2047<br />

A Multicenter, Prospective, 2048<br />

A Randomized Prospective Ope 1995<br />

A randomized study of standa 1997<br />

A study to evaluate the dyna 2028<br />

Association between Pre-S2 D 2040<br />

Association between ALT flar 2059<br />

Association of Interferon-ga 2036<br />

Baseline HBsAg titer allows 2045<br />

Biochemical, Virologic, and 243<br />

Biologic Responses to GS-477 2052<br />

Characterization of Hepatiti 2026<br />

Comparative Effectiveness of 244<br />

Comparison of the efficacy o 2055<br />

Decreasing risk of hepatocel 2012<br />

Discontinuation of Antiviral 2049<br />

Discontinuation of Effective 2019<br />

Dynamics of hepatitis b viru 2066<br />

Early Decline in Serum HBsAg 2014<br />

Early prediction of response 2001<br />

Efficacy and Safety of Pegin 2023<br />

Encapsidation and secretion 2064<br />

Final Results of Peginterfer 2002<br />

GalNAc-siRNA conjugate ALN-H 36<br />

Inhibition of Hepatitis B Vi 33<br />

Intrahepatic IP-10 expressio 2046<br />

Liver Gene Expression Profil 2027<br />

Long Term Efficacy and Safet 2004<br />

Long term follow up of chron 2017<br />

Longitudinal PAGE-B Scores P 2003<br />

Monthly dosing of ARC-520 in 2022<br />

Nucleos(t)ide Analogue Disco 1998<br />

Nucleos(t)ide Analogues agai 242<br />

Plasma MicroRNA Levels are A 2005<br />

Prolongation of interferon t 2033<br />

Reduction of HBsAg by peg-in 2029<br />

Reductions in cccDNA under N 32<br />

Safety and Efficacy of GS-47 2015<br />

Serologic and Virologic Chan 2043<br />

Serum HBV RNA is an Early Pr 245<br />

Serum hepatitis B core-relat 2063<br />

Tenofovir Disoproxil Fumarat 209<br />

The New Cyclophilin Inhibito 2035<br />

The baseline combination of 2044<br />

The incidence and predictors 2010<br />

The relationship between pat 2057<br />

Treatment and HBeAg-status D 241<br />

Urinary β2-microglobulin fo 1999<br />

Week 24 HBsAg levels predict 2013<br />

Whole-Genome Sequencing of P 2037<br />

A randomized trial evaluatin 2038<br />

Antiviral Activity of Tenofo 2021<br />

Antiviral treatment for chro 2060<br />

Bone toxicity during long-te 2053<br />

Characterization of HBsAg lo 2032<br />

Combined hepatitis B core-re 2020<br />

Descriptive Comparison of Ef 2025<br />

Development and Recurrence o 2065<br />

Development of a Direct RNA 2062<br />

Dual-gRNAs and gRNA-microRNA 34<br />

Entecavir and Tenofovir More 2016<br />

Excellent outcomes of hepati 2000<br />

Functional Activation of NK 2009<br />

HBV Does Not Modulate the Tr 35<br />

Impact of pre-treatment plat 2031<br />

Is efficacy of tenofovir in 2042<br />

Long-Term Clinical And Serol 2008<br />

Long-term Renal Safety of Ad 2056<br />

NGAL can act as a renal safe 2054<br />

Prediction of sustained off- 2050<br />

Reduction of hepatocellular 1996<br />

Safety, antiviral activity, 2058<br />

Sequential therapy with Peg- 2030<br />

Study on post-treatment rela 2061<br />

TKM-HBV, a Novel RNA Interfe 2007<br />

TKM-HBV, a Novel RNA Interfe 2034<br />

Tenofovir-associated Fanconi 2006<br />

The Efficacy And Safety Of T 2024<br />

The Relationship Between Tel 2051<br />

The combination of IFN-Lambd 2011<br />

The impact of timely initial 2018<br />

Update on the safety and eff 31<br />

Hepatitis C<br />

FO1. Virology, Pathogenesis, and<br />

Immunology<br />

A methyl donor, S-adenosylme 1022<br />

Alcohol increases the produc 30<br />

Altered frequencies and func 1007<br />

Analysis of genetic variatio 1028<br />

Daclatasvir and Sofosbuvir i 217<br />

Different HCV genotypes have 999<br />

Elevation of Galectin-9 is a 26<br />

Elucidating Mechanisms Under 1017<br />

Evaluatiin of protective and 992<br />

Evolution Of Resistance Asso 220<br />

Expression of IFNλ4 in live 1018


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) CATEGORY INDEX 1369A<br />

Fast and deep early HCV-RNA 222<br />

HCV RNA and HCV core antigen 1002<br />

HIV and HCV co-infection in 25<br />

Individualized DAA treatment 982<br />

Interferon lambda 3 (IL28B) 980<br />

Iron chelation restores mito 995<br />

Lipid-Free apolipoprotein E 1019<br />

Liver pDCs from HCV-Infected 988<br />

Long noncoding RNAs regulate 991<br />

MicroRNA 17 Host Gene Protei 1011<br />

Persistence of HCV-specific 1031<br />

Polymorphisms in irisin-prod 221<br />

Pre-treatment levels of miR- 1023<br />

Rapid normalization of antiv 28<br />

Role of very-low-density lip 997<br />

Selenoprotein P regulates th 990<br />

Serial change of resistant a 1027<br />

The Antibody response to hep 1033<br />

The Emergence of NS5B Resist 219<br />

The impact of HLA-DQs and IF 986<br />

The roles of unphosphorylate 1025<br />

The tumor suppressor IQGAP2 1015<br />

Transcriptomic analysis of C 1029<br />

Type III interferon response 1006<br />

Variability of the hepatitis 1021<br />

Very low prevalence of Hepat 1032<br />

De novo formation and matura 27<br />

A miR-122/miR-224-based mode 998<br />

Activation of Hepatic Stella 29<br />

Analysis of AL-516, a Novel 993<br />

Cell-specific peripheral inn 996<br />

Differential microRNA expres 1014<br />

Downregulation of let-7 fami 1005<br />

Effects of Resistance Mutati 985<br />

Essential Factors for Recons 1003<br />

Establishment and Characteri 1000<br />

Establishment of A Novel Hep 989<br />

Hepatitis C Genotype 4r Resi 1009<br />

Hepatitis C virus facilitate 984<br />

Interferon-based treatment a 1008<br />

Kidney involvement in HCV Ch 983<br />

LECT2 specifically induced b 1010<br />

MicroRNA-130a inhibit HCV pr 981<br />

Novel effect of NS5A inhibit 1013<br />

PD-1 blockade enhances cytop 1001<br />

Rapid Decline In Interferon 218<br />

Regulation of Hepatitis C Vi 994<br />

Single-strand RNA-induced mo 1016<br />

Substitution in amino acid 7 1020<br />

The FGL2:FcγRIIB Immunosupp 1012<br />

The Hepatitis C Virus modula 987<br />

The effect of sofosbuvir-bas 1030<br />

The expression of Immune-miR 1026<br />

Tim-3 Inhibits Monocyte Func 1004<br />

FO2. Diagnostics, Epidemiology, and<br />

Natural History<br />

MICA SNP, but not DEPDC5, PN 1824<br />

Hepatitis E in haematologica 1803<br />

A Highly Specific and Sensit 85<br />

A New Simple Quantitative Li 1849<br />

Acute HCV recurrence after l 1889<br />

Acute hepatitis and reactiva 1882<br />

Age and Risk Factor-based Se 1823<br />

Application of a Novel Hepat 1785<br />

Benefits of HCV eradication 1808<br />

Can HCV Antigen Testing Repl 1826<br />

Cannabinoid receptor 2 (CB2) 1837<br />

Cascade of Care for Hepatiti 1786<br />

Cascade of Care of HCV & HIV 1891<br />

Comparison of performance ch 1817<br />

Correlates of Successful HCV 1892<br />

Differential Progression to 1848<br />

Dynamic change of α-fetopro 1873<br />

Effectiveness and cost-effec 1845<br />

Evaluation of liver and sple 1862<br />

Evaluation of liver and sple 1879<br />

Evaluation of the role of he 1890<br />

Expanded eligibility to HCV 1861<br />

Fibrosis Regression After He 1876<br />

Healthcare Utilization and C 1883<br />

Hepatitis C Seropositivity i 1822<br />

Hepatitis C Virus (HCV) Mort 86<br />

Hepatitis C Virus (HCV) Prev 1859<br />

Hepatitis C Virus (HCV) Prev 1818<br />

Hepatitis C and Renal Diseas 1867<br />

Hepatocellular carcinoma (HC 1847<br />

Highly accurate HCV genotypi 1812<br />

Illegal drug use disclosure 1857<br />

Incidence and Predictors of 90<br />

Increasing Prevalence of Cir 88<br />

Investigation into the high 1865<br />

Linkage to Care and Treatmen 1874<br />

Low HCV Treatment for Chroni 1816<br />

Male gender, genotype 3, pre 1844<br />

Mathematical model for early 1821<br />

Modeling the probability of 1854<br />

Non-invasive assessment of l 1834<br />

Performance of HCV Ag quanti 1833<br />

Policy Implications of Dispr 1813<br />

Pragmatic Non-invasive test 1856<br />

Prevalence and Risk-Factors 1853<br />

Prevalence of diabetes and a 1877<br />

Prevalence of hepatitis B co 1850<br />

Prevalence of viral hepatiti 1796<br />

Progression to liver cirrhos 1806<br />

Rate and Predictors of Serum 1881<br />

Repeated biopsies within 23 1878<br />

Risk of hepatitis C virus in 1793<br />

Safety and Efficacy of Inter 1885<br />

Screening Urban Baby Boomers 1835<br />

Self-reported Alcohol Use at 1880<br />

Stability and prevalence of 1832<br />

The Utility and Limitations 1799<br />

The hepatitis C epidemic amo 1830<br />

The influence of Hepatitis C 1869<br />

The modeled impact of improv 1842<br />

The potential prevention imp 1794<br />

Times they are a changing: H 1871<br />

Treatment prioritization acc 1802<br />

Trends in end-stage liver di 1868<br />

Wisteria floribunda agglutin 1851<br />

Ability of EMR-based fibrosi 1791<br />

Access To Reimbursement Of I 1886<br />

An Integrated Linkage to Car 1810<br />

Benefits of Antiviral Treatm 1828<br />

Can Hepatitis C be eliminate 1872<br />

Cause of death in HCV patien 1789<br />

Characteristics of Patients 1792<br />

Comparative analysis of onli 1827<br />

Comparison of On-Treatment H 1860<br />

Development of End Stage Liv 1790<br />

Distribution of pre-existing 1795<br />

Does pregnancy protect again 1884<br />

Evaluation of both Fucosylat 1809<br />

Factors Associated with Hepa 1839<br />

Fibrosis Progression in Pati 1841<br />

From Care to a Cure: Improvi 1801<br />

HCV in Semen of HIV-infected 1852<br />

Hepatitis C Birth-Cohort Tes 1787<br />

Hepatitis C Eradication with 1866<br />

Hepatitis C RNA assay differ 1797<br />

Hepatitis C Virus (HCV) Erad 1825<br />

Hepatitis C Virus (HCV) Geno 1798<br />

Hepatitis C Virus Genotype A 1820<br />

Hepatitis C prevalence, dete 1846<br />

Heterogeneous IL28B/IFNL4 di 1855<br />

Hyperbilirubinemia in HIV-HC 1840<br />

Impact of Hepatitis C (HCV) 1807<br />

Impact of comorbidities on t 1863<br />

Improvements in the Hepatiti 1870<br />

Increasing Hepatitis C Virus 1811<br />

Linkage to Care Outcomes for 1804<br />

Liver fibrosis stage and com 1800<br />

Natural History of Decompens 1815<br />

Non-Invasive Fibrosis Scores 1843<br />

Patients with Hepatitis C (H 1814<br />

Perceived risk and HCV knowl 1864<br />

Positive Impact of Point of 1831<br />

Predictive value of APRI and 1887<br />

Racial Differences in Socioe 1805<br />

Rapid, sensitive, and accura 1858<br />

Rectal Shedding of HCV in HC 89<br />

Regression of Advanced Fibro 108<br />

Serum levels of Wisteria flo 1838<br />

The Association of Sustained 87<br />

The Power of the Law: Improv 1875<br />

The clinical profiles of Jap 1819<br />

The epidemiology of hepatiti 1829<br />

The use of hepatitis C core 1888<br />

Treatment With Statins Reduc 1788<br />

FO3. Therapeutics: Preclinical and<br />

Early Clinical Development<br />

A Single Dose of RG-101, a G 208<br />

Miravirsen Dosing in Chronic 2255<br />

Pharmacokinetics, Safety and 2260<br />

Pharmacokinetics, Safety, an 2259<br />

SB 9200 Shows Antiviral Acti 2262<br />

Safety, Efficacy and Tolerab 2256<br />

Short-term safety and pharma 2261<br />

The effect of microRNA let-7 2264<br />

Treatment with the Anti-miRN 2263<br />

Discovery of AT-337 and AT-3 2266<br />

Drug-Drug Interaction Studie 2265<br />

EDP-239, a novel HCV NS5A in 2257<br />

The impact of α-fetoprotein 2267<br />

Toxicology evaluation of TT- 2258<br />

FO4. Therapeutics: New Agents (not<br />

approved, phase 2-3)<br />

98%–100% SVR4 in HCV Genot 41<br />

A Randomized Controlled Tria 205<br />

Absence of significant drug- 723<br />

An Integrated Analysis of 40 42<br />

An integrated safety analysi 716<br />

Analysis of HCV Genotype 1 V 705<br />

Baseline HCV NS5A resistance 709<br />

Characterization of HCV Resi 718<br />

Daclatasvir exposure alone d 728<br />

Daclatasvir exposure does no 720<br />

Daclatasvir in combination w 252<br />

Daclatasvir plus sofosbuvir 206<br />

Drug-drug interactions betwe 710<br />

Efficacy and Safety of Co-Fo 708<br />

Efficacy and Safety of Ombit 714


1370A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

Efficacy of Grazoprevir (GZR 715<br />

High Efficacy of Grazoprevir 251<br />

High Efficacy of Grazoprevir 210<br />

High Efficacy of the Combina 701<br />

High SVR4 Rates Achieved Wit 250<br />

High SVR4 Rates Achieved Wit 248<br />

Improvement in liver disease 706<br />

Integrated Safety Analysis o 726<br />

No Clinically Meaningful Pha 725<br />

No Pharmacokinetic Interacti 731<br />

No Pharmacokinetic Interacti 730<br />

Ombitasvir/Paritaprevir/r an 722<br />

Predictors of Response to Gr 700<br />

Resistance Analysis of Treat 713<br />

SVR12 results from the Phase 39<br />

Safety and Tolerability of G 712<br />

Safety and efficacy of dacla 37<br />

Short-duration therapy with 702<br />

Sofosbuvir/GS-5816 Fixed Dos 249<br />

The Combination of Grazoprev 703<br />

A Pilot Evaluation of Twice 724<br />

C-EDGE CO-STAR: Efficacy of 40<br />

C-EDGE Co-Infection: Impact 729<br />

C-EDGE TN: Impact Of 12-Week 717<br />

Comparative Efficacy and Tol 711<br />

Efficacy, Safety And Pharmac 707<br />

High Rates of SVR in Treatme 247<br />

Pharmacokinetics of Coadmini 719<br />

Pharmacokinetics of Narlapre 721<br />

Prevention of allograft HCV 704<br />

Projected Long-Term Impact o 727<br />

Sofosbuvir/GS-5816+GS-9857 f 38<br />

FO5. Therapeutics: Approved Agents<br />

(THE SOFGER TRIAL) Sofosbuvi 1078<br />

IFNL4 ss469415590 Polymorphi 1149<br />

Adherence to All-Oral HCV Tr 1063<br />

Asian Patients with Chronic 1080<br />

Asian Patients with Genotype 1105<br />

Association of polymorphisms 1055<br />

Clinical Management of Ribav 1067<br />

Co-morbidities and co-medica 1140<br />

Combination of IFNL4 polymor 1153<br />

Combination therapy with dac 1125<br />

Comparison of 8 vs. 12 weeks 1036<br />

Demographic Predictors of Di 1048<br />

Early and persistent increas 1070<br />

Early dose adjustment of RBV 1162<br />

Effect of Chronic Kidney Dis 1088<br />

Effectiveness of Ledipasvir/ 93<br />

Effectiveness of Treatment o 1155<br />

Efficacy and Adverse Event P 1119<br />

Efficacy and Safety of Simep 1118<br />

Efficacy and Safety of Sofos 1165<br />

Efficacy and safety of IFN-f 1092<br />

Efficacy of Ledipasvir plus 1146<br />

Evaluation of the efficacy a 1191<br />

Global prevalence of pre-exi 1044<br />

Hepatic Decompensation and S 1185<br />

Hepatitis C treatment with S 1102<br />

Hepatitis C virus NS5A L31V 1082<br />

Hepatocellular carcinoma dev 1056<br />

High efficacy of a 12-week s 1179<br />

High efficacy of ledipasvir/ 1049<br />

Highly Successful Retreatmen 92<br />

IFN-λ inhibits miR-122 tran 1096<br />

IFN-free DAA regimens improv 1204<br />

Impact of baseline albumin l 1040<br />

Lower 25-OH Vitamin D Levels 1109<br />

NS3 Q80K Polymorphism in Vir 1127<br />

Ombitasvir/Paritaprevir/r an 1084<br />

Outcomes of ribavirin free r 1174<br />

Patient Decision-Making: The 1114<br />

Patient Expectations of the 1152<br />

Patient Reasons for Initiati 1139<br />

Pharmacokinetic Analyses of 1133<br />

Pre-Existing Co-Morbidities 1200<br />

Prediction for poor virologi 1062<br />

Prevalence of Pre-Treatment 91<br />

Quality-Adjusted Cost of Car 1079<br />

Quantifying and Predicting E 1145<br />

Real Life Treatment Outcomes 1154<br />

Real-world effectiveness and 1176<br />

Real-world effectiveness of 1205<br />

Resistance Analyses of Phase 1168<br />

Retreatment with sofosbuvir 1123<br />

Ribavirin levels at treatmen 1098<br />

Safety and Efficacy of All-O 1126<br />

Safety and Efficacy of Sofos 1057<br />

Safety and Efficacy of Treat 1128<br />

Simeprevir plus sofosbuvir f 1072<br />

Simeprevir, pegylated-interf 1192<br />

Sofosbuvir, Ledipasvir in IB 1075<br />

Sofosbuvir (SOF) and Ledipas 1035<br />

Sofosbuvir + Ledispasvir Com 1101<br />

Sofosbuvir Plus Ribavirin Wi 1094<br />

Sofosbuvir and Ledipasvir/So 1120<br />

Sofosbuvir and Ribavirin for 1083<br />

Sofosbuvir and ledipasvir fo 1081<br />

Sofosbuvir in combination wi 1091<br />

Sofosbuvir plus Ribavirin in 1076<br />

Sofosbuvir-based antiviral t 1158<br />

Sofosbuvir-based treatments 1156<br />

Sofosbuvir-based treatments 1060<br />

Sustained Virologic Response 1043<br />

The TOSCAR Study: Sofosbuvir 1077<br />

The effect of interferon-fre 1047<br />

Treatment Outcomes With 8, 1 94<br />

Treatment of Chronic HCV Gen 1170<br />

Treatment of Chronic Hepatit 1122<br />

Treatment of Hepatitis C Gen 1163<br />

Treatment of Hepatitis C Vir 1186<br />

Virological and Clinical Res 1190<br />

What does Fibroscan® measur 1137<br />

Is ribavirin actuall 1187<br />

ITPA polymorphisms are predi 1193<br />

Interferon Signaling and the 1132<br />

A Meta-Analysis of the Assoc 1129<br />

A comparison of direct seque 1061<br />

Adherence and Discontinuatio 1050<br />

An Actuarial View of the Agi 1069<br />

Approved All-oral Sofosbuvir 1034<br />

Assessment of serum hepatiti 1207<br />

Association between severe r 1138<br />

Awareness Of Hepatitis C Sta 1042<br />

Better Work Productivity and 1198<br />

Cardiac arrhythmia in patien 1194<br />

Clinical Significance of Blo 1203<br />

Clinical- and Economic-perfo 1095<br />

Combination therapies with d 1150<br />

Concomitant use of chemother 1160<br />

Cost-effectiveness of ombita 1143<br />

Daclatasvir plus sofosbuvir 1058<br />

Detecting Drug-Induced Liver 96<br />

Determination of on-treatmen 1103<br />

Differences in NK phenotype 1117<br />

Early response and efficacy 1167<br />

Effect of Different Immunosu 1045<br />

Effect of Proton pump inhibi 1199<br />

Effect of Renal Function on 1066<br />

Effectiveness of 12 or 24 we 1108<br />

Effectiveness of 8 or 12 wee 1046<br />

Efficacy and safety of dacla 1116<br />

Efficacy and safety of ombit 1107<br />

Efficacy of Daclatasvir/Asun 1097<br />

Efficacy of Sofosbuvir and S 1206<br />

Efficacy of combination ther 1157<br />

Efficacy, Change in MELD Sco 1106<br />

Evaluation of an Integrated 1136<br />

Frequency of Renal Impairmen 1099<br />

Frequent Emergences of Rare 1064<br />

HCV genotype 3: Meta-analysi 1134<br />

Health Outcomes and Cost Eff 1089<br />

Health resource usage in per 1173<br />

HepCure: An Innovative web-b 1151<br />

Hepatitis C Cure Could Avoid 1054<br />

Hepatitis C Cure Results in 1202<br />

Hepatitis C and HCC Outcomes 1124<br />

High prevalence of co-morbid 1052<br />

High ribavirin dose and exte 1201<br />

Hope and Hopelessness in HCV 1177<br />

IFN and/or RBV-Free Therapy 1189<br />

Improving liver function and 95<br />

Lifetime risks of liver morb 1087<br />

Lifetime risks of liver morb 1144<br />

Limited Effectiveness of Sof 1164<br />

Long-Term Efficacy of Ombita 1086<br />

Multicenter Experience using 1038<br />

Ombitasvir/Paritaprevir/r an 1161<br />

On-treatment HCV RNA Decline 1130<br />

Patterns of Care Since the A 1141<br />

Potential for drug-drug inte 1093<br />

Predictors of Treatment Adhe 1175<br />

Preliminary Experience Of Di 1148<br />

Preliminary Safety and Effic 1065<br />

Prevalence of drug-drug inte 1131<br />

Prospective Study for The Ef 1100<br />

RUBY-I: Ombitasvir/Paritapre 1039<br />

Rapid drop in serum glucose 1180<br />

Rapid virological response a 1110<br />

Real Life Experience of Dire 1104<br />

Real Life Experience with So 1121<br />

Real World Effectiveness Of 1188<br />

Real-Word Effectiveness of L 1053<br />

Resistance-associated varian 1115<br />

Resistant-associated variant 1197<br />

Ribavirin Levels and Transpo 1196<br />

Safety And Efficacy Of The C 1073<br />

Safety And Efficacy Of Two I 1147<br />

Similar Tolerability and Eff 1182<br />

Sofosbuvir Combination Thera 1184<br />

Sofosbuvir and Ledipasvir fo 1111<br />

Sofosbuvir and Ledipasvir in 1037<br />

Sofosbuvir and Ledipasvir ve 1074<br />

Sofosbuvir in the Treatment 1090<br />

Sofosbuvir is well tolerated 1178<br />

Sustained Virologic Response 1169<br />

Telaprevir in the Treatment 1112<br />

The Significance of Immunoal 1059<br />

The Time and Cost Investment 1172<br />

The healthcare cost burden o 1068<br />

The number needed to treat w 1166<br />

Therapeutic effect of dual o 1159<br />

Transient renal dysfunction 1113<br />

Treatment Outcomes of Vetera 1142<br />

Treatment of recurrent Hepat 1181<br />

Turquoise-III: 12-Week Ribav 1051<br />

Twelve Weeks of Sofosbuvir p 1041<br />

Utility of acoustic radiatio 1195


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) CATEGORY INDEX 1371A<br />

Utilization and Effectivenes 1135<br />

Utilizing claims data to und 1085<br />

Virological efficacy of new 1171<br />

“Real-life”-Experience w 1183<br />

Hepatobiliary Neoplasia<br />

GO1. Experimental<br />

Hepatocarcinogenesis<br />

Alpha fetoprotein harbors a 1977<br />

CD26 (DPP-4) as a molecular 1937<br />

Cholangiocarcinoma cells int 1993<br />

Co-activation of AKT and c-M 1934<br />

Colorectal liver metastasis 1969<br />

Correlation between Incidenc 1991<br />

Deregulated Interplay Betwee 1952<br />

Development of HCC is depend 1971<br />

EVI1 is a target for gene am 1980<br />

Efficacy and mechanism of de 1941<br />

Expression of Transforming G 1962<br />

FGF15 and FGF19 Induce Dispa 1986<br />

Fibrotic liver microenvironm 1970<br />

Glypican 3 (GPC3)-CD81 axis 1979<br />

Glypican 3 contributes to HC 1967<br />

HCV Core Gene Mutations with 229<br />

Hepatic Ruvbl1 is key regula 1963<br />

Hepatocyte-derived osteopont 1940<br />

Hyperinsulinemia, not hyperg 1981<br />

Liver Selective Acetyl-CoA C 1938<br />

Macrophage colony-stimulatin 1956<br />

Mevalonate pathway targets F 1974<br />

MicroRNA profiling of human 1992<br />

MicroRNA-207 controls Prp19 1964<br />

Notch-1 Up-regulated Rat Ova 1951<br />

Oxysterols Promote Malignant 1994<br />

P300 promotes TGF-β stimula 1947<br />

Platelet-Derived Growth Fact 1950<br />

Promoted malignant tumor phe 1968<br />

Promotion of Hepatocarcinoge 1957<br />

Steatosis Induces Wnt/Β-Cat 1945<br />

Survival of hepatocellular c 1960<br />

The Long Noncoding RNA SNHG6 1939<br />

The analyses of oxidative DN 1984<br />

The relationship between hep 1944<br />

Truncated Hepatitis B virus 1946<br />

Voluntary exercise opposes a 230<br />

Baicalein targets liver tumo 1936<br />

Branched chain amino acids p 1982<br />

Chronic alcohol accelerates 1958<br />

Classification of tumors bas 1935<br />

Continuous hepatocyte apopto 1985<br />

Critical role of Hippo casca 1948<br />

Cytotoxic synergy between so 1972<br />

DDB2 loss protects mice from 1988<br />

DNA methylome and cancer-spe 1983<br />

Differential Antitumor Effec 1966<br />

Hepatocyte-specific overexpr 1949<br />

Inactivation of Aldose Reduc 1961<br />

Inflammation contributes to 1989<br />

Inhibition of autophagy reve 1943<br />

Lymphotoxin–β regulated b 1987<br />

Osteopontin is a novel surro 1976<br />

Osteopontin-c isoform promot 1978<br />

PPPDE1 is a deubiquitinase t 1973<br />

PTEN Expression Affects the 1942<br />

Regorafenib inhibits ERK sig 1953<br />

Regulator of G-protein signa 1990<br />

Retrospective cohort study f 1975<br />

Significance of connective t 231<br />

Sitagliptin, a Dipeptidyl Pe 233<br />

Somatic Mutational Analysis 1965<br />

The Dead Box Protein P68 Can 1954<br />

Two-step forward genetic scr 234<br />

Variations of the host genom 1959<br />

p62 is a Key Regulator for I 232<br />

GO2. Diagnostics and Liver Imaging<br />

A mass spectrometry based se 1900<br />

ARID1A mutations may represe 1901<br />

Circulating tumor DNA analys 1895<br />

DNA Methylation Markers for 169<br />

Diagnostic Performance of Co 1897<br />

Efficacy of Delta-AFP in Pre 1899<br />

Evaluation of PIVKA-II effic 1898<br />

LI-RADS hepatocellular carci 1906<br />

Transforming acidic coiled-c 1902<br />

Transjugular Intrahepatic Po 172<br />

A Model to Estimate Survival 1907<br />

A Retrospective Cohort Study 1904<br />

Correlation of MR and pathol 1896<br />

High 18F-FDG uptake on preop 1893<br />

Reduced sensitivity of ultra 1894<br />

The role of AFP in screening 1903<br />

GO3. Clinical: Hepatocellular<br />

Carcinoma and Cholangiocarcinoma<br />

A Single Center Experience w 428<br />

A Study for the Risk Factors 131<br />

A potential biomarker, Brf1 460<br />

A stromal liver gene signatu 364<br />

Activation of Stemness Genes 366<br />

Albi score predicts survival 489<br />

Analyses of progressive dise 482<br />

Aspirin Use is Associated wi 397<br />

Cholangiocarcinoma, a signif 416<br />

Circulating Osteopontin and 170<br />

Clinical Features and Treatm 486<br />

Clinical significance of mic 426<br />

Combining transarterial chem 391<br />

Comparative Effectiveness of 357<br />

Comparison of balloon-occlud 411<br />

Diabetes, HBV infection and 132<br />

Differences between radiolog 413<br />

Differences in Presentation 387<br />

Doxorubicin Resistance in He 352<br />

Efficacy of Continuous Nucle 439<br />

Ethnicity impacts aggressive 362<br />

Etiology-Specific Effect of 415<br />

Hepatic function and tumour 444<br />

Hepatocellular Carcinoma Sur 442<br />

Heterogeneity of advanced he 473<br />

High gamma-glutamyl transfer 389<br />

Hong Kong Liver Cancer stagi 409<br />

Impact of Hepatic p62 Overex 456<br />

In patients with HCC and mac 421<br />

In patients with hepatocellu 449<br />

Influence of hepatitis B vir 405<br />

Intrahepatic Cholangiocarcin 372<br />

Lamivudine vs. Entecavir for 422<br />

Liver microRNA hsa-miR-125a- 431<br />

Liver-directed (LD) Therapy 481<br />

MESIAH score is an effective 470<br />

Meta-Analysis: Proportions o 367<br />

MicroRNA expression in hepat 429<br />

Model to predict recurrence 400<br />

Multi-platform analysis of T 129<br />

Non-Cirrhotic Hepatocellular 488<br />

Optimal age to start surveil 350<br />

Oral nucleos(t)ide analogues 349<br />

Outcomes following Radiofreq 174<br />

Outcomes of patients with re 491<br />

Patients with alcohol compar 373<br />

Preclinical characterization 394<br />

Predicting Tumor Death – T 369<br />

Prediction of Hepatocellular 381<br />

Prediction of liver decompen 450<br />

Prognostic Score predicting 467<br />

Prognostic assessment of pat 355<br />

Prognostic significance of i 376<br />

Progranulin Autoantibodies a 463<br />

Reappraisal of portal vein t 479<br />

Retreatment with TACE: Trans 447<br />

Risk Differences of Hepatoce 359<br />

Risk assessment of hepatocel 346<br />

Risk factors of complication 171<br />

Screening Practices for Hepa 401<br />

Serological markers of extra 441<br />

Short-term preoperative trea 417<br />

Significant Suppression of L 410<br />

Simple serological tests and 475<br />

Staging Hepatocellular Carci 425<br />

Stereotactic Body Radiothera 471<br />

Surveillance at six month in 374<br />

Survival After Intra-Arteria 344<br />

Survival analysis of second 472<br />

TBI 302: A first-in-class th 424<br />

Targeting the TGF-β Pathway 130<br />

The Doylestown algorithm – 371<br />

The autophagy marker LC3 is 398<br />

The effect of additional tra 464<br />

The life time frequency and 382<br />

The role of cytokines and ad 478<br />

Toronto Hepatocellular Carci 348<br />

Treatment not needed for hyp 407<br />

Trends in the burden of cirr 128<br />

Undetected asymptomatic alco 378<br />

When to perform surgical res 365<br />

Whole blood coagulation test 351<br />

Long term outcome after rese 423<br />

MicroRNA-26 expression is si 436<br />

Endothelial nitric oxide syn 435<br />

A New Clinically-Based Stagi 127<br />

A Pilot Safety Study of Natu 440<br />

A randomized controlled stud 402<br />

Albi score predicts survival 490<br />

Analytic Morphomics Predicts 418<br />

Antiviral Therapy for Chroni 207<br />

Blood alpha-fetoprotein leve 437<br />

Building of a Scoring Model 419<br />

Characteristics of etiology- 370<br />

Clinicopathologic analysis o 465<br />

Combination of modified Resp 468<br />

Comparison of Hepatocellular 451<br />

Comprehensive analyses of mu 406<br />

Conversion therapy with hepa 399<br />

Current knowledge of guideli 434<br />

Development of a model predi 377<br />

Dysregulated peripheral and 454<br />

Effect of Antiviral Therapy 368<br />

Efficacy and safety of mirip 383<br />

Efficacy of Radiotherapy in 433<br />

Elevated Serum Wisteria flor 408<br />

Etiological difference in bo 461<br />

Exploration of cancer-prone 430


1372A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

Hepatic Decompensation After 412<br />

Hepatocellular Carcinoma Can 361<br />

Hepatocellular carcinoma has 445<br />

Hepatocellular carcinoma in 462<br />

Higher Incidence of Hepatoce 458<br />

Higher Risk for Hepatocellul 388<br />

Identification of candidate 384<br />

Identification of novel biom 358<br />

Impact of Viral Hepatitis Et 483<br />

Impact of screening on Hepat 404<br />

Improved Survival of Hepatoc 457<br />

Integrator complex subunit 6 375<br />

Korean validation and compar 452<br />

Liver Transplant Recipients 414<br />

Long non-coding RNA integrat 353<br />

Long-term Survival Analysis 474<br />

Multicentric Study of Liver 469<br />

Mutations of p53, ARID1A and 347<br />

Myeloid-derived suppressor c 392<br />

Natural history of portal ve 393<br />

Nucleos(t)ide analogue thera 345<br />

Overexpression of periostin 403<br />

Plasma MicoRNA-122 level as 379<br />

Prediction of Recurrence aft 363<br />

Preventive Effect of Urea-ba 354<br />

Pro-angiogenic Tie2-expressi 476<br />

Prospective cohort study for 455<br />

Racial Differences in the Pr 386<br />

Racial Disparity in Hepatoce 420<br />

Regional and Etiological Dif 459<br />

Skeletal muscle depletion as 477<br />

Small intrahepatic cholangio 487<br />

Sorafenib Therapy in HIV-inf 438<br />

Stereotactic body radiation 484<br />

Surgical resection versus ra 432<br />

Surveillance for Hepatocellu 360<br />

Survival and prognostic fact 485<br />

Survival of Patients with Ad 395<br />

Survival of Patients with Ad 448<br />

The Casein kinase II (CK2) i 396<br />

The ITA.LI.CA Staging System 356<br />

The presence of histological 466<br />

The prognosis and the recurr 453<br />

The radiofrequency ablation 385<br />

The relationship between the 443<br />

Trends of Hepatocellular Car 380<br />

Validation of the Metroticke 446<br />

Value of PIVKA II and AFP se 480<br />

Variability in False-Positiv 427<br />

Vitamin K dosing during sora 173<br />

Hepatotoxicity<br />

HO1. Pathogenesis and Mechanisms<br />

A comparative analysis of th 595<br />

A protective role of C/EBP h 592<br />

Caspase Inhibition Switched 598<br />

Differentially altered gut m 581<br />

Golgi Stress Response Is Ass 588<br />

HMGB1-driven Feedforward Hep 178<br />

Inhibition of NF-kB by deoxy 176<br />

Oxaloacetic Acid Protects th 179<br />

Phenotypic and genotypic cha 583<br />

Sumoylation Regulates Lipopo 594<br />

The Clinical-Pathological Sp 597<br />

The Hepatic “Matrisome” 587<br />

The influence of drug proper 596<br />

Transcriptomic Analysis Of H 175<br />

Acetaldehyde Triggers HCV-In 589<br />

Critical role of sirtuin 1-m 177<br />

EPAC mediated protection aga 582<br />

Effect of Branched Chain Ami 584<br />

Hormesis in cholestatic live 591<br />

Mitochondrial Fission Factor 590<br />

Mitochondrial-shaping protei 599<br />

Persistent Generation of Inf 585<br />

Protective effects of connex 586<br />

STING deficiency protects fr 180<br />

The Effect of Treatment with 593<br />

HO2. Drug Metabolism, Toxicity, and<br />

Therapeutics<br />

3D culture strategies for im 1921<br />

AMPK activation prevents and 1909<br />

Application of the Drug-Indu 1929<br />

Caffeine Confers Hepatoprote 1917<br />

Combined Activities of JNK1 1924<br />

Drug-Induced Liver Injury As 1920<br />

Gab1 adaptor protein is an e 227<br />

Gender-specific glucuronidat 1927<br />

Long-term Outcome of HIV+ Pa 1908<br />

Protective role of miR-122 a 99<br />

Severe and prolonged jaundic 1923<br />

Sodium 4-phenylbutyric acid 226<br />

Statins prevent liver tumor 1919<br />

The non-nucleoside reverse t 1913<br />

Transcriptional Profiling Of 1916<br />

A Rapid Test for Acetaminoph 1914<br />

Azathioprine And 6-Mercaptop 1926<br />

Death and Liver Transplantat 1922<br />

Drug Induced Liver Injury (D 1928<br />

Human mercapto-albumin level 1912<br />

Intervention of RIPK-depende 1918<br />

Rescue of hepatocytes from l 223<br />

The role of CD36 in acetamin 1910<br />

Translocation of Iron from L 1925<br />

Understanding the Relationsh 1915<br />

Withaferin-A Reduces Acetami 1911<br />

HO3. Predictors of Drug Toxicity<br />

HLA-A*33:01 is strongly asso 225<br />

Genome-wide investigation id 1933<br />

Lipotoxicity Accounts for Li 228<br />

Minocycline hepatotoxicity: 1930<br />

Predicting In Vivo Hepatotox 224<br />

Liver Disorders that can Mas 1931<br />

The Epidemiology of Suspecte 1932<br />

Human Cholestatic<br />

and Autoimmune Liver<br />

Diseases<br />

IO1. PBC/PSC and Other Cholestatic<br />

Disease<br />

Association between elevated 616<br />

Bezafibrate: A novel and eff 603<br />

CD39 expression regulates Th 608<br />

Characterization Of A Patien 630<br />

Cholestatic liver disease an 631<br />

Clinical Epidemiology of Pri 75<br />

Colectomy is associated with 622<br />

Combination Anti-Retroviral 641<br />

Elevation of alkaline phosph 634<br />

Emperipolesis mediated by CD 638<br />

Enhanced and Prolonged FGF19 604<br />

Expression of matrix metallo 632<br />

Gender and IBD phenotype are 76<br />

Ig Repertoire of Autoantigen 77<br />

Impact of NGM282 on the Inci 627<br />

Incidence and Impact of Deco 73<br />

Increased expression of ORM1 614<br />

Integrated Analysis of Effic 645<br />

Long-term follow-up of a mul 605<br />

Long-term safety and efficac 628<br />

NGM282, A Novel Variant of F 106<br />

Patients with Primary Biliar 610<br />

Patients with childhood-onse 600<br />

Phenotypic Variations in Pri 607<br />

S-Alkaline phosphatase is no 633<br />

Sustained Improvement in the 609<br />

The IgG/IgG4 mRNA ratio by q 637<br />

The prevalence and outcomes 626<br />

Thyroid dysfunction in prima 629<br />

Validation of serum fibrosis 624<br />

A Novel Prognostic Model for 74<br />

Alkaline Phosphatase as Biom 623<br />

Clinician confidence in stra 636<br />

Convergence of Two Predictiv 601<br />

Differentially Expressed Pla 602<br />

Down-regulation of the secre 621<br />

Efficacy and Safety of Obeti 625<br />

Evaluation of the Posology o 635<br />

Gene expression profiling of 615<br />

High-risk primary biliary ci 611<br />

In PSC with dominant bile du 618<br />

Innate Immunity Drives the I 619<br />

Intrahepatic Cholestasis of 78<br />

Long term impact of fibrates 642<br />

Long-Term Safety of OCA in P 644<br />

Modeling biliary fluid dynam 617<br />

Natural killer cells regulat 640<br />

Phenotypic and Cytotoxic Ana 639<br />

Predictors of recurrence and 643<br />

Serum metabolomic profile of 620<br />

The Modulation of Co-stimula 613<br />

Unique Immunologlobulin Glyc 606<br />

Vitamin A deficiency promote 612<br />

IO2. Autoimmune Liver Disease<br />

Autoimmune hepatitis type 1 302<br />

Features and patterns of con 312<br />

Impact of SNP rs2187668 in H 297<br />

In autoimmune liver disease 300<br />

Killer cell immunoglobulin-l 309<br />

Macrophage migration inhibit 298<br />

Mean Corpuscular Volume as S 303<br />

Possible involvement of acti 310<br />

Soluble PD-1 levels are asso 301<br />

Tacrolimus or Mycophenylate 308<br />

Autoimmune Hepatitis Disprop 306<br />

Carcinogenesis in autoimmune 304<br />

Immunity in HLA-DR4 expressi 305<br />

Methotrexate therapy for Aut 311<br />

Outcomes and mortality of de 307<br />

Plasma levels of heparanase 299


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) CATEGORY INDEX 1373A<br />

Inflammation and<br />

Immunobiology<br />

JO1. Animal Models<br />

Novel in vivo and in vitro m 2122<br />

Allogeneic and xenogeneic tr 2130<br />

CD8+ T-cells drive liver inj 2128<br />

Chemokine CCL3/CCR5-recruite 2119<br />

Chronic ethanol feeding supp 182<br />

IL-17 signaling in hepatocel 186<br />

IL-17A plays a pivotal role 2125<br />

Liver Plasmacytoid Dendritic 2123<br />

Osteopontin ablation drives 2124<br />

Small specific-sized Hyaluro 2121<br />

Small-specific-sized hyaluro 2132<br />

Sparstolinin B, a TLR4-antag 2133<br />

Upregulated hepatic expressi 2126<br />

Vitamin D deficiency promote 2118<br />

Combined bone marrow plus li 2129<br />

Hepatocellular carcinoma is 2127<br />

IFN-α stimulates IFN-γ exp 2120<br />

The gut-liver axis coordinat 2131<br />

JO2. Innate Immunity<br />

Activation of hepatic innate 1290<br />

Chronic alcohol consumption 1294<br />

Effects of myeloid cell-sele 1295<br />

Galectin-3 mediates NLRP3 in 1285<br />

High baseline expression of 1288<br />

IL28B genetic variants deter 1287<br />

Impaired TRAF6 methylation a 1291<br />

Implants of matrix-embedded 1292<br />

Kupffer cells signature in a 1286<br />

Proportion of Intrahepatic C 1289<br />

The long noncoding RNA VL30- 1284<br />

Identification of liver mono 1293<br />

Liver sinusoidal endothelial 184<br />

Macrophage Specific β-caten 1283<br />

Mincle Signaling Exacerbates 1282<br />

The antifibrotic effect of I 183<br />

JO3. Adaptive Immunity<br />

Intracellular crawling of ly 1296<br />

MAIT cells are enriched in p 1298<br />

Alterations in liver dendrit 1299<br />

Hepatitis B X-interacting pr 1301<br />

Inhibition of the FGL2:FcγR 1297<br />

Soluble CD163 plasma concent 1300<br />

Toll-like receptor 9 and B-c 1302<br />

JO4. Mechanisms of Injury<br />

15-PGDH prevents LPS/GalN-in 185<br />

Inflammasome Activation Due 2141<br />

Knockdown of Receptor Intera 2135<br />

Lack of DNase II exacerbates 2136<br />

Liver Enzymes Elevation In T 2144<br />

Macrophage activation throug 2134<br />

Osteopontin promotes cholang 2143<br />

Prior Blockade of TNF-α Sig 61<br />

TRPV4 regulates inflammation 2137<br />

The attenuation of inflammat 2139<br />

The gut-liver axis plays a c 181<br />

Immune Regulation of Mesench 2138<br />

Role of phagocytosis in coor 62<br />

TNFα mediates the liver:lun 2142<br />

The origin and differentiati 2140<br />

Liver Fibrogenesis and<br />

Non-Parenchymal Cell<br />

Biology<br />

KO1. Basic Fibrosis Research and<br />

Stellate Cell Biology<br />

Branched-chain amino acids p 1377<br />

Cartilage oligomeric matrix 1382<br />

Characterization of obesity- 1386<br />

Curcumin blocks the loss of 1412<br />

Deactivation of human and ra 1418<br />

Deletion of Wntless in macro 1362<br />

Deletion of fibrocytes in mi 187<br />

Dual combination therapy dir 1367<br />

Efficacy of an ASK1 Inhibito 1359<br />

Endothelial Niche Derived No 1360<br />

Essential Roles of RNA-bindi 1396<br />

Exosome-mediated activation 1358<br />

HIV infected Kupffer cells m 1397<br />

Hepatic stellate cell-derive 1414<br />

Hepatocyte autophagy impairm 1364<br />

Human precision Cut Liver Sl 1409<br />

IL-22 enhances TGF-beta pro- 1380<br />

Identification of Axon guida 1405<br />

Induction of XBP1 Leads to H 1392<br />

Inhibition of microRNA-214 a 1395<br />

Integrin αvβ6 critically r 1368<br />

Liver fatty acid-binding pro 1361<br />

MeCP2 exerts global control 1381<br />

Melatonin Suppresses Activat 1398<br />

MicroRNA profiling of circul 1404<br />

Myofibroblastic conversion o 1354<br />

NLRP3 inflammasome modulatio 1371<br />

Novel treatment strategy for 1374<br />

Oral hepatotropic DPP4 inhib 1376<br />

Paracrine suppression of pro 1416<br />

Proof of liver regeneration 1375<br />

Reactive gamma-ketoaldehydes 1407<br />

Rev-erb and TGF-β Different 1389<br />

Selective antibody targeting 1379<br />

Single adenoviral delivery o 1399<br />

Targeting Activated Stromal 1402<br />

The AMPK-related kinase NUAK 1410<br />

The bile acid-phospholipid c 1411<br />

The paracrine effect of hepa 1373<br />

Versican: A Novel Modulator 1413<br />

Interleukin 15 regul 1370<br />

Antifibrogenic properties of 1391<br />

Binding of hepatic stellate 1356<br />

Calcium Mobilization through 1415<br />

Cancer associated fibroblast 1420<br />

Circulating exosomes from he 192<br />

Engineering Healthy And Cirr 1388<br />

Expression of ENTPD1/CD39 is 1394<br />

Fibrous network composition 1366<br />

GP38/PODOPLANIN marks a nove 1385<br />

Galectin-1 promotes the acti 1400<br />

Hepatocyte- and hepatic stel 1365<br />

Human neutrophil peptide-1 e 1421<br />

IL-10-producing regulatory B 1383<br />

IL-13Rα1 signaling and M2 m 1353<br />

Knockout of the substance P 190<br />

Liver Fibrosis in a Mouse Mo 1422<br />

Loss of Cytoglobin Exacerbat 191<br />

Mast cells interact with pro 1363<br />

Membrane type I-matrix metal 1378<br />

Mice lacking calmodulin kina 1387<br />

Modulation of Extracellular 1403<br />

Molecular mechanism responsi 1390<br />

Opioid Growth Factor Recepto 1369<br />

PDGFRα ubiquitination leads 1406<br />

Potential of connective tiss 1357<br />

Role of CREBH activation ind 1393<br />

SNAP-23 Heterozygous Mice Ar 1384<br />

Selective disruption of dyna 1355<br />

Sequence Profiling of miRNAs 1408<br />

Serine Protease Omi/HtrA2 Re 1419<br />

Splenectomy attenuates advan 188<br />

TRPC6 Channels Contribute to 1401<br />

The role of Mesothelin in th 189<br />

KO2. Imaging and Noninvasive<br />

Markers of Liver Disease<br />

Comparison of noninvasive fi 790<br />

A Novel Diagnostic Index to 774<br />

Assessment of biliary fibros 788<br />

Changes in liver stiffness b 777<br />

Collagen maturation measured 780<br />

Comparative 10-Years Prognos 786<br />

ELF Test Thresholds for dise 776<br />

ELF predicts clinical out 793<br />

Effect of inflammation on li 791<br />

Feasibility of three ultraso 802<br />

Glyco-isomer of Serum Mac-2- 796<br />

Liver stiffness assessed by 787<br />

Magnetic resonance elastogra 45<br />

Non-invasive fluorescent ima 44<br />

Predicting clinical outcomes 784<br />

Role of Acoustic Radiation F 800<br />

Serum Lipidomic Profiling fo 48<br />

Supersonic Shear Imaging Is 785<br />

The liver stiffness evaluati 795<br />

The predictive value of 13C- 46<br />

Usefulness of Shear Wave Ela 775<br />

Usefulness of ultrasound ela 781<br />

qFibrosis scoring for differ 782<br />

2D Shear Wave Elasto 792<br />

Differences in Diagn 789<br />

A Hepatic Stellate Cell Secr 778<br />

A clustering based fully aut 801<br />

AnnexinV/EpCAM (CD326) tumor 43<br />

Association of skin capsular 794<br />

Coffee consumption reduces l 47<br />

Diagnostic disparity of FIB4 779<br />

Fast macromolecular proton f 803<br />

High risk of misclassifying 773<br />

Liver fibrosis evaluation us 797<br />

Machine-learned Image Analys 798<br />

Real-time shear wave elastog 799<br />

Serum Wisteria floribunda ag 783<br />

KO3. Clinical and Translational<br />

Fibrosis Research<br />

Activated human hepatic stel 120<br />

Azathioprine exerts an antif 1445<br />

Collagen and tissue turnover 1437<br />

Collagen proportionate area 1440<br />

Collagen scoring by qFibrosi 1436<br />

Comparison of 16 blood and/o 1439


1374A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

Concerted effects of the two 1429<br />

Correlations between hepatic 1428<br />

Evaluation of EASL recommend 117<br />

Host microbiota dictates the 1427<br />

How gold is the “gold stan 1443<br />

Innate Lymphoid Cell (ILC) S 1430<br />

Lysyl oxidase-like-2 (LOXL2) 1442<br />

PEGylated TRAIL treatment am 115<br />

Reproductive Aging, Independ 1444<br />

Risk of cirrhosis in first d 1431<br />

Serum lysyl oxidase-like-2 ( 1435<br />

Serum wisteria floribunda ag 1446<br />

Statins, ACE inhibitors and 116<br />

Systemic administration of a 118<br />

The epidemiology of cirrhosi 1424<br />

Treatment of Hepatic Fibrosi 1423<br />

A marker of true type III co 1432<br />

A novel canine liver cirrhos 1447<br />

A novel glycobiomarker, Wist 1448<br />

Antifibrotic efficacy of a T 1434<br />

Comparing the liver stiffnes 1441<br />

Fibrosis is not just fibrosi 119<br />

Hepatitis C treatment and Pr 1450<br />

In vivo cell specific gene s 1426<br />

In vivo reprogramming of myo 98<br />

New Prognostic Model for 1-y 1438<br />

Novel assessment of hepatic 1425<br />

Safety, Pharmacokinetics, an 1433<br />

Second harmonic generation m 1449<br />

Liver Transplantation and<br />

Liver Surgery<br />

LO1. Cellular Immunobiology,<br />

Preservation, and Cell<br />

Transplantation<br />

Autologous hepatocyte transp 2<br />

Blockade of RIP kinase pathw 12<br />

Delayed Donor Bone Marrow In 54<br />

Identification of microRNAs 341<br />

Novel Vasodilator and Anti-i 340<br />

Previous ischemic preconditi 342<br />

Extracellular vesicles from 335<br />

Liver Fabrication Using Dece 336<br />

Promotion of liver transplan 338<br />

The innate immune cell micro 339<br />

Three-dimensional bioactive 337<br />

LO2. Donor and Allocation Issues,<br />

Living Donor and Split Liver<br />

Transplantation, and Hepatobiliary<br />

Surgery<br />

3D Printed Liver Models for 840<br />

A Failure of the Model of En 849<br />

A Novel Percutaneous Organ P 216<br />

Application of Gadoxetate Di 869<br />

Auxiliary partial orthotopic 867<br />

Clinical impact of 18F-FDG-P 862<br />

Comparison of Liver Transpla 49<br />

Controlled attenuation param 876<br />

Functional Status Predicts P 845<br />

Impact of UNOS Imaging Crite 10<br />

Liver Transplant Outcomes an 69<br />

Liver transplantation waitin 68<br />

Living donor liver transplan 860<br />

Living or Deceased Donor Liv 863<br />

Mechanisms of Gender Dispari 865<br />

Natural history of liver tra 873<br />

Older Liver Transplant (LT) 844<br />

Outcomes of Patients Undergo 854<br />

Physical function predicts l 841<br />

Recanalization of complete a 846<br />

Safety of live liver donatio 859<br />

The impact of non-identical 72<br />

Obesity is predictive of 839<br />

A shift to the left: A compa 838<br />

Association of Pre-Transplan 855<br />

Comparable Short And Long Te 70<br />

Coronary Angiography in Pati 874<br />

Could Share 35 disadvantage 6<br />

Effective and Safe postopera 71<br />

External Validation of a US- 853<br />

Graft Survival following Dom 868<br />

Interactive Effect Of Donor 878<br />

Lack of Radiologic-Pathologi 870<br />

Long Term Survival Of Liver 875<br />

Longer Term Efficacy of Simu 836<br />

Low serum testosterone predi 842<br />

Novel Sonar Guided Technique 858<br />

Objective Measurement of Kar 850<br />

PALBI-An Objective Score Bas 851<br />

Patient Safety in Living Don 847<br />

Predicting Futility of Liver 856<br />

Prediction of waitlist morta 9<br />

Predictors of Retransplantat 877<br />

Preliminary Data Analysis fr 837<br />

Region-Based Projections of 67<br />

Remnant Liver Ischemia Is As 857<br />

Significance of functional h 866<br />

The effect of recipient and 848<br />

The transportable machine pe 215<br />

Third Liver Transplantation 864<br />

Tool for efficient and effec 843<br />

Traveling for a Liver Transp 872<br />

Traveling for a Liver Transp 861<br />

Treatment with Optifast Redu 11<br />

Using Biomarkers to Predict 871<br />

Wait Times of 18 Mon 50<br />

“Oldest Old” (≥80 year 852<br />

LO3. Immunosuppression, Outcomes,<br />

and Complications<br />

15 first-days over-immunosup 1251<br />

Females, Hispanics, and Nona 1252<br />

A New Score based on explant 1232<br />

Aging of liver transplant re 1211<br />

Are cirrhotic patients await 1239<br />

Coffee consumption promotes 1222<br />

Comparative analysis of rena 1233<br />

Conversion from Sirolimus to 1247<br />

Downstaging or Bridging Ther 1262<br />

Early Cytomegalovirus infect 1268<br />

Early Liver Transplantation 1215<br />

Everolimus is Associated wit 1226<br />

Glycine is graft protective 1227<br />

Good results of liver transp 1269<br />

How intra-operative factors 1271<br />

IgG4 status in explanted liv 1246<br />

Impact of Positive Crossmatc 1254<br />

Lack of effect of CMV reacti 1230<br />

Liver Transplant (LT) Candid 1259<br />

Liver Transplantation For Al 1250<br />

MELD Based Outcomes of Simul 1249<br />

Metabolic Syndrome and Insul 1270<br />

Minimal hepatic encephalopat 51<br />

Morbid Obesity and Diabetes 1273<br />

Neurological Syndrome de nov 1256<br />

Outcomes of Transjugular Int 1280<br />

Post-Transplant Lymphoprolif 1275<br />

Pre-Transplant Sarcopenia Do 1267<br />

Pre-transplant portal vein t 1258<br />

Prediction of Posthepatectom 1213<br />

Recipient Risk Factors for A 1266<br />

Reduced Survival in Elderly 1208<br />

Reevaluating Risk Factors fo 1241<br />

Renal Function Improvement i 1281<br />

Renal function at month 1 an 1237<br />

Reporting Physical Function 1219<br />

Retained HLA Class II Donor 1236<br />

Risk Stratification for Opti 1277<br />

Role of Coronary Artery Calc 1255<br />

Rotational Thromboelastometr 1260<br />

Sarcopenia is a Risk Factor 1276<br />

Severity and Predictive Fact 1264<br />

Sublingual administration of 1217<br />

Synergistic anticancer effec 1278<br />

Systolic Dysfunction Post-Li 1261<br />

The Association of Pre-Trans 1242<br />

The Combination of Strongly 1228<br />

The Small Molecule TLR9 Anta 1209<br />

Two-Year Results of a Pilot 1214<br />

Ursodeoxycholic Acid as Adju 1224<br />

Outcomes Post-Liver 1243<br />

A Multi-Center Study to Deve 53<br />

A Multi-center Study on Down 4<br />

Acute Kidney Injury among Pa 1220<br />

Assessment of Energy Metabol 1235<br />

Blood Phosphatidyl ethanol i 1231<br />

Echocardiography for Liver T 1257<br />

Effective Systems of Care Im 1272<br />

Evaluation of risk indices t 1279<br />

Incidence of Acute Cellular 5<br />

Increase in Red Cell Distrib 1240<br />

Intrahepatic Cholangiocarcin 1274<br />

Liver Transplantation for Po 1216<br />

Liver Transplantation in Old 1245<br />

Long-Term Outcome of Extrahe 1234<br />

Ohio Solid Organ Transplanta 1265<br />

Outcomes of Patients with Fa 1263<br />

Outcomes of acute-on-chronic 1212<br />

Physical activity measured b 1248<br />

Pre-transplant echocardiogra 1218<br />

Primary Biliary Cirrhosis: N 1210<br />

Quantiferon-CMV testing afte 1225<br />

Recipient obesity is an inde 1223<br />

Renal Function Recovery and 1253<br />

Renal Recovery Outcomes afte 1221<br />

The Functional Basis for Cog 52<br />

The Gap Between Assessed Phy 1244<br />

The coming of age of the gra 1229<br />

Value of Magnetic Resonance 1238<br />

LO4. Viral Hepatitis<br />

Detectable HBVDNA in liver a 1739<br />

Determinants and impact of e 3<br />

Effect of simeprevir adminis 1754<br />

HBV cccDNA evaluation in HBV 1740<br />

Hepatitis E infection in imm 1752<br />

Impact of Sofosbuvir-Based R 204<br />

Liver Cancer in Patients Cur 1742<br />

Prevention of Recurrent Hepa 200<br />

Prevention of post-transplan 1751


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) CATEGORY INDEX 1375A<br />

The effect of ledipasvir/sof 1741<br />

Two Combined Genetic Markers 1<br />

An Evidence-Based Algorithm 1745<br />

Curing Decompensated Wait Li 202<br />

Current Trends in Liver Tran 199<br />

Effects of pre-transplant he 1749<br />

HCV Infection in Aging Baby 201<br />

Long-term Outcome of Entecav 1746<br />

Long-term complete prophylax 203<br />

Long-term outcomes after hep 1748<br />

Oral Interferon-free Antivir 1753<br />

Protective Birth Cohort Effe 1750<br />

Renal safety of nucleos(t)id 1744<br />

Treating HCV patients on the 1743<br />

Metabolic and Genetic<br />

Disease<br />

MO1. Hemochromatosis, Wilson<br />

Disease, and a-1 Antitrypsin<br />

Deficiency<br />

UGT1A regulation by a member 2108<br />

Accumulation of hepatitis B 2115<br />

Association between Psoas Mu 2110<br />

Clinical features, biochemic 2106<br />

Congenital glycosylation def 2102<br />

Dicer-Substrate Small RNAs a 2116<br />

Liver-specific deletion of t 197<br />

Mutational analysis of ATP7B 2107<br />

Pitfalls in Erythrocyte Prot 2111<br />

Variation in erythrocyte and 2112<br />

A Nationwide, Population-bas 2113<br />

Cardiac MRI for Assessment o 2104<br />

Combining high-throughput ge 193<br />

Functional analysis and drug 2103<br />

GNPAT Variant D519G in patie 2117<br />

Genetic variants associated 2105<br />

Hepatocellular Carcinoma in 2101<br />

Liver Fibrosis is Present in 138<br />

Relapse of Porphyria Cutanea 2100<br />

Relapse of porphyria cutanea 2114<br />

The real world experience of 2099<br />

Wilson disease: gestational 2109<br />

Midlevel Professionals<br />

(nurses, nurse<br />

practitioners, etc.)<br />

NO1. Behavioral and Quality Issues<br />

Interferon (IFN), Ribavirin 315<br />

TNF-alpha is the Most Consis 316<br />

Longitudinal alteration in h 314<br />

Monitoring Pruritus for Pati 317<br />

The Need for Improved Liver 313<br />

Uncertainty and Hepatitis C: 318<br />

NO2. Practice Issues<br />

Significant non–adherence 646<br />

Hepatology nursing in Canada 647<br />

Pharmacists mitigate effect 648<br />

Pediatric Hepatology<br />

OO1. Pediatric Liver Disease - Basic<br />

Science<br />

A novel mouse model for the 2098<br />

Enteral Obeticholic Acid Pre 194<br />

Enteral Obeticholic Acid Pro 1689<br />

Hepatic macrophage IL-1 beta 1687<br />

Molecular mechanism of IL-1 1688<br />

OO2. Biliary Atresia and Cholestasis<br />

B cell activation in the rot 1699<br />

Biliatresone causes GSH redu 195<br />

C5aR targets immune and epit 1698<br />

Exome Sequencing and de novo 1693<br />

Prophylactic endoscopic trea 134<br />

Role of transient elastograp 1697<br />

DXA Bone Density in Alagille 1694<br />

Diagnostic determination sys 1696<br />

Expansion of regulatory T ce 1695<br />

Human iPSC-derived hepatocyt 1690<br />

King’s-Variceal Prediction 1691<br />

Molecular genetic dissection 135<br />

Neurodevelopmental Outcomes 133<br />

The Rotavirus Induced Mouse 1692<br />

OO3. Viral and Autoimmune<br />

Hepatitis<br />

Autoimmune Liver disease in 1711<br />

Dissecting cellular entry me 1703<br />

Prediction of nucleotide ana 1700<br />

Quantitative Maternal Hepati 1701<br />

The Impact of Hepatitis B Va 1705<br />

Validation of transient elas 136<br />

“Hepatic Granulomas: 18 Ye 1712<br />

CD4+ helper T Cell Dysregula 1706<br />

Differentiating autoimmune h 1713<br />

HCV infection shifts CD8+ T 1708<br />

Long-term follow up of child 1710<br />

Pharmacokinetics of Once Dai 1707<br />

Pre-core region mutations an 1704<br />

Predictors of poor outcome i 1702<br />

The Natural History of Autoi 1709<br />

OO4. Metabolic and Genetic<br />

Diseases<br />

NOTCH2 variants in cholestat 1721<br />

A functional classification 196<br />

Congenital Lactic acidemia ( 1725<br />

Liver Involvement in Turnerâ 1720<br />

Autosomal Recessive Polycyst 1718<br />

Bile acid metabolism is alte 1726<br />

Hepatic Manifestations in Ad 1717<br />

Liver Disease in Subjects wi 1722<br />

Nocturnal Hypoxia is associa 1715<br />

Primary immunodeficiencies i 1723<br />

Recurrent recessive mutation 1719<br />

Risk Factors Associated with 1716<br />

Screening for Mitochondrial 1714<br />

Zinc monotherapy for young p 1724<br />

OO5. Pediatric Liver Transplantation<br />

AST-to-platelet ratio index 1733<br />

Epidemiology and outcomes of 1737<br />

Preformed and de novo DSA ar 1735<br />

Serum autoantibodies are ass 1728<br />

Surveillance biopsies in ped 1736<br />

Immunization Practic 1738<br />

Distance Affects Mortality o 1732<br />

Dynamics of pediatric liver 1734<br />

Hospitalizations for Respira 137<br />

Pediatric liver transplant r 1730<br />

Pre-transplant mortality ris 1727<br />

Racial and ethnic variations 1729<br />

Regulatory T cells in de nov 1731<br />

Portal Hypertension and<br />

Other Complications of<br />

Cirrhosis<br />

PO1. Experimental<br />

Adipose Tissue-Specific ATGL 1531<br />

Diabetes is Associated with 1527<br />

ELF and Angiopoietins acc 1530<br />

Elevated galectin-3 in cirrh 1529<br />

Enoxaparin treatment does no 1520<br />

Hepatocyte Tissue Factor Con 81<br />

Inflammation and hypoxia reg 1519<br />

L-carnitine prevents ammonia 1528<br />

Liver NK cells from NLG4-/- 1532<br />

Long term oral treatment of 1523<br />

Neuro-inflammation in Cirrho 1521<br />

Portal tract fibrosis is ind 1526<br />

Simvastatin prevents the agg 1518<br />

Statins activate the canonic 80<br />

Taurine lowers portal pressu 149<br />

The Evaluation of Portal Hyp 1524<br />

The effects of apoptosis inh 79<br />

Ammonia Produces Pathologica 1517<br />

Emricasan, a pan caspase inh 1522<br />

The soluble guanylyl cyclase 1525<br />

Usefulness of spleen stiffne 1533<br />

PO2. Ascites, Renal Dysfunction, and<br />

Hepatorenal Syndrome<br />

Acute Kidney Injury definiti 266<br />

Acute Kidney Injury in criti 275<br />

Acute kidney injury in cirrh 288<br />

Chronic Rifaximin Therapy De 271<br />

Clinical Utilization of Seru 280<br />

Coronary microvascular dysfu 268<br />

Deleterious effect of beta-b 265<br />

Diagnostic accuracy of the P 264<br />

Disturbance in diurnal renal 267<br />

Effect of Beta-Blockers on S 282<br />

Is Midodrine, Octreotide, an 286<br />

Midodrine and Tolvaptan in P 259<br />

Multicenter retrospective an 281<br />

New score CyBAC-Del incorpor 262<br />

Organ Failures Associated Wi 83<br />

Predictive factors for a goo 296


1376A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

Predictors of response and s 278<br />

Serelaxin increased renal bl 260<br />

Slow, continuous low-dose al 274<br />

Terlipressin given by contin 148<br />

The Trigger Matters – Outc 270<br />

The association of non-selec 82<br />

Urine cystatin C is a strong 277<br />

Acute kidney injury and shor 289<br />

Ascites Neutrophil Gelatinas 272<br />

Ascites neutrophil dysfuncti 261<br />

CA-125 significance in cirrh 295<br />

Efficacy and safety of intra 279<br />

Goal-Directed Treatment for 290<br />

Hepatiq Measure Of Hepatic F 269<br />

Increasing incidence and cos 283<br />

Is prevalence of spontaneous 285<br />

Low ascites levels of cell m 276<br />

Renal Biopsy Histopathologic 284<br />

Signicant Ascites in Childre 294<br />

TIPS with PTFE-covered stent 263<br />

The decreased urinary aquapo 287<br />

Use of non-selective beta bl 273<br />

Usefulness of the hepatic ve 291<br />

Validation of Ascites-specif 293<br />

What is the evolution of Gra 292<br />

PO3. Varices and Bleeding<br />

An early experience of the B 764<br />

A Randomized, Multi-center, 736<br />

Atorvastatin and propranolol 750<br />

Balloon-occluded retrograde 748<br />

Clinical and histologic corr 737<br />

Correlation between noninvas 739<br />

Direct oral anticoagulants h 761<br />

Does Haemodynamic or Clinica 772<br />

External validation of risk- 742<br />

Hemostatic profiles in liver 740<br />

Inter-observer variability i 744<br />

Ischemic hepatitis following 146<br />

Obesity is associated with i 735<br />

Occlusion of portosystemic s 760<br />

Prevention of Recurrent Vari 745<br />

Propranolol for Prevention o 733<br />

Risk factors for in-hospital 765<br />

Serelaxin reduced portal pre 753<br />

Serum lysyl oxidase-like-2 ( 746<br />

The impact of esophagogastri 759<br />

The predictive value of chan 771<br />

Transient Elastography and s 755<br />

Transjugular intrahepatic po 769<br />

Viral suppression with IFN-f 732<br />

Transient elastograp 762<br />

A prospective study comparin 752<br />

Acute kidney injury predicts 757<br />

Argon Plasma Coagulation Ver 758<br />

Baveno`s VI non-invasive app 84<br />

Comparison of Beta blockers 766<br />

Effect of nonselective beta- 768<br />

Efficacy and safety of treat 767<br />

Efficacy of carvedilol as an 738<br />

Establish risk scores for pr 751<br />

Factors Influencing The Firs 734<br />

Long-term results of balloon 756<br />

Placebo-Controlled, Randomiz 743<br />

Prospective validation of th 749<br />

Shunt or meso-portal bypass 763<br />

Spleen stiffness assessed wi 741<br />

The relationship between the 770<br />

Transjugular intrahepatic po 754<br />

PO4. Encephalopathy and Other<br />

Complications<br />

Acute decompensation of cirr 1482<br />

Ammonemia predicts severity 1510<br />

Anticoagulation in patients 150<br />

Autonomic cardiac dysfunctio 1495<br />

Bioelectrical impedance vect 1471<br />

Bromocriptine is effective i 147<br />

Characteristics, indications 1508<br />

Clinical Characteristics of 1464<br />

Conjugated bile acids suppre 1507<br />

Efficacy of Ornithine Phenyl 1509<br />

Erectile Dysfunction in pati 1476<br />

Gut Microbiota Alterations a 1463<br />

Increasing frequency of gram 1489<br />

Liver Injury Independently P 1488<br />

Low Serum Leptin Level Is A 1496<br />

Minimal hepatic encephalopat 1502<br />

Muscle mass in liver transpl 1468<br />

Myocardial strain as a nonin 1494<br />

Obesity Paradox: Obesity is 1469<br />

Perceptions of the ‘cirrho 1477<br />

Plasma Metabolomic Profiling 1470<br />

Portal vein thrombosis, mort 1501<br />

Preoperative plasma glucose 1492<br />

Progressive Loss of Hematopo 1483<br />

Proliferative lymphangiogeni 1514<br />

Real Time Pressure Volume Lo 1503<br />

Role of 100 % FiO2 PaO2 and 1513<br />

Sarcopenia and Sarcopenic-Ob 1481<br />

Six minute walking test perf 1484<br />

Six-week administration of l 1475<br />

Transient elastography relat 1466<br />

Transjugular Liver Biopsy Is 1498<br />

Use of betablockers, previou 1511<br />

Usefulness of Balloon-Occlud 1486<br />

Usefulness of Balloon-Occlud 1474<br />

Validation of Relationship b 1504<br />

Videogame Training Improves 1491<br />

Sarcopenic obesity i 1500<br />

A half of hepatic hydrothora 1499<br />

Comparison Of Thrombin Gener 1505<br />

Development of a Novel Clini 1485<br />

Development of an Electronic 1467<br />

In Vivo Assessment Of Portal 1506<br />

Inadequate Food And Water In 1493<br />

Is it safe for patients with 1472<br />

Lactulose treatment improves 1465<br />

Linkage and interaction betw 1516<br />

Myocardial blood flow is red 1480<br />

Non-invasive assessment of p 1490<br />

Osteopontin is a new prognos 1487<br />

Prospective Multicenter Obse 1478<br />

Randomized Controlled Trial 1473<br />

Relative Adrenal Insufficien 1515<br />

Sarcopenia predicts moratili 1479<br />

The World Congress of Gastro 1497<br />

The burden of osteoporotic h 1512<br />

PO5. Infections and Acute on Chronic<br />

Liver Failure<br />

25-hydroxyvitamin D deficien 2086<br />

Adding C-reactive Protein an 2089<br />

Cholangitis is a Leading Com 2084<br />

Circulating interleukin 6, 1 2075<br />

Comparison And Outcomes Of 5 103<br />

Effects of norfloxacin thera 145<br />

Lack of evidence for bacteri 2094<br />

Liver cirrhosis and short-te 2095<br />

Neutrophil to Lymphocyte Rat 2085<br />

Obesity is a Major Predictor 2081<br />

Polymorphisms in the interle 2087<br />

Presepsin as a new biomarker 2090<br />

Prevalence, risk factors and 2070<br />

Sepsis in Cirrhosis Patients 2093<br />

Short-term infusion of soyab 2088<br />

The Epidemiology and Clinica 2073<br />

The NDP52 Val248Ala variant 2077<br />

The macrophage activation ma 2071<br />

The relationship of fecal ba 2091<br />

A Broad Spectrum Ant 2067<br />

A prognostic model evaluatin 2096<br />

Assessment and relevance of 2078<br />

Characterisation of blood mi 2074<br />

Characteristics and outcome 2079<br />

Hispanic ethnicity is associ 2083<br />

Impact of genetic variation 2069<br />

Liver Cirrhosis Is Strongly 2082<br />

MELD score and liver stiffne 2080<br />

Pneumonia Outcome Score In C 2092<br />

Probiotic supplementation im 2068<br />

Sequential assessment of Acu 2072<br />

Systemic inflammatory respon 2076<br />

The relationship between pro 2097<br />

Steatosis and<br />

Steatohepatitis<br />

QO1. Steatohepatitis: Experimental<br />

A DPP-4 inhibitor, sitaglipt 966<br />

NorUDCA reduces liver injury 977<br />

The obese insulin resistant 978<br />

A Balanced Carbohydrate and 950<br />

Association between nonalcoh 964<br />

Adipose tissue-derived strom 947<br />

Alteration of Neutrophil-der 898<br />

An Intact Fgf15 Signaling Pa 916<br />

Ancestral starvation-activat 936<br />

Angiogenesis in the pathogen 954<br />

Anti-Fibrotic Effect of Auto 141<br />

Assessment of Donor and Envi 915<br />

Beneficial effects of obetic 893<br />

Blocking of IL-17A signaling 890<br />

CHOP and the Unfolded Protei 884<br />

Cholesterol crystallization 923<br />

DSS colitis has tumorinogene 967<br />

Deficiency of intestinal muc 59<br />

Deletion of G-protein-couple 885<br />

Differential carbonylation o 921<br />

Dual targeting of nuclear re 897<br />

Ectopic expression of N-acet 906<br />

Ethanol-Mediated Lipocalin 2 879<br />

Fatty Acid-Stimulated Inflam 935<br />

Fenofibrate and LXRα agonis 924<br />

Fermented soymilk prevent fr 951<br />

FoxO3 increases miR-34a to c 60<br />

Free fatty acid treatment re 975<br />

GFT505 (ELAFIBRANOR) prevent 974<br />

Glucagon-like peptide-1 pote 955<br />

Hepatic Estrogen Receptor Ne 956<br />

Hypoxia accelerates fatty ac 930<br />

Inhibition of Mitophagy in N 968<br />

L-carnitine ameliorates hepa 917<br />

Lipid-induced Endoplasmic Re 883


HEPATOLOGY, VOLUME 62, NUMBER 1 (SUPPL) CATEGORY INDEX 1377A<br />

Liver-Directed Allosteric In 957<br />

Metformin targets a phosphoS 958<br />

MicroRNA-21 is a Potential L 102<br />

Mixed Lineage Kinase 3 Media 895<br />

NADPH oxidase (NADPHox) play 896<br />

NFkB inhibition by Andrograp 969<br />

NOX2 deficiency attenuates e 910<br />

Non-invasive quantitative de 913<br />

Nuclear localization of auto 891<br />

Omentectomy prevents nonalco 939<br />

Over-expression of HNF4α at 941<br />

Peretinoin, an acyclic retin 144<br />

RIP3-dependent hepatocyte ne 940<br />

Reduction of Hepatic 27-Hydr 970<br />

Regulation of hepatocellular 929<br />

Regulatory T Cells Prevent A 909<br />

Role of Fatty Acid Desaturas 56<br />

Role of Fcgamma receptors in 971<br />

Role of the receptor tyrosin 973<br />

S-Adenosylmethionine Treatme 953<br />

Soluble Intercellular Adhesi 892<br />

Standard American Diet (SAD) 946<br />

Targeting enterohepatic bile 140<br />

The dual FXR/TGR5 agonist IN 882<br />

The dual FXR/TGR5 agonist IN 902<br />

Vitamin D through Induction 904<br />

miRNA-339-3p is increased in 960<br />

L-Fabp deletion attenuates b 911<br />

A novel transcriptional repr 925<br />

Aging associated impaired me 901<br />

Akkermansia Muciniphila Is D 932<br />

Aliskiren reduces liver and 965<br />

Alterations in NK cell pheno 938<br />

Analysis of T and myeloid ce 928<br />

Antihypertensive therapy imp 937<br />

Bile Acid Sequestrant Preven 139<br />

Blocking the NLRP3 inflammas 55<br />

CX3CR1 is a gatekeeper for i 100<br />

CXCL1 promotes steatohepatit 903<br />

DNA Methylation Profiles of 889<br />

DRX-065, the stabilized R-en 143<br />

Deletion of orphan nuclear r 920<br />

Effects of Dietary Different 926<br />

Enhanced expression of Rubic 97<br />

Exacerbation of non-alcoholi 959<br />

Fecal Alkaline Phosphatase I 905<br />

Genetic Ablation of Phosphat 900<br />

Genetic analysis of NAFLD wi 943<br />

Hepatic NOD-like receptors, 886<br />

Hepatic but not Intestinal S 919<br />

Hepatocyte-specific autophag 894<br />

Hepatocyte-specific depletio 880<br />

Hepatocytes Release Ceramide 58<br />

High Milkfat Diet Causes Mor 907<br />

High and Low Capacity Runner 948<br />

High calories intake and par 949<br />

Human amniotic epithelial ce 945<br />

Increased 53-binding protein 952<br />

Increased sensitivity and co 976<br />

Inhibition of β-hydroxybuty 979<br />

Inhibitor Of DNA Binding 2 G 944<br />

Key role of ASMase in the su 899<br />

Leucine synergizes with metf 933<br />

Liver MicroRNA-21 is Overexp 887<br />

Liver sinusoid endothelial c 927<br />

Lysosomal cholesterol in Kup 922<br />

Mast Cell Involvement in Non 914<br />

MiR-128-3p is enriched in th 57<br />

Ornithine transcarbamylase g 908<br />

Predicted prevalence of NAFL 931<br />

Regulation of Adaptive Therm 881<br />

Regulation of Hepatocellular 972<br />

Role of fibroblast growth fa 934<br />

TRPV4 deficiency enhances TL 962<br />

Thioesterase Superfamily Mem 942<br />

Treatment with the FXR-TGR5 142<br />

Two different aspects of PD- 888<br />

Western diet for 7 weeks in 961<br />

Wheat Bran Autolytic Peptide 918<br />

Intrahepatic expression leve 912<br />

β7-Integrins and MAdCAM-1 h 963<br />

QO2. Steatohepatitis: Clinical and<br />

Therapeutic<br />

A Meta-analytic Assessment o 2212<br />

A novel noninvasive diagnost 2156<br />

ANA Risk (Age-NAFLD-Apri Ris 2248<br />

Adults With Nonalcoholic Ste 237<br />

Application of transient ela 238<br />

Are the phenotypes of nonalc 2223<br />

Assessment of parameters rel 2254<br />

Association of Nonalcoholic 2219<br />

Beneficial effects of the du 162<br />

Can FibroMeter VCTE be us 2180<br />

Can we avoid liver transplan 2215<br />

Cenicriviroc and Pioglitazon 2247<br />

Characterization of insulin 2239<br />

Clinical features and risk f 2171<br />

Combination of serum HA, CK1 2160<br />

Combined effects of the pros 2195<br />

Controlled Attenuation param 2181<br />

Differences of lifestyle beh 2243<br />

Digoxin provides hepatoprote 2179<br />

Discovery of plasma biomarke 2232<br />

Effect of biopsy length on h 2153<br />

Efficacy and Safety of Vitam 107<br />

Exercise reduces liver fat, 2183<br />

Fibrometer Nonalcoholic Fatt 2235<br />

First degree relatives of su 2203<br />

Hepatic expression of the ap 2149<br />

High prevalence of undiagnos 2231<br />

High prevalence of PNPLA3 rs 2210<br />

Hypovitaminosis D in NAFLD r 2186<br />

IL-8 and MCP1 are Independen 2169<br />

Identification of Unique Mar 2194<br />

In Female Patients with Non- 2205<br />

Increasing severity of NAFLD 2188<br />

Liver Fat and Inflammation A 2187<br />

Longitudinal changes in FIB- 239<br />

MHC Class I Related Gene A ( 2173<br />

Macrophages (CD68+ Cells) an 2192<br />

Mechanism of Action of the A 2246<br />

Mitochondrial Mutator-Phenot 2217<br />

Molecular Characterization o 2221<br />

Naltrexone/Bupropion Extende 2175<br />

Non-Alcoholic Fatty Liver Di 2225<br />

Non-Invasive Estimation of D 2206<br />

Non-alcoholic Fatty Liver Di 2204<br />

Non-alcoholic fatty liver di 2238<br />

Non-invasive Diagnosis of No 2151<br />

Non-invasive hepatic fibrosi 2191<br />

Normal body mass index as a 2251<br />

Obesity fibrosis score: A ne 2182<br />

Older Age, Increasing Body M 2159<br />

Omega-3 fatty acids improve 240<br />

Outcome of bariatric surgery 2222<br />

Plasma Metabolomic Profiles 161<br />

Potential benefit of folate 2162<br />

Predictors of improvement in 2154<br />

Prevalence and severity of n 235<br />

Prevalence of hepatic steato 2253<br />

Prevalence of non-alcoholic 2189<br />

Relationships between very l 2228<br />

Relevance of the Oxidative S 2237<br />

Reverse correlation with FXR 2218<br />

Risk factors for hepatocellu 2166<br />

Serum marker of inflammasome 2185<br />

Short Sleep Duration is asso 2155<br />

Steatohepatitis and liver fi 2227<br />

Substantial variability of t 2164<br />

Subtyping nonalcoholic steat 2244<br />

Testosterone Is Associated w 2252<br />

The dual and opposite role o 2216<br />

The effectiveness of SGLT-2 2240<br />

The impact of alcohol consum 2241<br />

Transient Elastography is Fe 2220<br />

PNPLA3 and JAZF1 variants in 2158<br />

1H-Magnetic Resonance Spectr 2152<br />

A 32-gene poor prognosis sig 2174<br />

An international, phase 2 ra 105<br />

Changes in plasma biomarkers 2161<br />

Characteristics and Clinical 2250<br />

Clinical and Metabolic Effec 236<br />

Concomitant Non-Alcoholic Fa 2214<br />

Cross-sectional and longitud 2150<br />

Cross-sectional and longitud 2201<br />

DACRAs are novel therapeutic 2178<br />

DJ-1 labeling index is a nov 2165<br />

Diacylglycerol acyltransfera 2208<br />

Diagnostic accuracy for non- 2177<br />

Disturbance of regulatory T 2167<br />

Effect of Pradigastat, a Dia 2147<br />

Extending the Ballooning Sco 157<br />

Functional inhibition of spl 2198<br />

Haptoglobin Genotype 2 Allel 160<br />

How common is chronic liver 2168<br />

Importance of reduction in s 2202<br />

Improvement in Hepatic Steat 2200<br />

Increased Immune Responses o 2157<br />

Longterm Effects of Bariatri 2170<br />

Minimally deranged serum ala 2226<br />

Multi-parametric magnetic re 2190<br />

Non-alcoholic fatty liver is 2245<br />

Overweight in late adolescen 158<br />

Plasma Cytokeratin-18 Fragme 2148<br />

Polychlorinated biphenyl exp 2233<br />

Prebiotic supplementation wi 2176<br />

Preserved hemostatic status 2249<br />

Prospective comparison of no 2211<br />

Risk factors for hepatocellu 2207<br />

Risk factors for progression 2213<br />

Serum oxidative-anti-oxidati 2236<br />

Subjects with non alcoholic 2209<br />

The Association Between Sarc 2234<br />

The Lipidomic Signature Of D 2163<br />

The Performance of Vibration 2197<br />

The association of ALT level 2224<br />

The development of hepatocel 2230<br />

The glycosylation marker M2B 2199<br />

The hepatic and extra-hepati 2145<br />

The link between intestinal 2184<br />

Upregulated absorption of di 2146<br />

Usefulness of mac-2 binding 2193<br />

Weight Loss Results in Signi 2196<br />

Zone 1 and Zone 3 Steatosis 159


1378A CATEGORY INDEX HEPATOLOGY, October, 2015<br />

QO3. Alcohol: Clinical and<br />

Experimental<br />

Alcohol modifies the composi 1311<br />

An alarmin cytokine IL-33 in 1345<br />

Antimicrobial proteins Reg3b 111<br />

Ceramide Synthase May Regula 1314<br />

Cross talk between hepatocyt 1339<br />

Differential Changes In Pro- 1309<br />

Differential microRNA expres 1332<br />

Elevated leukocyte count com 1350<br />

Ethanol Exposure Inhibits He 1310<br />

Ethanol-impairs mTOR activit 1308<br />

Fat-specific Protein 27/CIDE 110<br />

Functional role of the Lin28 1349<br />

Genetic Ablation of MitoNEET 1338<br />

Hepatocyte-specific StARD1 d 1320<br />

Hepatocytes from mice on int 112<br />

Impaired Proteasome Function 1325<br />

Low Dose Zinc Sulfate (220mg 1303<br />

Microparticles Are Markers O 1344<br />

Molecular chaperone Hsp90 re 1351<br />

Nogo-B facilitates M1 polari 1322<br />

Positive familial metabolic 1328<br />

Role of Pre-mRNA Splicing Re 1324<br />

The Alcoholic Hepatitis Hist 1337<br />

The Epidemiology and Costs o 1329<br />

Short Term Abstinenc 113<br />

Creatine Supplementa 1347<br />

Ethanol-Induced Alte 1323<br />

Acrolein, a lipid-derived al 1304<br />

Alcohol stimulates macrophag 1315<br />

Alterations of gut microbiom 114<br />

Characteristic features of m 1326<br />

Coffee Consumption Protects 1316<br />

Cyr61 protection of hepatocy 1352<br />

Early prediction of response 1307<br />

Effects of Alcohol and Absti 1343<br />

Fibroblast growth factor 21 1306<br />

Gut Bacterial Load Influence 1330<br />

Hepatocyte-specific deletion 1313<br />

Increased IL-17 production i 1318<br />

Infection in alcoholic hepat 1321<br />

Inflammatory and apoptotic m 1334<br />

Interplay between microRNA-2 1317<br />

Intestinal HIF-1α Deletion 1305<br />

LPS-TLR4 pathway mediates du 1327<br />

Macrophage Migration Inhibit 1346<br />

Microbiota protects mice aga 1319<br />

Nosocomial Infection in Seve 1342<br />

Nuclear receptor Rev-Erbα f 1336<br />

Oral pentamidine (VLX103) pr 1333<br />

PDE4 inhibition attenuates a 1312<br />

Sarcopenia and adipopenia se 1340<br />

The L-carnitine alleviate he 1348<br />

The role of proinflammatory 1341<br />

Tributyrin, a butyrate pro-d 1331<br />

Zinc deficency inactivate he 1335

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!